laboratory simulation of reservoir-induced seismicity · laboratory simulation of reservoir-induced...

221
LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY by Winnie (Wai Lai) Ying A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Civil Engineering University of Toronto © Copyright by Winnie (Wai Lai) Ying 2010

Upload: vubao

Post on 10-Aug-2018

232 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

LABORATORY SIMULATION OF

RESERVOIR-INDUCED SEISMICITY

by

Winnie (Wai Lai) Ying

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Graduate Department of Civil Engineering

University of Toronto

© Copyright by Winnie (Wai Lai) Ying 2010

Page 2: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

ii

LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY

Winnie (Wai Lai) Ying

Doctor of Philosophy

Graduate Department of Civil Engineering

University of Toronto

2010

Abstract

Pore pressure exists ubiquitously in the Earth’s subsurface and very often exhibits a

cyclic loading on pre-existing faults due to seasonal and tidal changes, as well as the

impoundment and discharge of surface reservoirs. The effect of oscillating pore pressure on

induced seismicity is not fully understood. This effect exhibits a dynamic variation in

effective stresses in space and time. The redistribution of pore pressure as a result of fluid

flow and pressure oscillations can cause spatial and temporal changes in the shear strength of

fault zones, which may result in delayed and protracted slips on pre-existing fractures.

This research uses an experimental approach to investigate the effects of oscillating pore

pressure on induced seismicity. With the aid of geophysical techniques, the spatial and

temporal distribution of seismic events was reconstructed and analysed. Triaxial experiments

were conducted on two types of sandstone, one with low permeability (Fontainebleau

sandstone) and the other with high permeability (Darley Dale sandstone). Cyclic pore

pressures were applied to the naturally-fractured samples to activate and reactivate the

existing faults. The results indicate that the mechanical properties of the sample and the

heterogeneity of the fault zone can influence the seismic response. Initial seismicity was

Page 3: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

iii

induced by applying pore pressures that exceeded the previous maximum attained during the

experiment. The reactivation of faults and foreshock sequences was found in the

Fontainebleau sandstone experiment, a finding which indicates that oscillating pore pressure

can induce seismicity for a longer period of time than a single-step increase in pore pressure.

The corresponding strain change due to cyclic pore pressure changes suggests that

progressive shearing occurred during the pore pressure cycles. This shearing progressively

damaged the existing fault through the wearing of asperities, which in turn reduced the

friction coefficient and, hence, reduced the shear strength of the fault. This ‘slow’ seismic

mechanism contributed to the prolonged period of seismicity. This study also applied a

material forecast model for the estimation of time-to-failure or peak seismicity in

reservoir-induced seismicity, which may provide some general guidelines for short-term field

case estimations.

Page 4: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

iv

Acknowledgements

First, I would like to thank Professor R. Paul Young, my thesis supervisor, for his

guidance, advice and the opportunity he granted me to work in the field of rock fracture

dynamics. Special thanks are also due to Dr. Philip Benson (my experiment and thesis

advisor while he was at the UofT as Marie-Curie Research Fellow), who provided technical

assistance, valuable comments, and discussions throughout the course of the research; Dr.

Farzine Nasseri, who provided help during the sample preparation and assembly of

equipment; and Dr. Alex Schubnel, who provided the Fontainebleau sandstone blocks. I

extend my appreciation to Professor Michael King, Dr. Rosanna Smith, Dr. Ben Thompson

and Dr. Dave Collins who provided insightful comments and discussions. In addition, I

would like to thank my external examiner, Professor Bezalel Haimson, and my examination

committee: Professors John Curran, Brent Sleep and Bernd Milkereit for their valuable

comments to improve the quality of the thesis. I also gratefully acknowledge the assistance

of Professor Mark Kortschot and Billy Cheng for their assistance with the X-ray

micro-computed tomography scanner. Technical support was also received from the

technical team of the Civil Engineering Department (Giovanni Buzzeo, Alan McClenaghan,

Olga Perebatova) and from the Geology Department, Shawn McConville, who prepared the

thin sections for optical microscopy studies. Special thanks are due to Laszlo Lombos and

Dylan Roberts who provided technical support on the geophysical imaging cell (triaxial cell)

and Dr. Will Pettitt and his team who provided technical support on the InSite software.

I also gratefully thank my past and present colleagues: Dr. Toivo Wanne, Dr. Tatyana

Katsaga, and Xueping Zhao. They gave moral support throughout the course of the research.

Furthermore, I would like to thank Professor L. S. Chan and friends for their encouragement

and support. I am also very thankful for my parents’ and my sister’s support. Without their

caring, love, and patience, this path would not be possible.

Page 5: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

v

Table of Contents

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

Acknowledgements...................................................................................................................iv

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

List of Figures........................................................................................................................ viii

List of Tables ..........................................................................................................................xiv

List of Symbols and Abbreviations .........................................................................................xv

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

1.1 Fluid-induced seismicity....................................................................................1

1.2 Literature review................................................................................................3

1.2.1 Classification of reservoir-induced seismicity.......................................... 8

1.2.2 Modelling RIS ........................................................................................ 10

1.2.3 Acoustic emissions as earthquake proxies...............................................11

1.3 Thesis objectives and overview.......................................................................12

Chapter 2 Theory .....................................................................................................................15

2.1 Fundamentals of porous media........................................................................15

2.1.1 Definition of porosity ............................................................................. 15

2.1.2 Porosity measurement............................................................................. 16

2.1.3 Permeability............................................................................................ 17

2.2 Rock fracture mechanics..................................................................................21

2.2.1 Stress relations and Coulomb failure criterion ....................................... 21

2.2.2 Pore pressure effects ............................................................................... 26

2.2.3 Rock friction ........................................................................................... 33

2.3 Seismicity patterns...........................................................................................37

2.3.1 Aftershock sequence............................................................................... 38

2.3.2 Foreshock sequence ................................................................................ 41

Chapter 3 Techniques: The Use of Rock Physics and Laboratory Tools ................................42

3.1 Sample materials and preparation....................................................................44

3.1.1 Darley Dale sandstone ............................................................................ 45

3.1.2 Fontainebleau sandstone......................................................................... 45

3.2 Equipment........................................................................................................45

3.2.1 Triaxial deformation cell and triaxial compression loading machine..... 45

3.2.2 Servo-controlled permeameter system ................................................... 48

3.2.3 Confining pressure pump........................................................................ 50

Page 6: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

vi

3.2.4 Linear variable differential transformer (LVDT).................................... 51

3.2.5 Data acquisition (DAQ) hardware .......................................................... 51

3.3 Geophysical techniques ...................................................................................53

3.3.1 Ultrasonic-wave velocity survey and AE techniques ............................. 53

3.3.2 AE methods ............................................................................................ 62

3.4 Source of experimental error ...........................................................................72

3.5 X-ray micro-computed tomography ................................................................73

3.6 Optical microscopy..........................................................................................73

Chapter 4 Laboratory Simulation of RIS due to Oscillating Pore Pressures...........................75

4.1 Behaviour under compressive load tests..........................................................75

4.2 Behaviour under pore pressure reduction and subsequent increase ................79

4.3 Behaviour under cyclic pore pressure..............................................................80

4.4 Conclusions from Initial Testing......................................................................84

4.5 Modified Cyclic Pore Pressure Experiments...................................................86

4.6 Experiment setup and procedures ....................................................................91

4.7 Velocity survey ................................................................................................92

4.8 Results .............................................................................................................95

4.8.1 Fontainebleau sandstone intact sample experiment (F5 and F8)............ 95

4.8.2 Fontainebleau sandstone control experiment (F7)................................ 102

4.8.3 Fontainebleau sandstone saw-cut experiment (F6)............................... 105

4.8.4 Darley Dale sandstone experiment (DDS7) ......................................... 108

Chapter 5 Analysis and Discussion................................................................................112

5.1 Experiment F5 ...............................................................................................112

5.1.1 b-value analysis .....................................................................................118

5.1.2 AE source mechanism ...........................................................................119

5.2 Experiment F8 ...............................................................................................120

5.3 Control experiment on Fontainebleau sandstone...........................................122

5.4 Darley Dale sandstone experiment ................................................................126

5.4.1 b-values................................................................................................. 128

5.4.2 AE source mechanism .......................................................................... 130

5.5 Comparison between Fontainebleau sandstone and Darley Dale sandstone

experiments....................................................................................................132

5.6 Saw-cut experiment .......................................................................................133

5.6.1 Migration trends.................................................................................... 135

5.7 Post-experimental analysis ............................................................................135

5.7.1 X-ray micro-computed tomography analysis ....................................... 136

Page 7: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

vii

5.7.2 Optical microscopy............................................................................... 138

5.8 Conclusions ...................................................................................................144

Chapter 6 Forecasts of Reservoir-induced Seismicity ...................................................148

6.1 Material failure forecast method....................................................................149

6.2 Forecast of main slip......................................................................................150

6.3 Application to the Koyna reservoir (Protracted seismicity forecast).............154

6.4 Application to the Monticello reservoir (Initial peak seismicity forecast) ....156

6.5 Conclusions ...................................................................................................158

Chapter 7 Conclusions and Recommendations ..............................................................160

7.1 Conclusions ...................................................................................................160

7.2 Recommendations..........................................................................................163

References..............................................................................................................................165

APPENDIX I – Calibration Charts........................................................................................181

APPENDIX II – Sensor location and sensor file...................................................................191

APPENDIX III – Details of the F2 experiment.....................................................................194

APPENDIX IV – Response to Different Frequencies of Oscillation and Rate of Increase in

Pore Pressure………………………………………………………….195

APPENDIX V – Glossary .....................................................................................................196

APPENDIX VI – Information of Paper Published in GRL...................................................198

Page 8: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

viii

List of Figures

Figure 1.1 Worldwide distribution of reservoir-induced seismicity, with M ≥ 4.0………6

Figure 2.1 Schematic diagram of Darcy’s law…………….…………………………….18

Figure 2.2 Schematic figure of stress field…………..…………………………………..22

Figure 2.3 Mohr diagram for normal shear stresses produced by the principal stresses..22

Figure 2.4 Mohr stress circles for a series of tests showing failure according to Coulomb

failure criteria………..……..………………………………………………...23

Figure 2.5 Mohr circle representing the elastic effect of reservoir loading on the strength

of rock underneath the reservoir..……..……………………………………..24

Figure 2.6 Schematic figure of the change in stability of a fault plane relative to the

position of the reservoir..…………………………………………………….25

Figure 2.7 Mohr circle representing the fluid pressure effect on the strength of rock.…26

Figure 2.8 Pore pressure and shear stress at a faulted surface………………………….28

Figure 2.9 Showing the values of equation 2.17 for different values of z and t and D =

106 m

2/day..………………………………………………………………….31

Figure 2.10 Showing the values of equation 2.17 for different values of D and t, and z = 1

km and 125 mm, respectively…..……………………………………………31

Figure 2.11 Block-slider model demonstrating initiation of frictional instability……......35

Figure 2.12 Force displacement diagram showing a hypothetical case…………...….…..35

Figure 2.13 Shear stress plotted as a function of normal stress at the initial friction for a

variety of rock types………………………………………………………....36

Figure 2.14 Mogi’s classification of foreshock-aftershock patterns and their relationship to

the structures of materials and applied stresses…..…..……………………...38

Figure 3.1 Schematic diagram of the flow line from the experimental setup to data

acquisition to data processing to output and analysis..………………………44

Figure 3.2 Compression loading machine and triaxial deformation cell…………….….46

Figure 3.3 Cross section of the triaxial deformation apparatus showing the top and

bottom axial platens………………………………………………………….47

Page 9: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

ix

Figure 3.4 Rubber jacket holding lateral transducers and stacks in position……….…...48

Figure 3.5 North-East-Down (NED) coordinate system………..……..………………...48

Figure 3.6 Schematic diagram of the experiment setup…………..……………………..50

Figure 3.7 NI6255 data acquisition board……..………………………………………...52

Figure 3.8 Flow chart showing the LabVIEW data acquisition program………….……53

Figure 3.9 PAD pre-amplifiers used for amplifying signals by 60 dB…………….……57

Figure 3.10 Continuous waveform of a Darley Dale sandstone experiment and a discrete

acoustic emission waveform extracted from continuous waveform data at

different zoom levels…………………………………………………...……60

Figure 3.11 Flow chart showing the various AE data acquisition units for recording

passive and active AE data……..…………….……………………………...62

Figure 3.12 Showing typical P-wave arrivals received from each channel………...…….64

Figure 3.13 Examples of focal sphere and equivalent forces……………………………..67

Figure 3.14 The nine elementary force couples…………………………………………..68

Figure 3.15 Vector force representation of some source models…………………………70

Figure 4.1 Stress-strain curves of the Darley Dale sandstone samples under 10, 20, and

40 MPa confining pressures…...……………………………………………..77

Figure 4.2 Typical stress-strain curve of Fontainebleau sandstone with 20 MPa confining

pressure………………………………………………………………………78

Figure 4.3 Experiment F2 – Pore pressure cycles introduced after the rupture of the

specimen……………………………………………………………………..82

Figure 4.4 Activation of fault begins when the pore pressure reaches ~4.3 MPa (~87% of

the previous maximum)………………..……………………….……………82

Figure 4.5 a) The fractured F2 Fontainebleau sandstone sample; b) spatially and

temporally scattered seismic events induced by pore pressure cycles occurred

along the pre-existing faults……………………………………………..…..83

Figure 4.6 Showing the reference distance for the measurement of hypocenter locations

during initial tests…..………………………..………………………………84

Page 10: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

x

Figure 4.7 The number of events located at different ranges of hypocenter. The trend

indicated during the cyclic pore pressure period is similar to that during the

pre-peak period………………..…...………………………………..……….84

Figure 4.8 The three groups of initial testing and the output used for the design of cyclic

pore pressure experiments..………………………………………...………..86

Figure 4.9 The effect of different pore pressure oscillating frequencies…….….………89

Figure 4.10 An example of a Fontainebleau sandstone experiment in which the phase shift

could not be established…..……….………….……………………………..90

Figure 4.11 Typical velocity-time curve of Fontainebleau sandstone oscillating pore

pressure experiment………..…………………………...………………..…..94

Figure 4.12 Typical velocity-time curve of the Darley Dale sandstone oscillating pore

pressure experiment.……...……………………………….…………………94

Figure 4.13 Stress-time curve of the F5 experiment under the constant confining pressure

of 20 MPa.…………………………………………….………………..…..95

Figure 4.14 Stress-time plot of the F8 experiment under the constant confining pressure of

20 MPa...……………………………………………………………….….....96

Figure 4.15 Strain-time plot of the F5 experiment……………………………..…………96

Figure 4.16 Strain-time plot of the F8 experiment……………..…………………………97

Figure 4.17 Seismic rate-time plot of the F5 experiment…………………..…………….97

Figure 4.18 Seismic rate-time plot of the F8 experiment……………………………..….98

Figure 4.19 Upstream and downstream pore pressures during the post-peak stage of the F5

experiment…...………………………………………………………………99

Figure 4.20 Upstream and downstream pore pressures during the post-peak stage of the F8

experiment…………………………………………………………………..99

Figure 4.21 AEs occurred during the 5 cycles with sinusoidal pore pressure oscillated

between 2.5 and 18 MPa at 2-minute periods of the F5, mimicking a main

shock-aftershock sequence…….……………..………………..…………...100

Figure 4.22 AEs occurred during the 16 cycles with sinusoidal pore pressure oscillated

between 2.5 and 17 MPa at 2-minute periods of the F5, mimicking a

foreshock-main shock-aftershock sequence……….………………….…....101

Figure 4.23 AE locations of the F5 Fontainebleau sandstone experiment………………102

Page 11: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xi

Figure 4.24 Stress-time plot of the F7 experiment under the constant confining pressure of

20 MPa...………………………………………………………..…….…….103

Figure 4.25 Strain-time plot of the F7 experiment………………………..……………103

Figure 4.26 Upstream and downstream pore pressures during the post-peak stage of the F7

experiment……………………………………………………………….....104

Figure 4.27 Seismic rate-time plot of the F7 experiment…..….......................................104

Figure 4.28 Stress-time plot of the F6 experiment under the constant confining pressure of

5 MPa….………………………………………………………………..…105

Figure 4.29 Strain-time plot of the F6 experiment………………………………………106

Figure 4.30 Upstream and downstream pore pressures during the post-peak stage of the F6

experiment…....…………………………………………………………….106

Figure 4.31 Seismic rate-time plot of the F6 experiment………………………………..107

Figure 4.32 AE source location showing the events that occurred during each set of pore

pressure cycles………..…..…………..…………………………………….107

Figure 4.33 Stress-time plot of the DDS7 experiment under the constant confining

pressure of 20 MPa.…………………………………………..……………109

Figure 4.34 Strain-time plot of the DDS7 experiment……………………..……………109

Figure 4.35 Upstream and downstream pore pressures during the post-peak stage of the

DDS7 experiment….……..……..………………………………………….110

Figure 4.36 Seismic rate-time plot of the DDS7 experiment…………………………....110

Figure 4.37 AE source location showing the events that occurred during each set of pore

pressure cycles…………………………………….………………………..111

Figure 5.1 The seismic rate, pore pressure cycles, and axial strain change of the F5

experiment………………………………………………………………….114

Figure 5.2 Phase shift and average pore pressure at each cycle for the F5 experiment..115

Figure 5.3 Showing the reference distance for measurement of hypocenter locations...117

Figure 5.4 Population distribution of the seismic events occurred during pore pressure

oscillations……………………………………………………………….…117

Figure 5.5 Fontainebleau sandstone experiment: a) b-value analysis for the formation of

fault; b) b-value analysis for the aftershock and foreshock sequences......…119

Page 12: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xii

Figure 5.6 Focal mechanism solutions indicating the failure mechanisms during

oscillating pore pressure…...……………………………………………….120

Figure 5.7 The seismic rate, pore pressure cycles, and axial strain change of the F8

experiment………………………………………………………………… 122

Figure 5.8 The control experiment indicates pore pressure steps at 18 MPa for 10

minutes and 17 MPa for 32 minutes.……………………………………….124

Figure 5.9 Schematic diagram indicating the reduction of shear strength with cycles a to

d………………………………………………………………………….…126

Figure 5.10 The seismic rate and axial strain change of the Darley Dale sandstone

experiment……………………………………………………….…………127

Figure 5.11 Showing seismic rate, pore pressure cycles, and axial strain change of the

Darley Dale sandstone experiment, with pore pressure oscillated between a 18

and 2.5 MPa simulated aftershock sequence…………………………..…...128

Figure 5.12 b-value analysis of the Darley Dale sandstone experiment…………...........129

Figure 5.13 The distribution of seismic events that occurred during the oscillation of pore

pressure………………………………………………………………….….130

Figure 5.14 Focal mechanism solutions of the Darley Dale sandstone experiment,

indicating that the dominant failure mechanism during the cyclic pore

pressure stage is shear, with the corresponding double couple percentage...131

Figure 5.15 The F6 experiment indicated stable sliding along the saw-cut at a later stage,

when the average pore pressure was about 3.5 MPa…………………….…134

Figure 5.16 Typical X-ray micro-CT images along the vertical axis……...……..…...…137

Figure 5.17 Normalised crack area is calculated as crack area/minimum crack area of the

plots…..………………………………...……………………………….…..138

Figure 5.18 Optical microscopic views of the Fontainebleau sandstone…….………….141

Figure 5.19 Thin section of the Darley Dale sandstone tested sample……….…………142

Figure 5.20 Optical microscopy of the Darley Dale sandstone with sheared material in the

fractured zone……......…..…………………………………………………143

Figure 6.1 Fitting Ac equals 0.03 and γ equals 2 into failure forecast model………….151

Page 13: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xiii

Figure 6.2 Application of failure forecast model to the experimental foreshock sequence

obtained from the F5 experiment...………………………………………...152

Figure 6.3 Application of failure forecast model to short-term forecast of main slip #2 in

the F5 experiment……..…………………...…………………………….…153

Figure 6.4 Long-term forecast of the Koyna RIS……………….……………………..155

Figure 6.5 Short-term forecast of the Monticello RIS……………...………………….157

Figure A1 Calibration of the two LVDTs: A and B......….…………………..…………182

Figure A2 Calibration of the pore pressure transducers: Pa and Pb……………..…….183

Figure A3 Calibration of the permeameter cylinder volume: Pa and Pb………………184

Figure A4 Calibration of the lateral strain measurement device.……………………...185

Figure A5 Calibration of the axial load transducers, Z1 and Z2……………………..186

Figure A6 Calibration of velocity models during different periods of the Darley Dale

sandstone experiment…………………………………...………………..…188

Figure A7 Calibration of velocity models during different periods of the Fontainebleau

sandstone experiment…………………………………...………………..…189

Figure A8 Locating a synthetic acoustic emission to calibrate source location………190

Figure A9 Locations of the lateral sensors……….………..….………………………..191

Figure A10 Locations of the platen sensors……….………..….………………………..192

Page 14: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xiv

List of Tables

Table 3.1 Calibration for ultrasonic wave arrival time measurements…………………55

Table 3.2 Summary of experimental error sources and accuracy………..…….……….72

Table 4.1 List of experiments…………………………………………………………..87

Table 5.1 b-values during different periods of the F5 experiment……………………118

Table 5.2 Corresponding vertical movements implied by axial strain measurements..125

Table 5.3 b-values during different periods of the DDS7 experiment….…….……….130

Table 5.4 Comparison of the results of the two cyclic pore pressure experiments…...133

Table 6.1 Sensitivity analysis for the ‘long-term’ forecast of the F5 experimental

data…………………………………………………………………………151

Table 6.2 Sensitivity analysis for the forecast of peak seismicity of the RIS at the

Monticello reservoir………………………………………………………...158

Table A1 Summary of calibration parameters………………………………………...186

Table A2 Absolute errors in source location…………………………………………190

Table A3 Sensor file……………...……………………...……………………………193

Table A4 F2 experiment: Peak AE hits and the time delayed in response to the peak

pore pressure magnitude……………………………..…………………..…194

Table A5 AE response corresponding to different rate of increase in pore pressure…195

Page 15: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xv

List of Symbols and Abbreviations

The symbols and abbreviations are listed in alphabetical order, first in the Latin, then the

Greek alphabets. The point of first appearance is given in parenthesis, which refers to an

equation, figure or section. In Figure 3.16, the symbols P (pressure), B (null), T (tension) are

used for different meanings from the ones listed below. They represent minimum,

intermediate and maximum eigenvalues respectively; however those meanings are clear

within the context used.

Symbol/Abbreviation Description

A Anisotropy factor [Equation 4.1]

Ac Constant [Equation 7.1]

a Constant [Equation 2.19]

[ak] Vectoe containing k number of amplitudes [Equation 3.5]

AE Acoustic emission [Section 1.2.3]

B Skempton’s coefficient [Equation 2.13]

b The slope of the line measured from the linear descending portion

of the frequency-magnitude scaling relation [Equation 2.19]

CLVD Compensated linear vector dipole [Figure 3.16]

CT Micro-computed tomography [Section 3.5]

c The time offset parameter [Equation 2.20]

D Hydraulic diffusivity [Equation 2.11]

Ds Sample distance or sample diameter [Equation 3.1]

DAQ Data acquisition [Section 3.2.5]

DC Double couple [Section 3.3.2]

d Diameter of the sample [Equation 2.4]

dm Distance between receiver and the AE location [Equation 3.6]

dT Diameter of transducer [Equation 3.4]

F Applied force [Equation 2.11]

FN Normal force [Equation 2.16]

[Gki,j] Green’s function indicating the propagation effects [Equation 3.5]

Page 16: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xvi

g Gravitational constant [Equation 2.4]

H(t) Heaviside unit step function [Equation 2.13]

K Spring stiffness [Figure 2.16]

k Permeability [Equation 2.2]

L Length of the near field [Equation 3.4]

LVD Linear vector dipole [Figure 3.16]

LVDT Linear variable differential transformer [Section 3.2.4]

M Magnitude of seismic event [Equation 2.19]

Mamax

The largest magnitude of the main shock [Equation 2.21]

Mm Magnitude of the main shock [Equation 2.21]

[M] Vector containing moment tensor components [Equation 3.5]

m Constant [Equation 2.22]

mi Eigenvalues of moment tensor [Section 3.3.2]

mL Location magnitude of an AE between a set of events [Equation 3.6]

N Number of seismic events [Figure 2.14]

P Pore pressure [Figure 2.1]

P’ Pore pressure at the other end of the sample [Equation 2.1]

p The exponent that modifies the rate of change in seismic

frequency [Equation 2.20]

q Darcy velocity [Equation 2.1]

RIS Reservoir-induced seismicity [Chapter1]

S Cross-sectional area of sample [Figure 2.1]

S1, S2 Shear waves [Section 3.2.1]

Sn Number of sensors [Equation 3.6]

T The time difference between sending a pulse and receiving signal

[Equation 3.1]

To The travel time between pulsing and receiving a signal with no

sample present between receiver and pulser [Equation 3.1]

t Time period [Equation 2.10]

tc Time of the main shock occurrence [Section 2.3.2]

u Slip [Figure 2.12]

Page 17: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xvii

V Elastic wave velocity [Equation 3.1]

Vr Total volume/bulk sample volume [Equation 2.1]

Vf Volume of flow rate [Equation 2.4]

Vmax Maximum P-wave velocity [Equation 3.3]

Vmin Minimum P-wave velocity [Equation 3.3]

Vpore Pore volume [Equation 2.1]

Vsolid Volume of solid within the rock [Equation 2.1]

VI Virtual instrument [Chapter 3]

vu Poisson’s ratio [Equation 2.13]

W Amplitude [Equation 2.20]

Wi Waveform amplitude [Equation 3.7]

Win Input amplitude [Equation 3.2]

Wout Output amplitude [Equation 3.2]

X Sample length [Figure 2.1]

x Number of data points in waveform [Equation 3.7]

z Depth [Equation 2.10]

α Angle between the σ1 plane and the fault plane [Equation 2.5]

β Bulk compressibility of fluid-filled rock [Equation 2.11]

∆τ The change in shear stress [Equation 2.9]

∆S Incremental shear strength [Equation 2.9]

∆σn The change in compressive normal stress [Equation 2.9]

∆P The change in pore pressure [Equation 2.9]

Φ Porosity [Equation 2.1]

φ Angle of internal friction [Equation 2.7]

γ An exponent that measures the degree of non-linearity [Equation

7.1]

η Fluid viscosity [Equation 2.2]

λ Wavelength [Equation 3.4]

µ, tanφ Coefficient of internal friction angle [Equation 2.7]

θ Ray-path angle [Equation 3.3]

Page 18: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

xviii

ρ Density of the fluid [Equation 2.4]

σ Stress [Equation 2.8]

σ' Effective stress [Equation 2.8]

σ1 Maximum principal stress [Equation 2.6]

σ3 Minimum principal stress [Equation 2.6]

σn Normal stress [Equation 2.6]

τ Shear stress [Equation 2.5]

τ0 Cohesion [Equation 2.7]

Ω Precursory strain [Equation 7.1]

ω Angular frequency of water level changes [Equation 2.14]

Page 19: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

1

Chapter 1 Introduction

1.1 Fluid-induced seismicity

Pore fluids exists ubiquitously in the Earth’s subsurface and, in many cases, exhibit a

cyclical loading on pre-existing faults due to seasonal and tidal changes. It has been

speculated that non-volcanic tremor and low-frequency earthquake swarms are generated by

high fluid pressure, which enabled shear slip at plate interface asperities (Shelly et al., 2007).

These types of slow earthquakes exhibit episodic tremors and slips over a wide range of time

scales and with slow rupture propagation (Obara and Hirose; 2006; Obara, 2002; Linde and

Sacks, 2002). Studies have indicated that these earthquakes have low frequencies ranging

from 0.5 to 5 Hz (Obara, 2002; Rogers and Dragert, 2007). Ariyoshi et al. (2009) suggest that

interaction between asperities may cause the low-frequency nature of the earthquakes. In

addition, human activities, such as the impoundment and discharge of surface reservoirs, the

creation of underground reservoirs and the subsequent extraction of geothermal energy from

enhanced geothermal systems (Majer et al., 2007) and, injections in deep oil and gas wells

can generate fluctuating fluid pressures on fault zones (Zhao et al., 1995) and induce

seismicity. The study of fluid-induced seismicity provides a good context within which to

understand the physics of fundamental processes such as stress rotation (Faulkner et al., 2006;

Fitzenz and Miller, 2004), high pressure pulse induced aftershocks (Miller, 2004), and the

generation of seismic swarms (Yamashita, 1999; Kilburn, 2003; Benson et al., 2008).

Although the importance of pore fluids in induced seismicity is well known (Miller, 2004;

Richardson and Marone, 2008), the precise relationship between pore fluid pressure and the

mechanics of faulting in shallow crustal conditions is not fully understood, in particular, the

mechanism of fluid-induced protracted seismicity. Recent work has been focused on

Page 20: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

2

fluid-induced micro-seismicity (Benson et al., 2008; Miyazawa et al., 2008) and aseismic

fault movement (Rubenstein et al., 2007; Rogers and Dragert, 2007) in order to establish a

relationship between fault nucleation and slip, and the determination of observable foreshock

sequences (Lin, 2009; Umino et al., 2002).

The study of RIS provides an exceptionally good opportunity to understand the

mechanics of natural earthquakes and the hydraulic properties of the crust (Bell and Nur,

1978; Talwani and Acree, 1984/85; Roeloffs, 1988; Talwani, 1997). Furthermore, RIS can be

a step forward in earthquake forecast, as it provides identifiable foreshock–aftershock

sequences for the verification of various forecasting models (Gupta, 2002).

Since the causal association of seismicity with the reservoir impoundment of Lake

Mead, formed by the Hoover Dam on the Colorado River in the late 1930’s,

reservoir-induced seismicity (RIS) has been reported, due not only to the impoundment of

artificial surface reservoirs, but also to the injection of fluids into the ground (Evans, 1966)

and seasonal water level changes (Saar and Manga, 2003; Roeloffs, 1988). Since then, RIS

has drawn the attention of scientists, as it can cause property damage as well as the loss of

human life. The focus of this thesis is the effect of oscillating pore pressure of artificial

surface water on reservoir-induced seismicity.

The effect of oscillating pore pressure on fractured surfaces is highly complex, with

effects ranging from a transient reduction in grain-contact stress (Iverson and LaHusen, 1989)

to the localised nucleation of the pre-existing fault, resulting in periodic movement on a ‘slip

Page 21: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

3

patch’ (Richardson and Marone, 2008). Although the shallow crustal environment is under

less overall stress than at deep subduction zone, the damage can still be considerable due to

stress release at shallow focus. Several such earthquakes have occurred due to the

impoundment of large dams, such as the M 6.3 earthquake at the Koyna reservoir (India)

(Gupta, 2002 and 2005) and the numerous swarms at the Aswan reservoir (Egypt) (Selim,

2002).

1.2 Literature review

As the cases of artificial reservoir-induced seismicity accumulated, studies (Snow, 1972;

Bell and Nur, 1978; Talwani and Acree, 1984/85; Simpson, 1976, 1986; Simpson et al., 1988;

Roeloffs, 1988; Rajendran and Talwani, 1992) have been carried out to identify the effect of

reservoir loading on the existing stress field. Very often, the induced seismicity is associated

with the initial impoundment of the reservoir, which changes the effective stress conditions

of the fault zone due to the increase in pore pressure. However, in some cases, RIS has been

observed several years after the initial impoundment, while others have lasted for several

decades after the initial increase in pore pressure (Gupta, 1985; Simpson, 1976). For instance,

the RIS at the Koyna reservoir in India has continued since 1960s (Rastogi, 2003; Rao and

Singh, 2008). Four years after the initial impoundment of the reservoir, a devastating M 6.3

earthquake occurred in 1967 (Gupta, 2002). The earthquake caused over 200 deaths and

1,500 injuries (Gupta, 1992). There have been over 100,000 earthquakes of M ≥ 0 reported in

Koyna since 1963, of which over 150 were M ≥ 4 and over 17 were M ≥ 5 (Rastogi, 2003).

Protracted seismicity has prompted the need for a more rigorous understanding of the

mechanism of RIS. Some research, such as Roeloffs (1988), has suggested that cyclic

Page 22: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

4

variation in pore pressure may induce protracted seismicity, while others have attributed the

protracted phenomenon to the pore diffusion effect (Talwani and Acree, 1984/85; Simpson et

al., 1988; Talwani, 1997; do Nascimento, 2002).

There are four major factors that control RIS (Talwani, 1997): i) Reservoir water depth,

ii) Geological and tectonic settings of the area, iii) Availability of fractures in the substrata

and, iv) Hydromechanical properties of the underlying rocks. Among these factors, the first

three are useful considerations for future reservoir sites. However, for existing artificial

reservoirs with records of continued seismicity, only the last factor can be controlled for

mitigation measures. Stuart-Alexander and Mark (1976) studied the influence of reservoir

water depths in RIS. They found that as the water depth increases, the percentage of RIS

cases increases rapidly. Baecher and Keeney (1982) also found significant correlation

between RIS and reservoir depth, while the correlation of RIS with in situ stress and geology

was less significant.

The influence of reservoir water loading on induced seismicity can be subdivided into

three main effects (Bell and Nur, 1978):

i) Elastic stress increase due to filling of reservoir.

ii) The increase in pore fluid pressure in the saturated rocks, influenced by fluid diffusion

and the compaction of the water-saturated rock due to the weight of the reservoir.

iii) Variation in pore pressure, which varies with mechanical parameters, geology of the

substratum, and the frequency and amplitude of pore pressure fluctuation (inferred by

Roeloffs, 1988).

Page 23: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

5

Some studies classify RIS as ‘induced’ or ‘triggered’ according to the stress conditions

(McGarr and Simpson, 1997). ‘Induced’ seismicity involves a substantial change in crustal

stress or pore pressure with respect to its ambient state. This usually relates to the initial

reservoir impoundment or substantial water recharge/discharge. If the crust is sufficiently

close to a failure state due to natural tectonic processes (and only a small change in stress or

pore pressure is required to induce seismicity) this is referred to as ‘triggered’ seismicity. In

other words, the seismicity would have occurred due to the natural tectonic settings and the

impoundment of the reservoir caused it to happen earlier. However, discriminating between

the two types of seismicity can be difficult because it is impossible to prove that earthquakes

would have occurred without reservoir impoundment (Talwani, 2000). Therefore, in this

thesis ‘triggered’ and ‘induced’ seismicity are considered to be the same.

Up to 2000, RIS has been reported for 95 artificial reservoir sites (Figure 1.1). These

sites can be grouped according to the maximum seismic magnitude (Gupta 2002):

i) 4 sites with M ≥ 6, including the Xinfengjiang reservoir in China, the Kariba

reservoir in Zambia-Zimbabwe, the Koyna reservoir in India, and the Kremasta

reservoir in Greece.

ii) 10 sites with M 5 – 5.9.

iii) 28 sites with M 4.0 – 4.9.

iv) 53 sites with M < 4.0.

It should be noted that there are more sites of induced seismicity which have not been

recognized because of the lack of proper seismic surveillance, particularly in third world

countries (Gupta, 1992).

Page 24: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

6

Figure 1.1 Worldwide distribution of reservoir-induced seismicity, with M ≥ 4.0 (Gupta, 1992 & 2005).

As the tectonic stress in the Earth’s crust at some locations is often sufficiently close to

a critical stress, a small perturbation in the in situ stress field due to pore pressure variation at

critical locations can trigger seismicity (Talwani and Acree, 1984/85; Shapiro et al., 2006;

Roeloffs et al., 1979). Therefore, RIS sites very often coincide with these critical locations.

King et al. (1994) calculated the Coulomb stress change due to the main shock of the

Landers earthquake, which occurred on the 28th

June, 1992. They found that a stress increase

of less than 0.05 MPa could trigger earthquakes, which suggests that the stress conditions in

the areas must be very close to failure. Similarly, Grasso and Sornette (1998) analysed

induced seismicity cases and reported that both pore pressure change and mass transfer

leading to incremental deviatoric stresses of less than 1 MPa were sufficient to trigger

seismic events. Furthermore, Shapiro et al. (2006) stated that fluid-induced seismicity can be

triggered by pressure perturbations as low as 1 – 100 kPa at the hypocenters.

Page 25: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

7

In some countries, the occurrence of RIS has resulted in the major modification of civil

and engineering projects. For instance, the injection of waste fluid into the crust at the Rocky

Mountain Arsenal was discontinued due to induced seismicity (Evans, 1966). In addition, the

construction of the Auburn Dam in California was terminated in view of the potential

hazards (Allen, 1978), and later constructed with a modified design. In other countries, the

construction of large artificial water reservoirs (≥ 100 m high) has thrived for decades. This

is driven by the beneficial effects of large reservoirs, such as the generation of hydroelectric

power, flood control, irrigation and human consumption, etc. This rapid development of

large reservoirs manifests particularly in the developing countries. For instance, the Three

Gorges Dam on the Yangtze River in China was completed in 2008 and is the world’s largest

artificial reservoir, with a capacity of 39.3 km3 and a water level of 175 m. There are

hundreds of other artificial reservoirs under construction in China as of 2008. Many of these

reservoirs are large-scale and are located in seismogenic zones. Similarly, in other

developing countries such as India, the increasing demand for hydroelectric power has

sustained the increasing number of large artificial reservoirs.

After the May 2008 M 7.9 earthquake in Sichuan, China, there was debate about

whether the earthquake was a reservoir-induced (by the nearby Zipingpu reservoir) or natural.

After the completion of the initial impoundment of the Zipingpu reservoir in December 2006,

the water level was increased by 120 m. A week before the great earthquake, the reservoir

water level was rapidly reduced by ~58 m (Wang, 2008), which might have caused stress

perturbation. According to case histories, such as the Aswan reservoir in Egypt, RIS has been

reported during the reduction of surface water level. Furthermore, the Zipingpu reservoir is

Page 26: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

8

located 20 km from the epicentre of the M 7.9 earthquake (Ge et al., 2009; Kerr and Stone,

2009). Whether the M 7.9 earthquake was reservoir-induced remains questionable; however,

in view of the increasing number of large reservoirs, their potential hazards, and protracted

effects, it is essential to understand the mechanism of RIS and perhaps develop some reliable

RIS forecast models in order to mitigate or prevent induced earthquakes.

1.2.1 Classification of reservoir-induced seismicity

There are three main classifications of RIS, all of which are divided into temporal

categories. They include:

i) Rapid and delayed seismic responses. Rapid seismicity response follows immediately

after the initial loading of the reservoir or after a rapid change in reservoir water level.

According to Simpson et al. (1988), rapid response consists primarily of low magnitude,

swarm-like activity and is confined to the immediate reservoir area. They also suggest that

this type of RIS is caused by changes in elastic stresses or pore pressure change coupled to

the elastic stress and that pore pressure diffusion is not a major factor for inducing rapid

seismicity. Classic examples of rapid responses include the Nurek and Kariba reservoirs

(Gupta, 2002). Simpson et al. (1988) associate delayed responses with relatively larger

earthquakes and suggest that seismicity might extend significantly beyond the confines of the

reservoir. They suggest that diffusion of pore pressure is the mechanism responsible for these

spatial and temporal effects of RIS. Depending on the permeability and the fracture network

in the rock, it may take months or years for the pore pressure effect to spread into the crust.

When the pore pressure pulse finally reaches a zone of microcracks, it may force water into

the cracks and reduce the normal stress that holds the strained faults, consequently triggering

Page 27: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

9

delayed seismicity (Rastogi, 2003). Classic examples of this category are the Aswan and

Koyna reservoirs. The delayed response occurred seventeen years after the initial

impoundment at the Aswan reservoir (Simpson et al., 1988; Selim et al., 2002) and four years

after the initial impoundment at the Koyna reservoir (Talwani, 2000; Gupta, 2002; Gupta,

2005).

ii) Initial and protracted seismicity. Initial seismicity is associated with initial reservoir

impoundment or a large water level change. This applies to seismicity associated with water

level increases above the previous maximum attained. It results from the almost

instantaneous effect of loading (or unloading), as well as the delayed effect of pore pressure

diffusion (Talwani, 1997). The delay between the start of impoundment and the increase in

frequency and magnitude of seismicity varies from months to years and is associated with the

reservoir characteristics, local geology, and mechanical conditions. According to this initial

seismicity definition, both the rapid and delayed responses by Simpson et al. (1998) are an

integral component of initial seismicity. The initial increase in the frequency and magnitude

of earthquakes will reduce progressively, indicating the cessation of the coupled poroelastic

response to the impoundment (Talwani, 1997). Protracted seismicity occurs after the effect of

initial filling has diminished. It is often associated with the frequency and amplitude of water

level changes (Roeloffs, 1988), particularly with lower frequencies (longer periods). This

seismicity is observed both beneath the deepest part of the reservoir and in the surrounding

areas. This seismicity can persist for many years without decrease in frequency and

magnitude (Talwani, 1997). Classic examples of protracted RIS are the Koyna reservoir in

India, the Aswan reservoir in Egypt, and the Xinfengjiang reservoir (also known as the

Page 28: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

10

Hsinfengkiang reservoir in some references) in China.

iii) The third RIS classification is an integration of the previous two classifications. It is

divided into rapid response, delayed response, and continued seismicity by Gupta (2002), in

which the term ‘protracted seismicity’ is replaced by ‘continued seismicity’.

Although the classification by Gupta (2002) is the most descriptive of RIS, the one

suggested by Talwani (1997), i.e., initial seismicity and protracted seismicity, is used in this

thesis. This is because this research puts more emphasis on the protracted seismic effects of

pore pressure oscillations and, to a lesser extent, on the differences between the two initial

seismic responses: rapid and delayed.

1.2.2 Modelling RIS

Many analytical and numerical models have been developed to explain the phenomena

of RIS. Roeloffs (1988) classified the effects of RIS into three poroelastic approximations:

coupled effect (in which elastic stresses influence pore pressure and vice-versa), uncoupled

effect (in which the elastic stresses and pore pressure are independent), and decoupled effect

(in which the elastic stresses influence pore pressure, but not vice-versa). She concluded that

two-dimensional (2-D) uncoupled and coupled steady-state pore pressure solutions were

close for a large range of medium properties, while decoupled and coupled pressure solutions

were quite close for all medium properties. She suggests that the cyclic variation of water

level may induce seismic events. Other researchers such as Bell and Nur (1978), Simpson

and Narasimhan (1992), and Lee and Wolf (1998) have proposed different 2-D models to

Page 29: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

11

explain the cause of initial and ongoing seismicity in RIS, while Talwani (1997) uses a pore

pressure diffusion model to explain this phenomenon. Kalpna (2000) developed a

three-dimensional (3-D) model using the formulation suggested by Rice and Cleary (1976)

and assuming decoupled responses in RIS. Furthermore, do Nascimento (2002) and do

Nascimento et al., (2004, 2005a & 2005b) present a 3-D numerical model of the field case of

the Açu reservoir, Brazil. They incorporated hydrogeological aspects in the model and

simulated the spatial and temporal seismicity of the region due to the pore diffusion effect.

These hydrogeological factors included the heterogeneous hydraulic properties in the fault,

hydraulic conductivity variation with depth, specific storage coefficients, hydraulic

diffusivity and transmissivity. However, there is a severe lack of laboratory studies to

investigate the mechanism of protracted seismicity of RIS. One laboratory study of the effect

of cyclic pore pressure loading on saw-cut sandstone indicates that pore pressure oscillations

can induce stick-slip failures (Roeloffs et al., 1979).

1.2.3 Acoustic emissions as earthquake proxies

Many studies have shown that laboratory rock deformation experiments generate

behaviours similar to natural earthquakes and have used acoustic emissions as earthquake

proxies. An acoustic emission (AE) is a transient stress wave caused by the sudden release of

the impulsive strain energy of a material (Lockner, 1993), which travels as spherical

wavefronts in the material under stress. Kendall and Tabor (1971) suggest the use of acoustic

methods for monitoring the mechanical properties of rough interfaces. When a rock sample is

subject to deviatoric stress, AE is generated due to the formation of micro-fractures, the

opening or closing of fractures, or the shearing and sliding of pre-existing fractures.

Page 30: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

12

Brace and Byerlee (1966) and Dieterich (1979) illustrated that laboratory stick-slip

experiments are mechanically similar to crustal behaviours during earthquakes. In addition,

Scholz (1968) found that the microcracking events (AEs) generated during rock deformation

experiments radiate elastic waves in a manner similar to earthquakes and that these radiations

obey the Gutenberg and Richter frequency-magnitude relation. Furthermore, Lockner (1993)

found an analogy between AE produced by the brittle failure of rock at laboratory scale and

seismic waves generated by earthquakes. McGarr (1999) studied over 14 orders of

magnitude in seismic moment and found that stick-slip friction experiments provide insights

for interpreting earthquake energy-budget data over a broad range of hypocentral

environments. In addition, Thompson et al. (2009) demonstrated that stick-slip behaviour on

a pre-faulted sample shares similarities with large-scale complex fault zones. All these

studies indicate that AE laboratory experiments can be used to study large-scale earthquake

phenomena. This thesis applies observations from laboratory simulations of fluid-induced

seismicity experiments with the aid of the AE techniques to provide insights on the processes

and mechanics of field scale RIS.

1.3 Thesis objectives and overview

Many theoretical and numerical models have been developed to show the spatial and

temporal distribution of reservoir-induced seismicity (do Nascimento, 2002; do Nascimento

et al., 2004, 2005a & 2005b) and fluid-injection induced seismicity. Most of these models

address mechanical factors and focus mainly on explaining the rapid and delayed responses

of seismicity due to pore pressure variation. They do not model the seismic frequency and

Page 31: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

13

magnitude responses, which are both useful for RIS and/or earthquake forecasts.

Furthermore, the pore diffusion effect cannot fully explain the mechanism of protracted

seismicity with hypocenters that do not necessarily migrate greater distances. I hypothesize

that protracted seismicity can be influenced by the frequency and amplitude of pore pressure

oscillation, as inferred by Roeloffs (1988), to a greater extent than the mechanical properties

of the subsurface. In view of this, my research aims to use laboratory triaxial experiments

with the aid of acoustic emission techniques to study the effect of oscillating pore pressure on

RIS, in particular the effect on protracted seismicity. There are four main objectives of this

research:

i) To identify whether pore pressure fluctuation can reactivate pre-existing faults and

induce protracted seismicity.

ii) To investigate any evolution of the rate and magnitude of seismic events.

iii) To identify any spatial and/or temporal distribution trends of seismic events.

iv) To investigate the applicability of an existing failure forecast model to RIS.

Chapters 2 and 3 provide an account of the theories of reservoir-induced seismicity and

research techniques, respectively. Compressive loading tests were carried out for the

investigation using a triaxial deformation cell, which allows the formation of natural

fractures in the samples prior to cyclic pore pressure loading. Both mechanical and

continuous acoustic emission data were acquired during the experiments so that the spatial

and temporal distribution of acoustic emissions (AE) could be reconstructed for different

stages of the experiments. Initial testing was conducted to characterise the behaviour of

Darley Dale and Fontainebleau sandstone samples, as detailed in Chapter 4. These initial

Page 32: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

14

findings provided information for the design of the subsequent cyclic pore pressure

experiments, which simulate the effects of initial and protracted seismicity. The experimental

results are also presented in Chapter 4. Furthermore, discussions and analyses are provided in

Chapter 5. Two post-experimental analyses were also carried out to provide additional

evidence of the influence of cyclic pore pressure on induced seismicity (Chapter 5). These

include:

i) X-ray micro-computed tomography analysis, which yields an overview of the fracture

behaviour of the tested samples.

ii) Optical microscopy analysis, which aims to understand the microscopic-scale

phenomenon of the influence of cyclic pore pressure on the fractured zone.

The F5 Fontainebleau sandstone experimental data and two RIS field cases were used to

demonstrate the validity of a failure forecast model in Chapter 6. Lastly, the conclusions and

recommendations of the research are stated in Chapter 7. The calibration data and the sensor

files are listed in Appendixes I and II, respectively. The details of the F2 initial experiment

are contained in Appendix III, AE response to different rate of increase in pore pressure is

illustrated in Appendix IV, glossary of terms is listed in Appendix V and, the information of

the paper published in Geophysical Research Letters is enclosed in Appendix VI.

Page 33: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

15

Chapter 2 Theory

The study of RIS phenomena involves the interaction between fluid and rock, and hence

it is essential to understand the physics of porous media, and basic rock fracture mechanics.

Furthermore, the understanding of the seismic patterns of the two types of RIS: initial and

protracted RIS is also important for the study of RIS. Thus, this chapter is divided into three

sections to address the fundamentals of porous media, rock fracture mechanics, and

seismicity patterns.

2.1 Fundamentals of porous media

2.1.1 Definition of porosity

Porosity, Φ, is the measure of the pore volume within the rock. It is defined as the

fraction of rock volume Vr that is not occupied by solid matter. If the volume of solids within

the rock is denoted by Vsolid and the pore volume as Vpore = Vr – Vsolid, then porosity can be

defined as:

VolumeTotal

VolumePore

V

V

V

VV

r

pore

r

solidr ==−

=Φ Equation 2.1

There are two main groups of porosity: primary and secondary. Primary porosity is the

main or original porosity system in a rock or unconfined alluvial deposit. Secondary porosity

is a subsequent or separate porosity system in a rock, which often enhances the overall

porosity of a rock. This can result from the chemical leaching of minerals or the generation

of a fracture system. Secondary porosity can either replace the primary porosity or coexist

with it.

Page 34: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

16

Fracture porosity is the secondary porosity associated with a fracture system or faulting,

while vuggy porosity is the secondary porosity generated by the dissolution of large features

(such as macrofossils) in carbonate rocks, leaving large holes, vugs, or caves (Guéguen and

Palciauskas, 1994). When primary and secondary porosities overlap and interact, dual

porosity is formed. Dual porosity can be found in fractured rock aquifers where the rock

mass consists of primary porosity and the fractured system contributes to secondary porosity.

However, not all the pores or fractures in a system are interconnected and allow the passage

of fluids. For instance, dead-end pores and non-connected cavities do not allow fluid flow.

Only interconnected pores or cavities which allow fluid flow contribute to effective porosity.

2.1.2 Porosity measurement

There are several ways to estimate the porosity of a given material. The more common

methods include the direct method and the imbibition method. The direct method measures

the two volumes, Vr and Vsolid directly and, the porosity is given as 1-(Vsolid/Vr). This gives

the average porosity of the rock, because Vr-Vsolid includes all the pore spaces, both the

continuous pore network and pores that are not connected to the rock exterior.

For the imbibition method, a porous sample is immersed in a wetting fluid such as

distilled water, for a sufficiently long period of time, which allows the fluid to enter into all

the pore spaces that are connected to the rock exterior. The saturated weight of the sample is

then taken. The dry weight of the sample is measured after the sample is oven-dried for over

a sufficiently long period of time, such that the weight of the sample remains unchanged. The

temperature of the oven is limited to 80 oC to avoid inducing thermal cracks in the sample.

Page 35: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

17

The difference in weight will be ρVpore where ρ is the known density of the fluid and Vpore is

the pore volume. A volumetric displacement measurement of the saturated sample will give

the bulk volume, Vr, of the sample. Porosity is therefore determined by Vpore/Vr. This method

will yield the best values for the connected (effective) porosity.

In this study, porosity was determined using the standard ISRM water saturation

porosity technique. The samples were oven-dried for 72 hours at 80 oC before the dry

weights were taken. Then the samples were saturated with distilled water for more than 24

hours before the 100% saturated weights were taken. The difference between the dry and

saturated weights of the samples was used to calculate the effective porosity of the samples

using equation 2.1. The porosities of every sample used in the experiments were measured.

2.1.3 Permeability

Permeability describes the ability of a medium to transmit fluid and is greatly

influenced by the porosity of the medium. There are several ways of measuring the

permeability of a rock sample. The most common method is the steady state flow method,

which is based on Darcy’s law. This can be explained by considering the fluid volume which

crosses perpendicular to a cross-sectional area ‘S’ per unit area per unit time (Figure 2.1).

Page 36: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

18

Figure 2.1 Schematic diagram of Darcy’s law: elemental sample length X and surface area S.

A porous material that possesses a permeability of 1 Darcy under a pressure gradient of

1 atmosphere per centimetre produces a flow rate of 1 c.c./second for a fluid viscosity of 0.01

Poise through a 1 cm sided cube (Dullien, 1979; Guéguen and Palciauskas, 1994). One

Darcy is equivalent to 9.7 x 10-13

m2. However, Darcy’s law breaks down with turbulent flow

(with high Reynolds number). This suggests that the volume flow rate during the

measurement of permeability must be kept low. Typically, Darcy’s law applies when the flow

rate is less than 1 m/s (Guéguen and Palciauskas, 1994).

The Darcy’s law for horizontal flow is given as:

dX

kdPq

η−= Equation 2.2

where q is the Darcy velocity (volume/time/area) of the fluid, k is the permeability, η is fluid

viscosity (which is constant at a given temperature), and dP/dX is the pressure gradient over

the sample length.

X

P P’

S

q

Page 37: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

19

Equation 2.2 can be rearranged to calculate permeability k by substituting (P-P’) for the

differential pressure, where P is the pressure at one end of the sample and P’ the pressure at

the other end:

)'(/ PP

Xq

dXdP

qk

−≈=

ηη Equation 2.3

The Darcy fluid velocity, q, is defined as the fluid volume that passes through a

sectional area, S, over a period, t. This velocity can also be expressed as Vf/S, where Vf is the

volume flow rate (m3/second). For an experimental setup with a vertical flow through a

cylindrical sample of length X and diameter d, S can be expressed as πd2/4; consequently,

equation 2.3 can be written as:

2

4

)'( d

X

gXPP

Vk

f

πη

ρ±−= Equation 2.4

where g is the gravitational constant, and ρ is the density of the fluid.

This constant pore pressure method allows the determination of continuous changes in

permeability as the pressure changes and is likely to be the situation in nature. However, as

the name suggests, this method is not suitable for environments with variable pore pressure

conditions and very low-permeability rock will take a long time to achieve a steady-state

flow.

In the case of very low permeability rocks, the pulse decay method suggested by Brace

et al. (1968) is more efficient. This method measures the change in pressure through time at

one end of a sample due to a sharp pressure pulse, which is introduced at the other end

Page 38: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

20

(Hsieh et al., 1981; Trimmer, 1981). The disadvantage of this method is the interpretation of

the exponential pressure decay versus time curve, which requires processing after stable

experimental conditions occur after a few minutes. This method, therefore, disregards the

first few minutes of good quality data and relies on the latter data segment.

Another technique for measuring the permeability of rock samples is the oscillating

flow method. This technique requires the introduction of a well-controlled sinusoidal

pressure fluctuation upstream of the sample under test. The pore pressure oscillation is

superimposed upon the ambient pore pressure at one end of the rock sample. This method

allows continuous measurements during gradual changes in the state of the rock-environment

system. When both the amplitude ratio and the phase difference of the upstream and

downstream reservoir pressures are measured, permeability and hydraulic diffusivity can be

calculated. This allows the calculation of the magnitude and changes of the effective rock

porosity (Kranz et al., 1990; Fischer, 1992; Bernabé et al., 2006; Song and Renner, 2007).

However, there are several drawbacks to the oscillating flow method. First, the optimum

frequency of oscillation and the ratio of downstream-to-upstream pore pressures depend

upon sample size and the magnitude of permeability. For high porosity sandstones, the

oscillating period can be seconds, while for low porosity rocks such as granites, the pore

pressure oscillation period can be about an hour (Kranz et al., 1990). This imposes a limit on

the range of measurable permeability and hydraulic diffusivities. As for samples with very

high or very low porosities, the required oscillation period may either be unachievable or is

too time consuming. Second, the time required for the measurement of low porosity samples

may increase drastically; as in a typical laboratory setting, several tens of downstream

Page 39: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

21

pressure oscillations are required in order to perform a Fourier analysis of the signals in

moving time windows of at least several cycles (Bernabé et al., 2006). Third, the storativity

measurements usually suffer very large uncertainties, which greatly influence the accuracy of

the permeability calculation. The optimization method used for solving the non-linear

coupled equations are extremely sensitive to noise in the recorded data and, hence, may lead

to erroneous calculations. Finally, systematic errors may occur owing to early time transients,

pressure leaks, or long-term temperature variations (Bernabé et al., 2006). In view of these

limitations, the simple steady-state flow method was used in all the permeability

measurements performed in this research.

2.2 Rock fracture mechanics

2.2.1 Stress relations and Coulomb failure criterion

Figure 2.2 shows a two-dimensional stress field applied to a cylindrical rock specimen.

The normal stress σn and shear stress τ components are taken as the axes of abscissa and

ordinate, respectively. Normal stress lies perpendicular to the failure plane, while shear stress

lies along the failure plane. The angle between the σ1 plane and the fault plane is denoted as

α. Shear stress and normal stress can be expressed by equations 2.5 and 2.6, respectively.

τ = ( σ1 – σ3 ) sin2α / 2 Equation 2.5

and

σn = ( σ1 + σ3 ) / 2 – ( σ1 – σ3 ) cos2α / 2 Equation 2.6

where σ1 and σ3 are the principal stresses.

Page 40: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

22

Figure 2.2 Schematic figure of the stress field.

The relation between τ and σn with principal stresses σ1 and σ3 is represented in Figure

2.3.

Figure 2.3 Mohr diagram for normal shear stresses produced by the principal stresses.

In a triaxial test, σ1 and σ3 can be presented by a Mohr circle at any point in time. A

family of Mohr circles can be obtained from a series of failure tests with increasing values of

σ1 and σ3. From these circles, a failure envelope can be defined, with the combination of τ

and σn for which failure is reached (Figure 2.4). The failure envelope can be expressed by

σ1

σ3

σ1

σ3

τ

σn

α

σ

σ3

σ1

τ σn = (σ1+σ3) / 2 - (σ1-σ3) cos2α / 2

(σ1+σ3)/2 (σ1-σ3)/2

τ = (σ1-σ3) sin2α / 2

Page 41: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

23

equation 2.7:

τ = τ0 + σn tanφ Equation 2.7

where τ0 is cohesion, tanφ is the coefficient of internal friction (often showed as µ), and φ is

the angle of internal friction.

σn

Failure Envelope

σ1

τ0

τ

φ

Region of Stability

0

τ=τ0+σntanφ

σ3

Figure 2.4 Mohr stress circles for a series of tests showing failure according to Coulomb failure criteria.

The Coulomb failure criterion with the consideration of effective stress relates the

normal and shear stresses, which is effective in describing macroscopic-scale failure. It

describes the instantaneous condition of the material and is expressed independently of time.

Furthermore, it does not consider the contribution of rough rock joints and their surface

morphology. This simple criterion may provide a rough interpretation of initial or rapid RIS,

which is represented by a shift of the Mohr circle toward the failure envelope. However, in

the case of protracted RIS, variation in subsurface condition due to pore pressure oscillation,

mineral dissolution, etc. can vary the strength of the pre-existing faults with time and space.

This spatial and temporal variation in the strength of the fault zone cannot be addressed by

this macroscopic failure criterion.

Page 42: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

24

During initial reservoir impoundment or large water level change, shear failure may

occur on pre-existing fault planes or fractures. Studies by Gough et al. (1970) and Snow

(1972) focus on the elastic loading effects of reservoir impoundment and elastic stress

changes. The elastic response of the subsurface to reservoir loading causes changes in normal

and shear stress underneath the fault plane. This can be shown by Mohr circles in Figure 2.5,

where the loading of reservoir causes σ1 to increase and, hence, pushes the Mohr circle

towards the failure envelope.

σn

τ

σ3 σ1 (after)σ1 (before)

Figure 2.5 Mohr circle representing the elastic effect of reservoir loading on the strength of rock underneath the

reservoir. The dashed semi-circle represents the strength before loading and the solid line semi-circle represents

the strength after loading.

If an increase in elastic loading is the dominant mechanism for RIS, seismicity would be

triggered soon after reservoir impoundment and the seismicity would soon be stabilized

when the additional strain has been released. The elastic loading effect of water

impoundment would only be effective in thrust fault and normal fault settings and would not

be effective in strike-slip fault settings, as the increase in vertical load would not influence σ1

and σ3 of strike-slip faults. Figure 2.6 shows the change in stability of a fault plane relative to

the position of the reservoir in thrust fault and normal fault settings. In the case of a

Page 43: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

25

strike-slip fault, there is no change in ∆τ; therefore, the elastic theory would suggest that this

fault setting is not prone to RIS. However, many RIS case histories have been found in

strike-slip fault setting, for instance, the earthquakes at the Açu reservoir in Brazil (do

Nascimento, 2002), the Koyna reservoir in India (Gupta, 2002), the Xinfengjiang reservoir in

China (Talwani and Chen, 1998), and the Hoover dam in the USA (Guha, 2000) are all

located on strike-slip faults. This suggests that the net increase in impoundment load is not

the dominant mechanism in these RIS cases. Furthermore, numerous case histories of

protracted RIS have continued for decades, such as that at the Koyna reservoir (Gupta, 2002)

and the Xinfengjiang reservoir (Talwani and Chen, 1998). This seismicity cannot be

explained by the elastic loading effect of reservoir impoundment. Other protracted RIS cases

such as that at the Aswan reservoir, Egypt, had protracted RIS recorded during the period

when the surface water level was lowered (Selim et al., 2002). This further suggests that the

increase in elastic loads is not the dominant mechanism for RIS.

Figure 2.6 Schematic figure of the change in stability of a fault plane relative to the position of the reservoir (do

Nascimento, 2002).

Hanging

wall

Foot

wall

Hanging

wall

Foot

wall

reservoir reservoir reservoir reservoir

σ1

σ1 σ3

σ3

a) Normal fault b) Thrust fault

Legend:

unstable

stable

Page 44: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

26

2.2.2 Pore pressure effects

Pore fluid has both mechanical and chemical effects on induced seismicity. However,

the current understanding of the nature of RIS places emphasis on the physical effect of pore

pressure enhancement due to the increase in surface water levels. Mechanically, pore

pressure can reduce the effective normal stress that acts on the existing faults or fractures via

the law of effective pressure (Equation 2.8) and, hence induce seismicity. If the medium is

porous and contains fluid with pore pressure P, the principal stresses are reduced to the

effective values:

σ’ = σ – P Equation 2.8

where σ’ is the effective stress. In a Mohr diagram, an increase in pore pressure (with σ1 and

σ3 being constant) moves the Mohr circle toward failure (Figure 2.7).

σn

τ

P

σ3-P σ1σ1-Pσ3

Figure 2.7 Mohr circle representing the fluid pressure effect on the strength of rock. The dashed semi-circle is

the stress before fluid pressure is applied and the solid line semi-circle is the stress after the fluid pressure is

applied.

Page 45: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

27

For RIS, the instantaneous increase in pore pressure in the substratum is due to the

increase in impoundment load at the surface. This increase in pore pressure will decrease the

normal stress that holds the fault and, thereby, induce slips along the pre-existing faults.

The long-term chemical effect of water may change the strength of fractured rocks by

decreasing the coefficient of friction (Shen et al., 1974; Kisslinger, 1976; Talwani and Acree,

1984/85; Chen and Talwani, 1998; and Chen, 2001). The chemical interactions of pore fluid

with the rock matrix are discussed by Chen (2001) and will not be discussed in this thesis, as

the focus of the research is on the mechanical effects of pore fluid. In addition, no chemical

effects of pore fluid would have taken place during the relatively short period of experiment.

According to Coulomb failure criterion, the incremental shear strength, ∆S, along a

pre-existing fault plane due to reservoir impoundment, is given by Bell and Nur (1978) and

Talwani (2000):

∆S = µ ( ∆σn – ∆P ) – ∆τ Equation 2.9

where ∆τ is the changes in shear stress on the fault, ∆σn is the change in compressive normal

stress across the fault, µ is the coefficient of internal friction, and ∆P is the change in pore

pressure. Negative values of ∆S indicate fault weakening, while positive values imply fault

strengthening. A decrease in ∆S is due to a decrease in ∆σn, or an increase in pore pressure or

∆τ (Figure 2.8). The calculation of ∆S as a means of assessing the stability of a fault depends

on knowledge of the geometry of the fault, the assumed magnitude and orientation of the

regional stress, and the assumed value of the friction coefficient.

Page 46: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

28

shear stress = τ = τ0 + µ ( σn – P )

Figure 2.8 Pore pressure and shear stress at a faulted surface (Simpson and Narasimhan, 1992).

Pore pressure acts in two different deformation regimes: drained and undrained

(Detournay and Cheng, 1993; Guéguen and Palciauskas, 1994). Each regime has a different

effect on RIS.

i) Undrained Response (Rapid Response of Initial Seismicity)

Due to the increase in water level at the surface, there will be an instantaneous increase

in pore pressure in the substratum. This increase in pore pressure due to elastic compression

leads to an undrained response, which weakens the fault and may cause failure. If this

mechanism is dominant, the corresponding RIS will be instantaneous and only minimal delay

between the peak of the water level and the maximum seismic activity will be observed (do

Nascimento, 2002).

Simpson (1976) considered the effects of elastic stress changes and pore pressure

changes due to reservoir impoundment. He showed that elastic stress changes decreased

quickly away from a reservoir in the case of finite reservoir length, while the increased pore

pressure did not dissipate as quickly with distance from the reservoir. This increase in pore

τ

σn P

Page 47: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

29

pressure, due to the undrained response, can persist until pore pressure slowly dissipates into

the surrounding fractures. As a result, the reduction in effective pressure can last for a period

of time depending on the mechanical properties of the environment. Simpson and

Narasimhan (1992) proposed that heterogeneities in rock properties can also cause a

localised pore pressure increase due to the undrained response. Consequently, this causes a

rapid seismic response due to reservoir impoundment.

ii) Drained Response (Delayed Response of Initial Seismicity and Protracted RIS)

When sufficient time is allowed for fluid to move through the rock mass, pore pressure

can decrease to the original value; hence, the drained response is associated with delayed RIS.

In most RIS cases, there is a time lag between reservoir impoundment and the onset of

seismicity. Howells (1974), Talwani (1976), and Talwani and Acree (1984/85) suggest that

the pore pressure diffusion effect is a possible mechanism for delayed RIS. Simpson (1976),

Bell and Nur (1978), Roeloffs (1988), and Rajendran and Talwani (1992) use the coupled

elastic effect and pore pressure diffusion effect to explain the phenomenon of RIS. The

change in pore pressure is governed by the basic diffusion equation given by Jaeger et al.

(2007):

t

P

Dz

P

∂∂

=∂∂ 1

2

2

Equation 2.10

where t is the time and D is the hydraulic diffusivity, which can be expressed as follows:

ηβk

D= Equation 2.11

Page 48: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

30

where k is the permeability of the rock, η is the viscosity of the pore fluid, and β is the bulk

compressibility of fluid-filled rocks.

Analytically, the delay can be expressed in terms of the distance below the ground

surface and the hydraulic diffusivity, which is a function of permeability, k. If a pressure P(z

= 0, t = 0) is imposed at the surface and retained for t > 0, and D is a constant, the solution to

equation 2.10 is given by Carslaw and Jaeger (1959):

)2

(1)0,0(

),(

Dt

zerf

P

tzP−= Equation 2.12

where P(0, 0) is the original pore pressure change at the bottom of a reservoir due to

reservoir impoundment, P(z, t) is the pore pressure with depth z at time t after the

impoundment, and erf is the standard error function.

By assuming D = 106 m

2/day and that distance z varies from sample distance of 125 mm

to field distance of 0.4 to 5 km, a set of curves can be plotted by using equation 2.12 (Figure

2.9). The plot indicates that the pressure front arrives at shallow depths first and the time

required for the pressure front to reach greater depths increases significantly. In the case of

the laboratory experiment with 125 mm long sample, it takes less than a few seconds for the

pore pressure front to arrive from the upstream (applied pore pressure end) to the

downstream at the other end of the sample. The effect of D, at a depth of 1 km and at sample

depth of 125 mm is illustrated and compared in Figure 2.10, which shows that the time

required for the pressure front to reach a certain depth increases with decrease in D and,

decreases with decrease in distance.

Page 49: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

31

Figure 2.9 Showing the values of equation 2.12 for different values of z and t and D = 106 m

2/day.

Figure 2.10 Showing the values of equation 2.12 for different values of D and t, and with z = 1 km and 125 mm,

respectively.

Page 50: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

32

Roeloffs (1988) discussed the fault stability changes induced beneath a reservoir with

cyclic variations in water level, with consideration of the fully coupled effect of pore

pressure and elastic stresses, i.e., the elastic stresses influence pore pressure and vice-versa.

She calculated stress and strength changes produced by a steady periodic variation of the

water level on the surface of a uniform porous elastic half-space and modified equation 2.12

to include the effect of undrained response. For a unit step increase in pore pressure at the

surface, P(0, t) = H(t), she calculated the pore pressure at a depth z after time t, P(z, t). For a

one-dimensional case, the pore pressure at depth z and time t becomes:

))((2

)1(),( tHDt

zerfctzP δδ +

−= Equation 2.13

where erfc is the complementary error function, H(t) is the Heaviside unit step function, δ =

B(1+vu)/3(1-vu) for isotropic conditions, B is the Skempton’s coefficient, and vu is the

Poisson’s ratio measured under undrained condition. The coupled response may be

dominated by the undrained response immediately on impoundment and be primarily caused

by diffusion at a later stage.

Roeloffs (1988) also showed that in the case of periodic reservoir water level change,

the magnitude of pore pressure change due to diffusion at any depth depended on the

frequency of water level changes, i.e., the higher the frequency of water level changes, the

lower the magnitude of pore pressure change. She showed that at depths below z, there was

almost no change in pore pressure due to diffusion:

z = 2 π (2D/ω)1/2 Equation 2.14

Page 51: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

33

where ω is the angular frequency of water level changes. This concept of the fully coupled

effect of pore pressure and elastic stresses has been used to explain the seismicity at Lake

Mead, the Nurek reservoir (Roeloffs, 1988), the Koyna reservoir, the Oroville reservoir

(Rajendran, 1992), the Monticello reservoir (Rajendran, 1992; Rajendran and Talwani, 1992),

and the Xinfengjiang Reservoir (Shen and Chang, 1995).

Under drained conditions (long-term), other factors such as the hydrochemical

properties of the substratum, the chemical composition of the fluid, the nature of gouge and

clay that clogged the fractures, etc. can influence the permeability of the fractured rock and,

hence, affect the rate of pore fluid diffusion and the time lag in seismic response (Chen,

2001).

2.2.3 Rock friction

Friction is of great importance in rock fracture mechanics. Its effects can be observed on

all scales: ranging from the microscopic scale, in which friction occurs between contact

surfaces; to larger scales such as interaction between grains, aggregates, or surfaces of joints

and fractures; and to even larger scales on the order of square meters or square kilometres,

where friction acts between fault surfaces. The frictional resistance of two surfaces in contact

is defined as the ratio of shear resistance to the normal stress. This concept is referred to as

Amontons’ law, which is similar to but predates the Coulomb friction model by almost a

century (Lockner and Beeler, 2002):

τ = µσn Equation 2.15

Page 52: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

34

where τ is the shear stress, σn is the normal stress, and µ is the static friction coefficient.

If there is any variation of frictional resistance during sliding, a dynamic instability can

occur, resulting in a sudden slip associated with stress drop. This dynamic instability is

followed by a period of no motion during which the stress is recharged, followed by

instability. This variation of frictional resistance can be caused by the presence of pore fluid,

which can mechanically affect friction by pore pressure, and/or the chemomechanically

weakening of the solid (Scholz, 2002). Under saturated conditions, the basic friction

coefficient is reduced to an apparent friction coefficient.

The mechanism of dynamic instabilities can be demonstrated using a spring with

stiffness K, pulling a block subject to a normal force FN (Figure 2.11). When the applied

force reaches the frictional resistance of the block, the block is free to slide on the rigid

surface. The hypothetical case in Figure 2.12 shows a maximum resistance followed by a

decrease with continued slip. During the slip, the spring unloads and follows the line with

slope K. After a tangent point B is reached, the frictional resistance F decreases at a rate

greater than K. Instability therefore occurs due to a force imbalance, which causes

acceleration of the block. Beyond point C, resistance becomes greater that the force in the

spring and the block decelerates and come to rest at point D. The shaded area between points

B and C is equal to that between points C and D (Scholz, 2002). The condition for instability

is given by:

Kdu

dFN > Equation 2.16

Page 53: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

35

Figure 2.11 Block-slider model demonstrating initiation of frictional instability, where FN is the normal force

acting on the block, K is the spring stiffness, and F is the applied force. When the force reaches the frictional

resistance of the block, the block is free to slide on the surface (Scholz, 2002).

Figure 2.12 Force displacement diagram showing a hypothetical case in which frictional resistance falls with

displacement at a faster rate than the system responds (Scholz, 2002).

Byerlee (1978) conducted laboratory measurements of frictional strengths for silicate

and carbonate rocks and proposed linear relationships to distinguish their frictional behaviour

under low normal stresses (σn < 200 MPa) and high normal stresses (200 MPa < σn < 2000

MPa) (Figure 2.13).

τ = 0.85 σn for σn < 200 MPa Equation 2.17

and

F

Slip, u

B

C

D

Slope = K

F

FN

K

Page 54: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

36

τ = 50 + 0.6 σn for 200 MPa < σn < 2000 MPa Equation 2.18

This relationship is known as Byerlee’s law, which indicates that rock types including

silicates and carbonates have friction coefficients that vary from 0.6 to 0.85. However, the

surface energy of silicates can be reduced by the presence of water (Scholz, 2002).

Figure 2.13 Shear stress plotted as a function of normal stress at the initial friction for a variety of rock types.

For crustal faults, the friction coefficients may vary from ~0.5 to 1.0 (Zoback et al.,

1987; Townend and Zoback, 2000), which is a wider range of friction coefficients than that

given by Byerlee’s law. The lower-end friction coefficients are caused by the presence of a

fault gouge with hydrated mineral phases which weaken the fault, while the upper-end

friction coefficients are caused by time-dependent fault healing and the presence of

heterogeneities. Byerlee’s law is limited to certain rock types and does not account for

dynamic instabilities which depend on second-order friction properties such as roughness

Initial Friction

0

200

400

600

800

1000

1200

1400

0 500 1000 1500 2000

Normal Stress (MPa)

Sh

ear

Str

ess

(M

Pa)

τ = 0.85σn

τ = 50 + 0.6σn

Page 55: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

37

and sliding velocity. Despite the differences in scales, there has been extensive evidence that

the strength of the continental crust is governed by Byerlee’s law. However, the laboratory

values of the rate and state variables that control stability must be extrapolated cautiously to

larger scales (Scholz, 2002).

2.3 Seismicity patterns

Earthquakes seldom occur as isolated events, but as part of a sequence. Before large

earthquakes occur, local seismicity rates often show a significant increase in number. This

seismicity is referred to as foreshocks, while smaller earthquakes that follow main shocks are

termed aftershocks. This earthquake sequence has well-defined characteristics. Earlier

laboratory experiments by Mogi (1963) found that homogeneous media are characterised by

seismicity with no foreshocks, slightly heterogeneous media have a number of foreshocks

preceding the main shock, and that extremely heterogeneous media are characterised by a

swarm type of earthquake activity. Extremely heterogeneous media are often defined as

media with variations in material type as well as materials created by previous deformations

(Couples and Lewis, 2007), while slightly heterogeneous material can be interpreted as

originally homogeneous material that has undergone deformation and has formed

heterogeneous faulting. These three types of foreshock–aftershock sequences are also

exhibited by natural earthquake sequences (Figure 2.14). Mogi (1963) inferred that the

mechanical structure of the media and the nature of the applied stresses could be responsible

for these patterns. Since the stresses could be regarded as nearly uniform for tectonic

earthquakes, the pattern of the earthquake sequences would be mostly influenced by the

degree of heterogeneity of the media.

Page 56: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

38

Figure 2.14 Mogi’s (1963) classification of foreshock-aftershock patterns and their relationship to the structures

of materials and applied stresses. N is the number of seismic events and t is the time. Type I: in the case of

homogeneous material and uniformly applied stress, a main shock occurs without any foreshock and is followed

by numerous elastic aftershocks. Type II: when the material has a rather heterogeneous structure and/or the

applied stress is not uniform, small elastic shocks occur prior to a main shock and many aftershocks occur

following the main shock. Type III: when the structure of the material is extremely heterogeneous and/or the

applied stress has a considerable concentration, a swarm type of activity occurs consisting of a number of

elastic shocks with magnitudes increase gradually and then decrease after some time.

2.3.1 Aftershock sequence

There are three empirical scaling relations that describe earthquake aftershock

sequences: i) the Gutenberg-Richter frequency-magnitude scaling, ii) the modified Omori

Page 57: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

39

law for the temporal decay of aftershock rates, and iii) Båth’s law for the difference in the

magnitude of a main shock and its largest aftershock (Shcherbakov et al., 2004). The first

two relations were used in Chapter 5 for analysing the aftershock sequences of the pore

fluid-induced seismcitity.

i) Gutenberg-Richter scaling

The Gutenberg-Richter scaling is the frequency-magnitude scaling for aftershocks.

Under a wide variety of conditions, the number of earthquakes in a specified region and time

window with magnitudes greater than M is given by the relation:

log N(≥ M) = a – bM Equation 2.19

where N is the number of events with magnitudes greater than M on the Richter scale, and a

and b are constants, in which b is the slope of the line measured from the linear descending

portion of the graph.

This relation is valid for earthquakes both regionally and globally, with b-values

typically between 0.5 and 1.5 (Von Seggern, 1980), while the constant, a, gives the logarithm

of the number of earthquakes with a magnitude greater than zero. Studies have shown that

aftershocks correspond to the Gutenberg-Richter scaling relation with b-values similar to

those for main shocks (Kisslinger, 1976). The physical meaning of b-value equals unity is

that there are 10 times more events of a smaller magnitude than that of a magnitude which is

greater by 1. The change in b-value indicates the change in frequency and magnitude of

earthquakes over time. Because precursors often indicate an increase in frequency and

Page 58: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

40

magnitude prior to a main shock, the corresponding b-value is expected to decrease with the

growth of larger magnitude events.

Laboratory experiments of brittle rock deformation have been carried out to study the

fracture mechanism of micro-cracking events (Scholz, 1968; Main et al., 1988). These

studies show that rock deformation experiments in the laboratory and deformation of the

crust exhibit similar mechanisms. In addition, the fluctuations in b-value are consistent with

the intermediate and short-term earthquake precursors (Main et al., 1988).

ii) Modified Omori law

The decay of aftershock sequences follows the Omori law. The modified Omori law is

defined as (Utsu, 1961):

ptc

WtN

)()(

+= Equation 2.20

where N(t) is the number of aftershocks in an interval of time t, W is the amplitude, c is the

time offset parameter, and p is the exponent that modifies the decay rate, typically found to

be very close to 1 (Scholz, 2002).

iii) Båth’s law

Båth’s law states that the differences in magnitude between a main shock and its largest

aftershock are approximately constant, typically about 1.2 (Båth, 1965) for large, shallow

earthquakes. This relationship is independent of the magnitude of the main shock. The

relationship is given by:

Page 59: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

41

∆M = Mm – Mamax

Equation 2.21

where Mm is the magnitude of the main shock and Mamax is the largest magnitude aftershock.

2.3.2 Foreshock sequence

Jones and Molnar (1979) found that foreshock sequence can be fit with an empirical

relation:

N = at–m

Equation 2.22

where t is the time before the origin time of the main shock, N is the frequency of foreshocks,

and a and m are constants. Papazachos (1975) and Kagan and Knopoff (1978) found that the

exponent m is close to 1. Later, Helmsetter et al. (2003) found that foreshocks obey the

inverse Omori law, which states that the rate of earthquakes prior to a main shock increases

on average as a power law proportional to 1/(tc-t)p of the time to the main shock occurring at

time, tc and p is the exponent that modifies the rate of change of the foreshocks.

Page 60: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

42

Chapter 3 Techniques: The Use of Rock Physics and Laboratory Tools

This chapter describes sample preparation, triaxial loading equipment configuration,

and a data acquisition system. An account of the acoustic emission (AE) processing

techniques, including the procedures for extracting discrete AE from continuous AE records

and the methods of calculating source locations and source mechanisms, is illustrated.

Furthermore, descriptions of X-ray micro-computed tomography (CT) techniques and optical

microscopy are provided.

A schematic flow line diagram of the experiment and analysis is indicated in Figure 3.1,

and the author’s contributions are highlighted below. The sketch is divided into five sections,

as follows:

i) Triaxial experiment: the experiment involves the fracturing of a rock sample followed

by pore pressure oscillations to induce AEs that simulates RIS. These experiments were

designed by the author. However, the addition of an intermediate step was suggested by Dr.

Phil Benson to separate AEs obtained during the formation of fault from that obtained during

cyclic pore pressure stage. Furthermore, all the samples were prepared by the author, i.e.

coring, cutting and polishing.

ii) Data acquisition: this is divided into ultrasonic data acquisition and mechanical data

acquisition. Mechanical data are acquired by an NI data acquisition unit. The ultrasonic data

are divided into velocity data, triggered data, and continuous data. The author was involved

in the initial setting up of data acquisition units and trouble-shooting of the interfacial

problems of mechanical and electronic systems.

iii) Data processing: this involves the use of the InSite program for the processing of both

Page 61: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

43

velocity survey data and discrete AE data. Continuous data recorded by a Giga RAM

recorder are processed by a SeisAcq event harvester, while continuous data recorded by a

‘Richter’ system are processed by a Streamer harvester. All data processing and analysis were

carried out by the author.

iv) Output: this includes the velocity structure of the sample over time, AE source locations

and mechanisms, and the evolution of b-values. The mechanical data output (which includes

permeability, stress, strain, flow volume, and pore pressure) are automatically logged,

converted, and displayed by a LabVIEW Virtual Instrument (VI) program, which the author

took part in writing the program for this purpose.

v) Analysis of fractured sample: this includes X-ray tomography and optical microscopy.

The X-ray tomography scanning was carried out by the Department of Chemical Engineering

and Applied Chemistry, while the thin sections were prepared by the Department of Geology.

However, the thin section photo captures and both X-ray tomorgraphy and optical

microscopy analyses were performed by the author.

Page 62: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

44

Figure 3.1 Schematic diagram of the flow line from the experimental setup to data acquisition to data

processing to output and analysis. Post-experimental analysis includes X-ray tomography and optical

microscopy.

3.1 Sample materials and preparation

The rock types featured in this thesis are Darley Dale and Fontainebleau sandstones.

Cylindrical samples 125 mm in length by 50 mm in diameter were prepared, conforming to

the 2.5 length/diameter ratio recommended by the ISRM (International Society of Rock

Mechanics). The samples were prepared using a hollow coring drill, with the end faces

ground and polished to within 0.01 mm parallelism using a diamond grinding wheel. Prior to

experimentation, the samples were vacuum-saturated with distilled water for at least 72

hours.

Passive data

by AE DAQs

Active data by

velocity survey

Triggered data

Continuous data

Velocity data

InSite Processor

SeisAcq / Streamer event harvester

Discrete AE

Ultrasonic data

Mechanical data (stress, strain, volume, pore pressure)

NI data

acquisition

Permeability, strain, stress, flow volume, pore pressures

Sample

Velocity structure

AE source location AE source mechanism

b-values

Triaxial Experiment Data Acquisition Data Processor Output

Experiment Section

Analysis on fractured sample: i) X-ray tomography, ii) Optical microscopy

Page 63: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

45

3.1.1 Darley Dale sandstone

Darley Dale sandstone is a poorly sorted, well-indurated felspathic sandstone from the

North of England, with grain sizes ranging from 0.08 mm to 0.8 mm (Ayling et al., 1995), an

average length of 0.22 mm, and a porosity of 13%. The cementing material is predominantly

silicious (Baud and Meredith, 1997). It was chosen for its relatively homogeneous nature and

highly reproducible stress-strain behaviour, which is ideal for the study of frictional sliding

along natural fractures. The intact samples used in this research have a permeability of 5.7 x

10-15

m2 under 5 MPa hydrostatic pressures.

3.1.2 Fontainebleau sandstone

Fontainebleau sandstone is found in the Ile-de-France region South of Paris. It is

composed of more than 98% quartz, with small quantities of accessory clay minerals. It is a

well-sorted sandstone (David and Darot, 1993), with an average grain size of approximately

0.25 mm. However, as the degree of silicification differs from one sample to another, the

bulk porosities and the characteristics of the porous networks show a wide range of variation

(Bourbié and Zinzner, 1985). In this research, the Fontainebleau sandstone samples typically

have a low porosity of about 4% and the permeability of the intact samples is 1.3 x 10-17

m2

under 5 MPa hydrostatic pressures.

3.2 Equipment

3.2.1 Triaxial deformation cell and triaxial compression loading machine

During a triaxial experiment, the axial load is applied by an external hydraulic actuator

operated under a displacement-rate control mode (Figure 3.2). The axial load is controlled by

Page 64: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

46

a fast-acting hydraulic valve and measured by an internal load cell to an accuracy of ±0.005

MPa (or ±10 Pa). The actuators are equipped with a servo-controlled feedback system such

that a complete stress-strain curve can be recorded. Under constant strain rate loading, a

feedback signal is sent from the displacement transducer and is compared to the programmed

signal. If a discrepancy occurs between the two values, the hydraulic pressure is adjusted

until the feedback signal matches the programmed value. The raw signals are recorded in

volts and are converted to stresses in MPa by an automated LabVIEW program.

Figure 3.2 Compression loading machine and triaxial deformation cell.

The triaxial deformation cell is equipped with 18 piezoelectric transducer crystals (12

lateral and 6 axial transducers) for acoustic emission (AE) detection and three pairs of

ultrasonic wave transducers (stacks) for contemporaneous elastic velocity measurement

Triaxial

deformation

cell

Loading

Machine

Page 65: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

47

along the three axes of the sample, for P-, S1- and S2-waves. The lateral piezoelectric

transducers, as well as the lateral velocity stacks, are held in position by the rubber jacket,

which separates the saturated sample from the confining oil (Figures 3.3, 3.4). The face of

these lateral transducers and velocity stacks are curved to fit the surface of cylindrical

samples. A North-East-Down (NED) coordinate system is used to describe the position of the

transducers with respect to the sample (Figure 3.5). Prior to the experiment, pencil-lead break

testing is performed at the face of each transducer to identify the sensitivity of the

transducers. A detailed sensor location file is attached in Appendix II.

Figure 3.3 Cross section of the triaxial deformation apparatus showing the top and bottom axial platens. The

apparatus is equipped with 6 axial transducers positioned in the faceplate of the platens and an ultrasonic

velocity measurement system made of piezo-ceramic stacks comprise of three channels (P-S1-S2).

Naturally fractured

sample

Page 66: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

48

Figure 3.4 Rubber jacket holding lateral transducers and stacks in position.

Figure 3.5 North-East-Down (NED) coordinate system.

3.2.2 Servo-controlled permeameter system

The servo-control permeameter system consists of two identical hydraulic intensifiers,

which provide high-pressure pore fluid to each end of the sample (Figure 3.6). Each

intensifier has a maximum internal volume of 75 ml and is filled with distilled water. The

N

D

E

Lateral transducers

inserted in rubber

jacket

Lateral strain

measurement device

Lateral

stack

Rubber

jacket

Page 67: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

49

intensifiers are servo-controlled using analogue electrical feedback from an external pressure

transducer to maintain any pressures up to 100 MPa, to an accuracy of +/-0.005 MPa. The

intensifiers are fitted with integral displacement transducers which monitor the position of

the high-pressure pistons, thus enabling the intensifiers to be used as volumometers. For

steady-state flow permeability measurements, the upstream (pore pressure port at the bottom

platen) and downstream (pore pressure port at the top platen) intensifiers are set to slightly

different control pressures in order to maintain a small, constant pressure gradient across the

sample (typically a 0.5 MPa difference). During the pre-peak stress constant pore pressure

stage of an experiment, the pore pressure is kept at a constant value (average of 5 MPa) and

the two intensifiers is controlled independently. As a result, permeability is not measured

continuously during pore pressure oscillation. Steady-state flow measurement of

permeability was measured at intervals that would not affect the oscillations of pore

pressures. Thus, they were measured prior to the oscillations and after the completion of

oscillations. Figure 3.6 indicates the schematic diagram of the permeameter and the

connection to the triaxial deformation cell. Another pair of pressure transducers is located in

close proximity to the triaxial deformation cell for the detection of upstream and downstream

pressures. This minimises the head loss along the pipelines and can provide more accurate

measurements. Raw signals are recorded as voltages and are converted into pressures in MPa

by the LabVIEW program.

Page 68: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

50

Figure 3.6 Schematic diagram of the experiment setup. Showing a pair of servo-controlled intensifiers

connected to the downstream (top platen) and upstream (bottom platen) of the triaxial deformation cell.

3.2.3 Confining pressure pump

Confining pressure is also controlled by a pair of servo-controlled pressure pumps. Each

Page 69: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

51

intensifier has a maximum internal volume of 68 ml and is filled with hydraulic oil. The two

intensifiers are controlled in tandem, using analogue electrical feedback from two external

pressure transducers. The confining pressure pump can maintain any pressure up to 100 MPa,

with an accuracy of +/-0.05 MPa. Another confining pressure transducer is attached close to

the triaxial deformation cell, which monitors confining pressure in the triaxial deformation

cell and provides accurate measurements. The signals are recorded by a data acquisition

board in voltages, which are then converted to pressures in MPa by the LabVIEW program.

3.2.4 Linear variable differential transformer (LVDT)

The corresponding axial shortening of the sample due to axial loading is measured by

two linear variable differential transformers (LVDTs). This LVDT is a common type of

electromechanical transducer that converts mechanical rectilinear motion into a

corresponding electrical signal. They are used for monitoring the axial displacement of the

sample due to compression and during cyclic pore pressure loads and are spaced radially

from the triaxial deformation cell (Figure 3.3), such that the measurements are not affected

by the elastic shortening of the entire loading train. The measured signals are in voltages and

are converted to strain by the LabVIEW program.

3.2.5 Data acquisition (DAQ) hardware

All mechanical experimental data are logged as a function of time using a National

Instrument (NI 6255) high-speed multifunction data acquisition (DAQ) board (Figure 3.7),

running on a computer equipped with LabVIEW 8.3 data acquisition software. This

high-speed device allows 16-bit accuracy with measurements of all channels operating at

Page 70: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

52

maximum speeds. During the experiments, data is acquired at a rate of 100 kHz and an

average of 100 points. The acquisition board is equipped with 40 differential analog input

ports or 80 single-ended input ports, referenced to the ground. Differential inputs provide

higher resolution data than single-ended inputs, as the differential amplifier amplifies the

difference in voltage between the two inputs, which influences the output voltage in opposite

ways. During an experiment, the differential input settings are used for readings that require

high accuracy, such as pore pressures and LVDT measurements. Other measurements such as

axial loads and confining pressures are acquired through the single-ended ports.

Figure 3.7 NI6255 data acquisition board.

All raw data acquired by the DAQ are in voltages and are imported into the LabVIEW

data acquisition program, also known as a virtual instrument (VI), for data conversion, data

Page 71: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

53

storage, as well as for displaying various real-time plots and values during experiments.

Equipment is calibrated prior to the experiments. By applying linear regression, the gradients

and intercepts of each transducer calibration line can be obtained. These regression results

(gradients and intercepts) are plotted in section 3.7. The calibration parameters are imported

into the logging software so that all measurements are converted into the appropriate

physical quantities (e.g., transducer pressures in MPa, volumometer volume in ml, and

LVDT measurements in strain). The converted measurements are then collectively stored in a

data file and displayed in the program panels. The schematic block diagram of the LabVIEW

VI program is illustrated in Figure 3.8.

Figure 3.8 Flow chart showing the LabVIEW data acquisition program.

3.3 Geophysical techniques

3.3.1 Ultrasonic-wave velocity survey and AE techniques

During an experiment, AE are received by an array of 16 receiver channels, while

ultrasonic-wave velocity surveys are performed at regular time intervals during the pre-peak

stress stage and between each set of pore pressure cycles during the post-peak stage.

Page 72: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

54

a) Ultrasonic-wave velocity survey

Ultrasonic-wave velocity measurements are particularly valuable when used in

conjunction with active AE measurements because the surveys provide data on the change in

the elastic properties of the material associated with rapid strain release. An ultrasonic wave

is generated by a synthetic source with a known location. By processing these waveform data,

the change of seismic wave velocities and amplitudes over time can be determined. This

information also implies the evolution of transport properties and the damage of the sample

over time.

To assess the velocity structure of a rock sample during an experiment, active 3-D

velocity surveys were performed at ten minute intervals prior to the rupture of the sample,

using the three pairs of designated transducers along the three axes. The signal-to-noise ratios

were improved by stacking the signals 50 times. During the subsequent pore pressure

oscillation stage, velocity surveys were acquired prior to or after the completion of a cyclic

pore pressure sequence, in order to avoid disturbing the pore pressure oscillations.

During a velocity survey, an electrical pulse is generated by a pulse generator, which

excites the piezoelectric transducer to produce a mechanical signal. The signal then

propagates through the rock sample and is detected by a receiver (transducer) at the other end,

which transforms this mechanical signal into an electrical signal. As the piezoelectric

material is located some distance behind the end cap metal of the transducers, the P- and

S-wave arrival times through a specimen have to be calibrated. This can be done by

performing a face-to-face test of the end platens or by inserting a specimen (e.g., aluminium

Page 73: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

55

cylinder) with known elastic properties between two lateral stacks. With a known

dimension/diameter and the elastic wave velocities of the specimen, the times-of-flight can

be back-calculated. These times-of-flight are subtracted from the arrival times acquired for

velocity surveys to calculate accurate velocities (Equation 3.1).

V = Ds/(T-To) Equation 3.1

where Ds is the sample distance or sample diameter; T is the time difference between sending

a pulse and receiving a signal via a high-frequency data-acquisition system (Cecchi digital

storage oscilloscope), with the presence of a sample; To is the travel time between pulsing

and receiving a signal with no sample present between the receiver and pulser; and V is the

elastic wave velocity. The calibration table of To values of P-, S1- and S2-wave

measurements in the three axes are tabulated in Table 3.1. The automatic and/or manual

picking of elastic wave arrivals is performed by finding the first break signal which deviated

from an amplitude level of 0 volts. The manual picking of elastic wave arrival times requires

some level of experience and the picking of all channels and events needs to be consistent.

Table 3.1 Calibration for ultrasonic wave arrival time measurements.

To along X-axis (µs) To along Y-axis (µs) To along Z-axis (µs)

P S1 S2 P S1 S2 P S1 S2

23.155 25.455 26.765 23.185 25.505 26.935 16.997 34.605 35.255

A Milne AE recorder allows 3-D ultrasonic velocity surveys to be performed at regular

time intervals. During a survey, each of the 16 selected transducers is excited sequentially by

an electrical pulse to produce a mechanical signal. This signal is recorded by the other 15

Page 74: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

56

transducers in the array. Consequently, a maximum of 240 potential ray paths can be

generated.

b) Passive acoustic techniques

In this research, AE were detected using transducers consist of lead-zirconate-titanate

piezoelectric ceramic material (PZT-5A, Navy Type II) with a frequency range of

approximately 20 kHz to 2 MHz. Each transducer has PZT element with diameter of 2 mm

by 1.4 mm thick. These methods can describe the localisation and extent of induced

seismicity and monitor the propagation of events. Elastic waves emitted from the initiation

and propagation of cracks during the formation of a natural fracture and shearing or sliding

on existing fractures during the pore pressure oscillations are recorded across an array of

transducers, which consists of 16 transducers (typically 12 from the lateral, 2 from the top

platen, and 2 from the bottom platen). Each transducer/receiver is assigned as a channel for

the recording of elastic waves (Appendix II). Each receiver is connected to a Physical

Acoustics Corporation (PAC) pre-amplifier (model 1220 A) or a Pulser-Amplifier Desktop

Units (PAD) (Figure 3.9), which has a filter range of 10 kHz to 1.0 MHz. The signals

received are amplified by 60 dB. The relationship between decibel and amplification is

shown in Equation 3.2, which indicates that 60 dB is equivalent to 1000-times amplification

in amplitude. The subsequent inversion of the elastic waves’ arrival times for each event

provides a localisation of the source of the waves.

dB = 20 log (Wout/Win) Equation 3.2

where Wout is the output amplitude and Win is the input amplitude.

Page 75: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

57

Figure 3.9 PAD pre-amplifiers used for amplifying signals by 60 dB

During an experiment, two types of passive data are recorded, namely discrete data and

continuous data. Both sets of data are acquired by two different monitoring systems. Discrete

AE data are monitored by a Milne AE recorder (ASC, 2007) and a Hyperion Ultrasonic Giga

RAM Recorder (ESG, 2002a), while continuous data are recorded by the Hyperion

Ultrasonic Giga RAM Recorder and a recently developed ‘Richter’ system. The Milne AE

recorder is part of the OMNIBUS acoustic emission and ultrasonic monitoring system (ASC,

2007), which is used for collecting discrete data of more than 50 events-per-second. This unit

provides 10 MHz 12-bit full-waveform acquisition. Signals received by sensors are

pre-amplified by a PAD pre-amplifier system (Figure 3.9). It also records the number of hit

counts (number of AEs) per channel. A ‘hit’ is recorded when the waveform signal voltage

exceeds a pre-defined threshold within a pre-defined time window. The threshold voltage

depends upon the experiment. In this research, this threshold is typically set to 60 mv, which

was about 3 to 4 times the peak noise level, and events with amplitudes of less than 60 mv

would not meet the triggering criterion.

Page 76: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

58

The Hyperion Ultrasonic Giga RAM Recorder (ESG, 2002a) was used as a secondary

system for acquiring discrete data. During experiments, triggered AE waveform data were

recorded with a maximum capacity of 16 events per second. A minimum of 5 channels was

set as the triggering criteria and each channel is assigned an individual channel threshold.

When at least 5 channels record trigger threshold crossings with a 5 ms time window, an AE

‘hit’ is recorded (ESG, 2002a). These triggered data were sampled at a rate of 10 MHz and

recorded by First-In-First-Out memories.

Conventional AE recorders have limited capacity for recording triggered events during

high AE activity, as ‘mask-time’ occurs during the transfer of data from volatile memory to

permanent storage (Thompson, 2006). However, this inadequacy was overcome by the

Hyperion Giga RAM recorder developed by Young and Bowes (2002), which allows

continuous data acquisition and reconstruction of a complete AE catalogue by eliminating

AE system saturation. Data were digitized at 14-bit resolution, with an input voltage range of

±2.5 V, continuously streamed to a 40 GB RAM buffer, and sequentially copied to hard drive.

The experiments described in this thesis used sampling frequency of 10 MHz, resulting in a

segment of 134.8 s continuous waveform on 16 channels, which can be viewed at various

resolutions (Figure 3.10).

Discrete AE events were extracted from the continuous ultrasonic record using the

SeisAcq program developed by ESG (2002a). This program essentially replays the

continuous record after the test and applies a trigger logic to extract events. In this thesis,

events were extracted when the amplitude of at least five channels exceeded a trigger

Page 77: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

59

threshold of 60 mv in a 100 µs window. These events were then processed as conventional,

discrete AE events for source location and source mechanism analysis.

Page 78: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

60

Figure 3.10 Continuous waveform of a Darley Dale sandstone experiment and a discrete acoustic emission

waveform extracted from continuous waveform data at different zoom levels.

Page 79: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

61

Similar to the Giga RAM Recorder, the ‘Richter’ system was recently developed for

continuous AE data acquisition. The limitation of its capacity is the storage capacity of the

hard drives. During an experiment, the most critical AE data are those that occurred during

pore pressure oscillation; consequently, this system is only activated for a 2-hour period

during pore pressure oscillation. The system consists of 8 individual computer systems,

designated B0052 to B0058. Each computer is responsible for recording continuous

ultrasonic waveforms to the allocated channels. System B0052 is assigned as the ‘master’

computer and linked to the other seven ‘slave’ computers. In order to activate all eight

computers for data recording, a signal is sent from the B0052 to the ‘slave’ computers, which

are all synchronised prior to an experiment. Similarly, by reconstructing and ‘re-harvesting’

the continuous data, discrete events of AE can be restored. These events can then be used for

AE source location for the study of migration trends or the nucleation of events. The

software program used for the ‘re-harvesting’ of events is called Streamer, which is another

subset of the InSite software (ASC, 2007). Figure 3.11 indicates a flow chart which

illustrates the acquisition of passive and active AE data by different monitoring and recorder

systems.

Page 80: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

62

Figure 3.11 Flow chart showing the various AE data acquisition units for recording passive and active AE data.

3.3.2 AE methods

AE waveform data are processed to provide information on the micromechanics

involved in the evolution of damage, including nucleation and propagation processes, and in

rock mass structure and competence (Young et al., 2004). These methods include:

i) Source location and parameter analysis: Analysis of spatial and temporal trends of AE

activity due to pore pressure oscillation.

AE source locations or hypocentres were calculated using first arrival times recorded by

an array of transducers. This technique enables the mapping of the temporal and spatial

Page 81: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

63

evolution of microcracks. First arrival times were picked by an automatic picking routine

using the InSite waveform processor (ASC, 2007), or manually picked to enhanced accuracy

(Figure 3.12). In order for the auto-picking routine to work effectively, different velocity

models were assumed in the different time periods of an experiment. The velocity model was

expected to change as with changes in differential stress. Microcracks were expected to

orient parallel to the maximum compressive force at the beginning and change the sample

from an initially isotropic to transversely isotropic structure. A transversely isotropic (TI)

structure generally develops with the fast axis parallel to the maximum compressive stress,

due to cracks opening parallel to the maximum loading axis. The velocity data was fitted to a

model in the form:

V(θ) = (Vmax + Vmin)/2 - (Vmax-Vmin)/2 cos (180-2θ) Equation 3.3

where V is the modelled P-wave velocity as a function of the ray-path angle, θ, and Vmax and

Vmin are the maximum and minimum P-wave velocities, respectively.

Page 82: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

64

Figure 3.12 Showing typical P-wave arrivals received from each channel. The solid blue line indicates the first

arrival of the P-wave. The dotted pale blue line indicates the peak amplitude of the first arrival of the P-wave.

Page 83: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

65

There are four unknowns in a source location equation. The first three describe the

location of the event, which includes x0, y0 and z0, while the fourth is t0, which is the time of

the AE. The unknowns were solved using a Simplex algorithm, which is an iterative

procedure that minimises errors between the measured and calculated arrival times (Nelder

and Mead, 1965; Press et al., 1994). The procedure uses a geometrical shape known as

Simplex, which has one extra vertex as compared to the number of dimensions of the defined

geometry. The error distance for each vertex is minimised through an iterative process. At

least seven arrival times are required for solving the source location in order to minimise the

AE location error.

There are three major errors in source location. First is the error associated with picking

the first arrival times; second is the uncertainty in the velocity model; and third is the ±2 to

±4 mm uncertainty in the location of transducers. In order to reduce the location error, travel

time residuals between the calculated and measured arrival times for a channel are limited to

two sample points; if the residual exceeds the limit, the channel is omitted from the location

algorithm. The most accurate method of assessing location errors is by locating a synthetic

source, i.e. locate the transmitters during ultrasonic wave-velocity surveys and then calculate

a residual between the actual location and the calculated location. This allows fine tuning of

the source location velocity structure. This source location calibration is illustrated in

Appendix I, which indicates that absolute source location error ranged from ±2 to ±4 mm.

According to Pettitt (1998), the error at the centre of the sample was approximately half that

at the edge. Therefore, for the experiments conducted in this thesis, a source location error of

approximately ±1 to ±2 mm was expected near the centre of the sample. Subsequently, AEs

Page 84: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

66

closer to the centre of the samples were chosen for source mechanism analysis due to their

relatively higher location accuracy, as compared to events occurred near the edge of the

sample.

ii) Source mechanism studies: Analysis of the failure mechanics of AEs as a function of the

compression and tension forces acting at the source and the orientation of failure, giving

information about movement on existing microstructures and new crack and damage zone

creation.

Source mechanism analysis determines the modes of failure operating in terms of how

the sample is reacting to external forces. The simplest method to determine the source

mechanism of an AE is to calculate a fault plane solution as described by Gibowicz and

Kijko (1994). This method requires the polarity of P-wave arrivals to be identified at a

number of sensors in an array. The polarity recorded at a sensor indicates whether the first

motion at the source is dilation (positive) or compression (negative) for a particular ray path.

The polarities are plotted on a lower hemisphere stereograph representing the focal sphere of

the source. If a shear type event is assumed, then the equivalent forces are modelled as a

double couple (DC). Orthogonal nodal planes are drawn to separate regions of dilation and

compression. One of the planes is the fault plane, while the other is an auxiliary plan. The

differentiation of the two planes requires additional information.

The source mechanisms of AE are often characterised using a ratio of first arrival

polarities, with a tensile event having all positive values and a compression event having all

Page 85: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

67

negative values (Figure 3.13). For an event with mixed polarity data, the success of fitting a

fault plane solution can define the event source as either shear or complex.

Figure 3.13 Examples of focal sphere and equivalent forces. Shaded regions represent dilation and white

regions represent compression (modified from Aki and Richards, 2002).

Another method for analysis source mechanism was developed by Pettitt (1998) based

on a time invariant moment tensor approach. The error quantification routine was improved

by Collins et al., (2002). This approach consists of two main stages, namely the inversion of

moment tensor and its composition into a source model. A brief account of moment tensor

approach is summarized in the following.

A seismic source can be described by 9 elemental force couples (a couple being a pair of

equal and oppositely oriented forces, and the moment of each, Mij, comprises the moment

Page 86: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

68

tensor of the source. Conservation of angular momentum for the equivalent force couples

means the moment tensor is a symmetric second order tensor (3 x 3 matrix) requires only 6

components (Figure 3.14).

Figure 3.14 The nine elementary force couples which describe general source mechanism when combined in

moment tensor. The matrix notation of subscripts 1, 2, 3 correspond to x, y, z in Cartesian axes at the source

(Aki and Richard 2002).

By assuming a point source, the procedure can be simplified to a time invariant process.

Only channels receiving far-field signals will be considered in the moment tensor analysis,

while those receiving near-field signals will be disregarded. This further simplifies the source

radiation pattern. The near-field distance can be calculated by Equation 3.4 (Savage and

Aronson, 2005).

Page 87: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

69

L = dT2/4λ Equation 3.4

where L is the length of the near field; dT is the diameter of the transducer; and λ is the

wavelength, which equals velocity divided by frequency.

The moment tensor is calculated by the inversion of P-wave first arrival amplitudes

recorded by the receiver array (Equation 3.5). The Green’s function matrix inversion is

performed using Singular Value decomposition after De Natale and Zollo (1989). As the

particle response of the sensors is not known, the procedure performs a relative amplitude

inversion, and calculates a relative event magnitude. At least 7 first motion P-wave

amplitudes are used in the moment tensor analysis, and the residual between the modeled and

observed amplitudes is used for checking the quality of the inversion. Since good focal

coverage is important for the accuracy of MT analysis, events are selected only from the

central third of the sample for this analysis. Furthermore, these selected events are processed

manually to enhance the accuracy. Finally, the stability of each inversion is tested by

dropping random arrivals (Thompson et al., 2009).

[ak] = [Gki,j][Mij], or

where [ak] is a vector containing n number of amplitudes, [Gki,j] is the Green’s function

=

a1

:

:

:

:

ak

G11,1 G11,2 G11,3 G12,2 G12,3 G13,3

: : : : : :

: : : : : :

: : : : : :

: : : : : :

Gk1,1 Gk1,2 Gk1,3 Gk2,2 Gk2,3 Gk3,3

M11

M12

M13

M22

M23

M33

Equation 3.5

Page 88: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

70

indicating the propagation effects between the source and the receiver, and [Mij] is a vector

containing the 6 moment tensor components.

The physical mechanism of moment tensor can be interpreted using 3 eigenvectors, ai,

and 3 eigenvalues mi. The eigenvectors are mutually orthogonal in space, representing the

principal stress axes of the source; while the eigenvalues represent the magnitude and

polarity of the forces. The convention used for the principle stress axes are pressure (P), null

(B) and tension (T), representing minimum, intermediate and maximum eigenvalues

respectively. The forces can be represented in linear vector dipoles in Figure 3.15.

Figure 3.15 Vector force representation of some source models: a) negative Linear Vector Dipole (LVD), b)

Double Couple (DC), c) pure implosion, d) pure explosion, e) positive Compensated Linear Vector Dipole

(CLVD), and f) negative Compensated Linear Vector Dipole (CLVD) (Pettitt, 1998).

iii) Magnitude-frequency (b-value) trends for the identification of spatial and temporal

changes in rock mass response.

a) b) c)

d) e) f)

-1

-1

-1

1

1

1

-0.5

-0.5 0.5

0.5

-1 1

1 -1 -1

Page 89: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

71

Previous studies have shown that AE hits have a power law frequency-magnitude

relationship similar to the Gutenberg-Richter b-value relationship (Equation 2.19) for

earthquakes. It was found that the b-values depend primarily on stress, and are much higher

in ductile rock than that observed in brittle rock (Scholz, 1968). For AE, magnitude is

calculated as location magnitude mL using equation 3.6.

=∑

=

n

n

m

mRMS

LS

dW

m

)(

log 1 Equation 3.6

where mL is an estimate of relative magnitude between a set of events, Sn is the number of

sensors and dm is the distance between receiver m and the AE location and WRMS is

calculated according to equation 3.7.

x

W

W

x

i

i

RMS

∑== 1

2

Equation 3.7

where Wi is the waveform amplitude and x is the number of data points in waveform.

The power law frequency-magnitude relationship for AE is the same as the

Gutenberg-Richter relationship for earthquakes, except that the magnitude M is replaced by

mL. Thus, equation 2.19 can be written as:

LL bmamN −=)(log Equation 3.8

where N is the number of events with magnitudes greater than mL, and a and b are constants,

Page 90: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

72

in which b is the slope of the line measured from the linear descending portion of the graph.

3.4 Source of experimental error

There are two main categories of experimental error: physical experimental

measurement errors and human errors involved in the interpretation of physical results (Table

3.2). Physical errors can be minimised through standard good practices such as averaging

physical measurements and performing accurate calibrations. Human errors are often a

greater source of error. These errors involving the interpretation of physical results are

inevitable, but can be minimised by interpreting the results using consistent criteria, such as

during the picking of the first arrivals of elastic waves.

Table 3.2 Summary of experimental error sources and accuracy

Experimental Errors Absolute Accuracy

Sample length/diameter ±0.01 mm

Sample surface ±0.01 mm

Sample parallelism ±0.01 mm

Sample mass ±0.001 g

Axial load ±0.005 MPa

Displacement ±0.0005 mm

Confining pressure ±0.05 MPa

Pore pressure ±0.005 MPa

P-wave arrival time ±0.2 µs

Page 91: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

73

3.5 X-ray micro-computed tomography

A three-dimensional X-ray micro-computed tomography (CT) technique has been

applied to geomaterials to provide 3-D information on fault zones (Mees, 2003; Benson et al.,

2003; Salvo et al., 2003; Thompson et al., 2009). This 3-D information is obtained from a

series of 2-D measurements. The basis of X-ray micro-CT is X-ray radiography: an X-ray

beam is sent through a sample and the transmitted beam is recorded on a CCD-based detector

(Salvo et al., 2003). In this study, the 3-D X-ray micro-CT technique is adopted to understand

the variation of flow path along the vertical axis of the sample. In order to achieve a higher

resolution of X-ray tomography (34.89 µm x 34.89 µm x 34.89 µm), the volume of scanning

is limited to the central most representative range of the sample. The reconstructed 3-D

images can indicate the complexity of the fractured network within the samples and provide

an account of the homogeneity or heterogeneity of the fractured zone. A representative area

is chosen from the scanned image (slice), which is then used for the calculation of the

relative crack area. The resulting areas are normalised by the minimum crack area of all the

slices of the sample. These normalised crack areas are plotted according to the sequence of

the slice. The resulting profiles illustrate the relative ease of fluid flow along the vertical axis

during oscillation of pore pressure. This analysis can indicate zones with relatively narrow

flow paths, in which transient pore pressure accumulation is expected during pore pressure

cycles.

3.6 Optical microscopy

One sample each of tested Fontainebleau and Darley Dale sandstone is imbibed with

epoxy to stabilize the fractures prior to obtaining optical microscopy specimens. Three thin

Page 92: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

74

section specimens are prepared for each sample: one on the xy-plane, one on the xz-plane,

and one on the yz-plane. Precise microscopic information regarding the microstructure of the

porous networks can be related to the variations of the measured transport properties.

Microscopy photos are taken with a CMOS colour microscopy camera, with 1.3 mega-pixel

resolution. The analysis of the microscopy can provide an account of the difference between

the fractured zone and the far-field area. In addition, the comparison of the microscopy of the

two rock types (Darley Dale sandstone and Fontainebleau sandstone) can illustrate the

difference in the behaviour of the samples under cyclic pore pressures.

Page 93: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

75

Chapter 4 Laboratory Simulation of RIS due to Oscillating Pore Pressures

In RIS, the variation of surface water level affects the pore pressure at the bottom of the

reservoir. It is appropriate to assume that the water level cycles are equivalent to the pore

pressure cycles at the bottom of the reservoir (Talwani, 1997). Furthermore, do Nascimento

et al. (2005) numerically show that the use of sinusoidal water level variation has only a

minor effect on the resulted diffusivity, as opposed to the use of real seasonal water level data.

Therefore, in the experiments, sinusoidal pore pressure variation was used as a proxy of

surface water level change.

Initial tests were carried out to identify three important parameters for the design of

cyclic pore pressure experiments: first, to identify the compressive strength and behaviour of

samples under different confining pressures; second, to identify the behaviour of the sample

during reduction in pore pressure and the different rate of increase of pore pressure; and third,

to investigate the behaviour of samples under different frequencies of pore pressure

oscillations. The findings from these initial sets were used as inputs for the design of the

subsequent cyclic pore pressure experiments. The peak value of the compressive strength of

the samples determined which pair of axial platens to be used during the cyclic pore pressure

experiment. Titanium platens were designed to deform samples with compressive strength

less than 500 MPa, while hardened stainless steel platens were designed to deform samples

with compressive strength up to about 600 MPa.

4.1 Behaviour under compressive load tests

Three triaxial compressive loading experiments were carried out on the saturated Darley

Page 94: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

76

Dale sandstone under a strain rate of 2 x 10-6

s-1

and confining pressures of 10, 20, and 40

MPa, respectively, to provide the parameters used for the cyclic pore pressure experiments.

The behaviour of the rock became more ductile and the failure angle range reduced from 71o

to 66o, as higher confining pressures were used. Figure 4.1 shows the loading curves of three

initial tests. Although only three initial tests, each at different confining pressure was

conducted to provide the strength and behaviour of the samples, the number of initial tests

were considered adequate because Darley Dale sandstone has been widely studied and is

known to have consistent behaviour under deformation tests (e.g. Baud and Meredith, 1997).

In addition, the failure behaviour of the samples (ductile or brittle failure) was of a greater

importance than the absolute value of the peak strength in the case of cyclic pore pressure

experiments in the post-peak regime. Furthermore, the stress-time plots of the Darley Dale

sandstone illustrated in section 4.8 showed that the failure behaviour of the sandstone was

consistent with that indicated in the initial testing.

The experimental results suggest that 40 MPa confining pressure creates ductile

deformation, in which the stress-strain curve does not provide the clear stress drop and

rupture occurrence essential for indicating the end of stage 1 of the experiment. Conjugate

sets of faults were formed during the experiment, which enhance the difficulty in the study of

the migration trends of seismic events. Therefore, 40 MPa confining pressure was not used in

the cyclic pore pressure experiment.

Page 95: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

77

Figure 4.1 Stress-strain curves of Darley Dale sandstone samples under 10, 20, and 40 MPa confining pressures.

The behaviour of Fontainebleau sandstone was very brittle. Samples subjected to 15 and

20 MPa confining pressure both failed with a brittle ruptured, which provided a clear mark

for the beginning of the post-peak regime. However, under these confining pressures, the

samples did not generate a clear shear failure plane. Axial splitting and complex shear

conjugates were developed (Figure 4.2). Although only two initial tests, each at different

confining pressure was conducted to provide the strength and behaviour of the samples, the

number of initial tests were considered adequate because Fontainebleau sandstone has been

widely studied and is known to have consistent behaviour under deformation tests (e.g.

Bourbié and Zinzner, 1985). Similarly, the failure behaviour of the samples (ductile or brittle

failure) was of a greater importance than the absolute value of the peak strength in the case

of cyclic pore pressure experiments in the post-peak regime. In addition, the stress-time plots

Page 96: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

78

of the Fontainebleau sandstone experiments illustrated in section 4.8 showed that the failure

behaviour of the sandstone was consistent with that indicated in the initial testing.

With the information from these initial compressive load tests of the two sandstone

types, it was decided that a confining pressure of 20 MPa, which represents conditions at 800

m in the crust, would be used for cyclic pore pressure experiments for both types of

sandstone for comparison purposes. This confining pressure not only allows the development

of a clear sample rupture mark (i.e., a sharp stress reduction) for both sandstone types, but

also allows a larger variation in pore pressure, which has to be kept below the confining

pressure at all times during the experiment to avoid damaging the rubber jacket.

Figure 4.2 Typical stress-strain curve of the Fontainebleau sandstone with 20 MPa confining pressure.

Page 97: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

79

4.2 Behaviour under pore pressure reduction and subsequent increase

Furthermore, trial experiments were carried out to investigate the responses of fractured

Darley Dale sandstone due to pore pressure variations. After the failure of the sample, the

strain rate was paused. The pore pressure was then reduced to half of that used during the

loading stage to observe the acoustic behaviour of the sample. During this stage, the

quiescence period was obtained with the seismic rate reduced to background levels. This is

due to the increase in the effective stresses that hold the faults together, which then lock the

faults. This quiescence period is essential for the separation of residual events caused by

sample failure from those caused by pore pressure variation.

After quiescence has been observed, the pore pressure was then ramped up beyond the

previous maximum to investigate the possibility of the activation of faults due to pore

pressure increase. Three experiments (DDS4, DDS5 and DDS6) were carried out to

investigate this possibility. A list of the rate of increase in pore pressure that has been used in

various experiments is illustrated in Table A5 (Appendix IV). Two of the experiments

(DDS4 and DDS5) used a constant pore pressure of 1 MPa in the pre-peak stage, and then

lowered to 0.5 MPa during the quiescence period. The subsequent targeted pore pressure was

1.5 MPa, which exceeded the previous maximum. The rates of increase of pore pressure were

0.2 MPa/minute and 0.5 MPa/minute, respectively. In both experiments, no activation of

fault was observed, i.e., there was no distinguishable increase in AE due to the increase in

pore pressure. Consequently, the third experiment (DDS6) was modified to use a higher rate

of increase and a higher amplitude of pore pressure in order to investigate the activation of

faults. During the pre-peak stage, a constant pore pressure of 5 MPa was used, which was

Page 98: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

80

then lowered to 2.5 MPa during the post-peak for quiescence purposes. The pore pressure

was then increased at a rate of 5 MPa/minute to reach a designated pore pressure of 7.5 MPa.

The result suggests that the activation of faults by pore pressure increase is possible and that

the rate of increase in pore pressure used in the subsequent cyclic pore pressure experiments

could be kept at ≥ 5 MPa/minute.

4.3 Behaviour under cyclic pore pressure

Roeloffs (1988) inferred that the frequency of the cyclic pore pressure can play an

important role in RIS. Thus, a third set of initial testing was conducted primarily to identify

the seismic response of faulted samples due to the different frequencies and/or amplitudes of

pore pressure cycles. Because the number of AE responses of Fontainebleau sandstone

during the post-peak regime surpass those recorded in the Darley Dale sandstone

experiments, the third set of initial testing was carried out on the saturated Fontainebleau

sandstone. During the experiment, the sample was loaded to rupture under a constant

confining pressure, a constant strain rate of 2 x 10-6

s-1

and a constant pore pressure of 5 MPa.

The strain rate was then paused and pore pressure was reduced to 2.5 MPa for the locking of

faults and quiescence. For the ease of pore pressure control during this stage, the two pore

pressure ports were linked to perform as a pair, i.e., both ports received the same command

of pore pressure change during the variation in frequency and amplitude of pore pressure.

The details of the pore pressure cycles used in the F2 experiment are tabulated in Appendix

III.

The results of the F2 experiment indicate that seismicity can be activated when the pore

Page 99: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

81

pressures reach or exceed the previous maximum. However, seismicity reduced rapidly

during the subsequent cycles with the same amplitudes. No significant seismicity was

observed when the pore pressure was lower than the previous maximum (Figure 4.3).

However, it is noted that the activation of seismicity began when the pore pressure was lower

than the previous maximum (Figure 4.4). Consequently, it is speculated that if cyclic pore

pressure peaks could be maintained at a higher percentage of the previous maximum, faults

might be reactivated after a greater number of cycles. This postulation was then

demonstrated by the subsequent cyclic pore pressure experiments (F5 and F8).

The AEs generated due to cyclic pore pressure were spatially and temporally scattered

along the pre-existing fault (Figure 4.5). There were no clusters of events or migration trends

observed from the reconstructed temporal seismicity distribution. This behaviour may be

explained by the existence of a heterogeneously faulted structure, consisting of axial splitting

and shear fractures, which causes anisotropy in diffusivity within the sample and, hence the

scattered locality of the pore pressure development that induced the seismic events.

Page 100: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

82

Figure 4.3 Experiment F2 – Pore pressure cycles introduced after the rupture of the specimen.

Figure 4.4 Activation of fault begins when the pore pressure reaches ~4.3 MPa, which is lower than the

previous maximum.

Page 101: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

83

a) b)

Figure 4.5 a) The fractured F2 Fontainebleau sandstone sample; b) spatially and temporally scattered seismic

events induced by pore pressure cycles occurred along the pre-existing faults.

The seismic data of this initial testing was also used for the investigation of the

distribution of seismicity along the sample. With the reference points set at both ends of the

sample where pore pressure was applied (Figure 4.6), the hypocenter distances were then

grouped into 10 mm bins. The results indicate that the number of events tends to increase

toward the middle of the sample (Figure 4.7).

Figure 4.6 Showing the reference distance for the measurement of hypocenter locations during initial tests.

62.5

0

0

Page 102: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

84

Figure 4.7 The number of events located at different ranges of hypocenter. The trend indicated during the cyclic

pore pressure period is similar to that during the pre-peak period.

It was speculated that by applying pore pressure at one end of the sample and recording

the pore pressure response at the other end, more information about the pore pressure

gradient along the fractured sample could be provided. In addition, this procedure provided

longer distances for the migration of events, if any, and hence eased the study of migration

trends. Consequently, this modified procedure would be used in the subsequent cyclic pore

pressure experiments.

4.4 Conclusions from Initial Testing

Initial testing indicates that the amplitude and rate of increase of pore pressure and the

frequency of pore pressure oscillations can influence the activation of RIS. The effect of

cyclic pore pressure frequency on seismicity is demonstrated in Figure 4.9. The specific

frequency range that influence seismicity depends on the mechanical properties of the sample

and would not be determined in this thesis, as it is beyond the scope of this research. The

findings and modifications of the cyclic pore pressure experimental procedures are

summarized in Figure 4.8. The initial experiments indicate that the procedures can be divided

Page 103: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

85

into three stages:

i) Loading of an intact rock sample under a strain rate of 2 x 10-6

s-1

, constant pore

pressure of 5 MPa, and confining pressure of 20 MPa, which replicates shallow crustal

condition of 800 m.

ii) After the failure of the sample, a period of quiescence must be introduced by pausing

the strain rate and reducing the pore pressure to 2.5 MPa. This provides a clear separation of

the residual events caused by the rupture of sample and the events caused by pore pressure

variation.

iii) Oscillating pore pressure with a rate of increase of pore pressure greater than 5

MPa/minute would be used. The peak pore pressures should exceed the previous maximum

for the activation of faults, while peaks at ~95% of the previous maximum would be used for

the investigation of the reactivation of faults due to oscillating pore pressures. The pore

pressure would be applied at the bottom of the sample (upstream port) and measured at the

top of the sample (downstream port). This method is similar to that used for measuring

permeability of intact rock samples (Kranz et al., 1990; Fischer, 1992; Bernabé et al., 2006;

Song and Renner, 2007), which has been discussed in section 2.1.3.

Page 104: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

86

Figure 4.8 The three groups of initial testing and the output used for the design of cyclic pore pressure

experiments.

4.5 Modified Cyclic Pore Pressure Experiments

Two different sandstones: i) Fontainebleau sandstone with a relatively low porosity of

4%, and an initial permeability of ~1.3 x 10-17 m2 and ii) Darley Dale sandstone with a

relatively high porosity of 13%, and an initial permeability of ~5.7 x 10-15

m2 were used in

the cyclic pore pressure experiments. This provides a good comparison of the effect of pore

pressure oscillation on different permeability samples. Furthermore, one control and one

saw-cut experiment were carried out on Fontainebleau sandstone samples. The purpose of the

Page 105: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

87

control experiment was to distinguish the significance of cyclic pore pressure on induced

protracted seismicity and the reactivation of pre-existing faults from that of a step increase in

pore pressure, while the saw-cut Fontainebleau sandstone experiment provided a clear single

fracture for the investigation of any migration trends of AEs. The list of experiments is

tabulated in Table 4.1:

Table 4.1 List of experiments

Fontainebleau sandstone experiments Description

F5 Naturally-fractured sample underwent

oscillating pore pressure

F6 Saw-cut sample underwent oscillating pore

pressure

F7 Naturally-fractured sample with pore pressure

step changes, as a control

F8 Naturally-fractured sample underwent

oscillating pore pressure

Darley Dale sandstone experiment

DDS7 Naturally-fractured sample underwent

oscillating pore pressure

Sinusoidal pore pressure cycles were applied from one end of the sample and recorded

at the other in all the experiments except the control experiment. This procedure required

trial-and-error to identify the oscillation frequency or oscillation period that could be used in

the fault activation and reactivation experiment. The upstream and downstream pore

pressures in Figure 4.9 indicate that the response of the downstream pore pressure was

affected by the frequency of the applied pore pressure. When the frequency was too high, the

downstream pore pressure might not respond according to the applied pore pressure

Page 106: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

88

frequency. In addition, seismicity was not induced when the average pore pressure of the

sample did not exceed the previous average; nevertheless, the applied pore pressure was set

beyond the previous peak values. In this case, existing faults were not activated unless

sufficient time was allowed for the downstream pore pressure to develop. If the oscillating

frequency was too low (or the rate of increase in pore pressure was too low), the downstream

pore pressure synchronised with the upstream pore pressure (Figure 4.9). The optimal pore

pressure frequency for the reactivation experiment was obtained by several trials, such that a

phase shift between the upstream and downstream pore pressures was observed. This phase

shift corresponds to an observable repositioning between the applied upstream pore pressure

and the measured downstream pore pressure. However, a phase shift is not always achievable.

It depends on two main factors. First, the permeability of the sample: when the permeability

is too high, the upstream and downstream pore pressures can equilibrate almost

instantaneously. Second, the structure of the faulted sample: if the naturally-generated fault

exhibits a through connection between the upstream and downstream (i.e., a major axial

splitting through the sample as indicated in Figure 4.10), then the upstream and downstream

pore pressures can also be equilibrated almost instantaneously, without phase shifts.

Page 107: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

89

Figure 4.9 The effect of different pore pressure oscillating frequencies. Left: In a Fontainebleau sandstone

experiment in which the oscillating frequency was too high, the downstream pore pressure (Pressure A) did not

respond according to the applied cyclic signals at the upstream (Pressure B). Although the upstream pore

pressure peaks exceeded the previous maximum, the sample seemed to not experience the effect of the increase

in pore pressure due to the high oscillation frequency. There was no seismicity during this period, as the

downstream pore pressure did not exceed the previous maximum of 5 MPa. Right: In another Fontainebleau

sandstone experiment in which the oscillating frequency was too low, the downstream pressure synchronised

with the upstream pore pressure almost instantaneously.

Page 108: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

90

a) b)

Figure 4.10 An example of a Fontainebleau sandstone experiment in which the phase shift could not be

established. The sample failed with a major axial splitting fault. a) Top view of a failed Fontainebleau sandstone;

b) side view of a failed Fontainebleau sandstone.

The laboratory experiments involve an approximate size-frequency scaling relationship

that has been established in the literature for analysing millimetre-scale laboratory

experiments as compared to field observations in kilometre-scale (e.g. Burlini et al., 2007;

Benson et al., 2008). This size ratio is ~106. In addition, a relatively high frequency of pore

pressure oscillation is required for inducing seismicity in the laboratory. However, in the

field case, when the distance between the upstream and the downstream is large, a much

lower pore pressure oscillation frequency (e.g., seasonal pore pressure oscillations) may

induce seismicity (Saar and Manga, 2003; Roeloff, 1988). Therefore, the laboratory cyclic

pore pressure frequency is in the order of 10-2

Hz, whilst in the field scale, the seasonal water

level oscillation generates a corresponding cyclic pore pressure frequency in the order of

~10-8

Hz; giving an approximate frequency ratio of ~10-2

/10-8

= 106. Similarly, the frequency

of the P-waves of the acoustic emissions in the laboratory experiments is in the range of 500

kHz to 800 kHz, while in the field the frequency of the waveform is in the order of 0.5 Hz to

Page 109: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

91

5 Hz, resulting in an approximate scaling of ~105 to 10

6. Furthermore, the rate of increase of

pore pressure in the laboratory scale to that of the field is also in the order of 106. The details

of the ratio of the rate of increase in pore pressure are indicated in Appendix IV.

4.6 Experiment setup and procedures

In the cyclic pore pressure experiments, initial RIS were investigated by introducing

cyclic pore pressure peaks that exceeded previous maxima, while the investigation of the

reactivation of pre-existing faults used cyclic pore pressures peaks at ~95 % of the previous

maximum (e.g., when the previous set used 18 MPa peak pore pressure, the subsequent set

should use 95% x 18 = ~17 MPa peak pore pressure). Each experiment was divided into

three parts. The first part was the preparation of a ‘naturally’ faulted sample, during which an

intact sample was loaded under a constant strain rate (2 x 10-6 s-1), a constant confining

pressure (20 MPa) and a constant pore pressure (5 MPa) until failure occurred. The second

stage started after the rupture of a sample in which strain rate was paused and pore pressure

was lowered to 2.5 MPa to allow the locking of faults and to obtain quiescence for the clear

observation of fault reactivation in the subsequent stage. The last part of the experiment

began with a series of cyclic pore pressure loading on the fractured sample, to simulate

induced seismic events along the pre-existing fractures. The seismic responses were recorded

by a streaming system, which allows continuous data acquisition for the reconstruction of the

spatial and temporal distribution of seismic events. Different amplitudes of sinusoidal cyclic

pore pressure were used for the study. The most suitable frequency of the sinusoidal cyclic

pore pressure was obtained by testing the response of the fractured sample during the

experiment.

Page 110: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

92

A control experiment was carried out to identify the effect of pore pressure without

oscillation. This experiment used the same procedures as in stages 1 and 2. However,

constant pore pressure steps were applied in stage 3, instead of pore pressure oscillations.

The pore pressure steps were 18 MPa for 10 minutes and 17 MPa for 32 minutes, which were

the equivalent durations of the 5 cycles of main shock-aftershock sequence and the 16 cycles

of foreshock-main shock-aftershock sequence, respectively, in the F5 experiment.

A saw-cut Fontainebleau sandstone sample was prepared for the F6 experiment to

investigate the migration trends of the AE events during the oscillation of pore pressure. The

sample was pre-cut at 60o and it was expected that all the AEs would occur along the cut. The

experimental procedures were slightly modified. During the experiment, the sample was first

loaded at 5 MPa hydrostatic pressure, with a constant pore pressure of 1 MPa. An axial strain

rate of 2 x 10-6

/s was then applied until the sample was close to failure (i.e., when the sample

began to indicate ductile behaviour with a reduction in the rate of increase of axial stress in

response to the applied constant axial strain. The strain was then stopped and the pore

pressure was reduced to 0.5 MPa for the quiescence period. After that, a series of pore

pressure oscillations was introduced to reactivate the fault.

4.7 Velocity survey

Velocity surveys were conducted periodically during each experiment, and each

consisting of 9 velocity components, namely: P-, S1- and S2-waves each along x-, y-, and

z-axes. Initially, the sample was isotropic with velocities in all three axes similar to each

other; however, as the differential stress increased, so did the percentage of anisotropy. This

Page 111: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

93

anisotropy, A, can be calculated as follows:

100max

minmax xV

VVA

−= Equation 4.1

where Vmax and Vmin are the maximum and minimum velocities. These anisotropy factors

and P-wave velocities provide the velocity structures of the sample during the experiment,

which are essential for the auto-picking algorithm of P-wave arrivals. Only a few velocity

surveys were performed after the failure of the sample. These surveys were taken either prior

to or after the completion of pore pressure oscillations, to avoid disturbing the oscillation

effects.

The plots of the evolution of elastic-wave velocities (Figure 4.11 and 4.12) show that a

maximum reduction in velocities occurred during the sample rupture and that the change

during the post-peak stage was relatively small compared to that of the pre-peak to rupture

stage. This significant reduction in velocities was due to the creation and coalescence of

microcracks during the failure stage. The velocity changes along the horizontal axes were

greater than that along the vertical axis during the rupture stage, which indicated that there

were more cracks oriented parallel to the vertical or sub-vertical axis than parallel to the

horizontal axis. This anisotropy in the velocity structure was taken into account for accurate

source locations (Equation 4.2). The details of the velocity models are listed in Appendix I.

Page 112: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

94

Figure 4.11 Typical velocity-time curve of Fontainebleau sandstone oscillating pore pressure experiment. The

development of transverse anisotropy indicates cracks were preferentially oriented in the direction parallel to

the vertical or sub-vertical axis (i.e., axial splitting).

Figure 4.12 Typical velocity-time curve of Darley Dale sandstone oscillating pore pressure experiment. There

was a greater reduction in velocity in the horizontal axis than in the vertical axis during the sample failure (i.e.,

transverse anisotropy).

Page 113: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

95

4.8 Results

4.8.1 Fontainebleau sandstone intact sample experiment (F5 and F8)

Two sets of cyclic pore pressure were applied in the Fontainebleau sandstone F5 and F8

experiments during the post-failure stage. These experiments simulated fluid-induced initial

seismicity by increasing pore pressure amplitude to a level higher than the previous

maximum. The applied pore pressure at upstream oscillated between 2.5 and 18 MPa at

2-minute periods, while the downstream pore pressure was measured. Figures 4.13 and 4.14

show the axial stress-time plots of the F5 and F8, respectively, while Figures 4.15 and 4.16

show the strain-time plots of the F5 and F8, respectively. During the post-peak period, the

fractured sample responded with an increase in seismicity, when the average pore pressure

exceeded the previous maximum (5 MPa) (Figure 4.17 and 4.18).

Figure 4.13 Stress-time curve of the F5 experiment under constant confining pressure of 20 MPa.

Page 114: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

96

Figure 4.14 Stress-time plot of the F8 experiment under constant confining pressure of 20 MPa.

Figure 4.15 Strain-time plot of the F5 experiment.

Page 115: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

97

Figure 4.16 Strain-time plot of the F8 experiment.

Figure 4.17 Seismic rate-time plot of the F5 experiment.

Page 116: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

98

Figure 4.18 Seismic rate-time plot of the F8 experiment.

The oscillating pore pressure of the two stages (the first stage with a peak pore pressure

of 18 MPa and the second stage with a peak pore pressure of 17 MPa) of the F5 and F8

experiments were plotted in Figures 4.19 and 4.20, respectively. Depending on the

naturally-fractured network of the sample, the downstream pore pressure of the F8

experiment tended to catch up with the applied upstream pore pressure faster than in the F5

experiment.

Page 117: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

99

Figure 4.19 Upstream and downstream pore pressures during the post-peak stage of the F5 experiment (Ying et

al., 2009).

Figure 4.20 Upstream and downstream pore pressures during the post-peak stage of the F8 experiment.

Page 118: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

100

With the aid of non-destructive acoustic techniques, seismicity that occurred during the

cyclic pore pressure could be located. Figures 4.21 and 4.22 show the AE events that

occurred during each pore pressure cycle during the main shock-aftershock and

foreshock-main shock-aftershock sequences, respectively. Figure 4.23 indicate the collective

events that occurred in the two sequences of the F5 experiment. The InSite software (ASC,

2007) allows the replay of events over time, from which no trends of event migration or

magnitude growth over time was observed. The pore pressure induced seismicity occurred

mostly along the pre-existing fractures and most of these events have relatively small

location magnitudes, ranging from -3 to -2. Similar results were found in the F8 experiment

and there were no trends of event migration found in either experiment.

Cycle 1 Cycle 2 Cycle 3 Cycle 4 Cycle 5

Figure 4.21 AEs occurred during the 5 cycles with sinusoidal pore pressure oscillated between 2.5 and 18 MPa

at 2-minute periods of the F5, mimicking a main shock-aftershock sequence.

Page 119: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

101

Cycle 1 Cycle 2 Cycle 3 Cycle 4 Cycle 5

Cycle 6 Cycle 7 Cycle 8 Cycle 9 Cycle 10

Cycle 11 Cycle 12 Cycle 13 Cycle 14 Cycle 15 Cycle 16

Figure 4.22 AEs occurred during the 16 cycles with sinusoidal pore pressure oscillated between 2.5 and 17 MPa

at 2-minute periods of the F5, mimicking a foreshock-main shock-aftershock sequence.

Page 120: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

102

Fractured sample Set 1 Set 2

Figure 4.23 AE locations of the F5 Fontainebleau sandstone experiment. Set 1: Cumulative seismic events

occurred during the 5 cycles of the main shock-aftershock sequence. Set 2: Cumulative seismic events occurred

during the 16 cycles of the foreshock-main shock-aftershock sequence. Each seismic event is represented by a

dot and the colour of the dots indicates different magnitude of events.

4.8.2 Fontainebleau sandstone control experiment (F7)

Figures 4.24 and 4.25 show the stress-time and strain-time plots of the F7 experiment,

while Figure 4.27 indicates the seismic rate-time plot of the experiment. The pore pressure

steps used in the experiment are illustrated in Figure 4.26.

Page 121: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

103

Figure 4.24 Stress-time plot of the F7 experiment under constant confining pressure of 20 MPa.

Figure 4.25 Strain-time plot of the F7 experiment.

Page 122: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

104

Figure 4.26 Upstream and downstream pore pressures during the post-peak stage of the F7 experiment. The

upstream and downstream pore pressures equilibrate almost instantaneously (Ying et al., 2009).

Figure 4.27 Seismic rate-time plot of the F7 experiment.

Page 123: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

105

4.8.3 Fontainebleau sandstone saw-cut experiment (F6)

The stress curve began to show ductile behaviour at about 16 MPa (Figure 4.28). The

strain rate was paused and subsequent pore pressure oscillation was applied (Figure 4.29).

However, due to the high permeability zone of the saw-cut, the upstream and downstream

pore pressures equilibrated almost instantaneously (Figure 4.30). The seismic rate-time plot

is illustrated in Figure 4.31.

Figure 4.28 Stress-time plot of the F6 experiment under constant confining pressure of 5 MPa.

Page 124: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

106

Figure 4.29 Strain-time plot of the F6 experiment.

Figure 4.30 Upstream and downstream pore pressures during the post-peak stage of the F6 experiment. The

upstream and downstream pore pressures equilibrated almost instantaneously.

Page 125: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

107

Figure 4.31. Seismic rate-time plot of the F6 experiment.

The source location, with location error of about 4 mm, showed that the events occurred

along the pre-cut during pore pressure oscillation and that the relative location magnitudes of

the events were small, ranging between -3.3 and -2.2 (Figure 4.32).

Figure 4.32 AE source location showing the events that occurred during each set of pore pressure cycles.

Page 126: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

108

4.8.4 Darley Dale sandstone experiment (DDS7)

Experiment DDS7 (Darley Dale sandstone), provides a good comparison to the

low-permeability Fontainebleau sandstone experiment, as Darley Dale sandstone has

relatively high porosity (~13%) and high initial permeability (~5.7 x 10-15 m2). Due to this

high permeability, phase shift could not be achieved even when the period of cycles was

reduced to half of that used in the Fontainebleau sandstone experiment (i.e. 1-minute period).

This suggests that the pore pressure of the entire system was almost at equilibrium. A total of

eight sets of pore pressure cycles were performed in this experiment. These sets involve pore

pressures cycled between a minimum of 2.5 MPa and a maximum of 7.5, 7.0, 10.0, 9.5, 12.5,

12.0, 18.0, and 17.0 MPa, respectively. The 7.5, 10.0, 12.5, and 18.0 MPa peak pore

pressures are the sets with increases in pore pressure that exceed the previous maximum. The

7.0, 9.5, 12.0, and 17.0 MPa pressure sets are those with lowered peaks as compared to the

preceding sets, i.e., at approximately 95 % of the previous maximum. These sets were used

for investigating fault reactivation and foreshock–main shock–aftershock sequences. Figure

4.35 shows the period in which the pore pressure oscillations were peaked at 18.0 and 17.0

MPa. This partial plot of the pore pressure oscillating stage indicates that the upstream and

downstream pore pressures equilibrated almost instantaneously. The results indicate that an

existing fault was activated at the four step increase in pore pressure peak that exceeded the

previous applied maximum. The activations were accompanied by significant axial strain

change. The seismic rate in the subsequent oscillations reduced more rapidly than that in the

Fontainebleau sandstone experiments (F5 and F8). Although there was occasional seismicity

during the reduced peak pore pressure sets, no significant axial strain change was observed.

As a result, no reactivation of faults was identified during these periods. The stress-time and

Page 127: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

109

strain-time plots (Figures 4.33 and 4.34) indicate the four stages of fault activation, when the

pore pressure exceeds the previous applied maximum. These four stages of activation were

accompanied by an increase in AEs, as indicated in Figure 4.36.

Figure 4.33 Stress-time plot of the DDS7 experiment under constant confining pressure of 20 MPa.

Figure 4.34 Strain-time plot of the DDS7 experiment.

Page 128: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

110

Figure 4.35 Upstream and downstream pore pressures during the post-peak stage of the DDS7 experiment.

Partial plot of the oscillating pore pressure stage, showing the upstream and downstream pore pressures

equilibrate almost instantaneously.

Figure 4.36 Seismic rate-time plot of the DDS7 experiment.

Page 129: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

111

The AEs generated during the post-peak cyclic pore pressure regime occurred along the

pre-existing fault. In this experiment, no trends of event migration or magnitude increase

with time were observed. Figure 4.37 shows the collective seismic events that occurred

during the eight cyclic pore pressures sets.

Figure 4.37 AE source location showing the events that occurred during each set of pore pressure cycles. Sets 1

to 8 correspond to pore pressure cycles with peak pressures at 7.5, 7.0, 10.0, 9.5, 12.5, 12.0, 18.0, and 17.0 MPa,

respectively. Most of the events occurred during the sets with pore pressure peaks exceeding the previous

maximum. Few events occurred during the sets with pore pressure peaks lower than the previous maximum.

Page 130: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

112

Chapter 5 Analysis and Discussion

5.1 Experiment F5

By plotting the strain, seismic rate, and pore pressure oscillation on the same graph

(Figure 5.1), a clear picture of the seismic response with respect to the oscillating pore

pressure can be observed. During the first set of applied pore pressures with oscillation

between 2.5 and 18 MPa, there was a corresponding cyclic axial strain change (Figure 5.1)

with the cumulative axial strain change during the 5 cycles equivalent to a sample shortening

of ~0.019 mm. The strain changes in the first three cycles were greater than the last two

cycles. In Roeloffs et al. (1979), where cyclic pore pressure experiment was carried out on

saw-cut sandstone samples, a greater sliding along the pre-cut was also found to be greater

during the first few cycles than the subsequent cycles.

The aftershock peak seismicity of the cycles decayed according to the modified Omori

law (Utsu, 1961) with p-value about 0.0039, but at a slower rate compared to tectonic events.

This is a common characteristic in fluid-induced seismicity (Gupta, 1972 & 2005). After

main slip 1 in the aftershock sequence, the seismic rate continued in the subsequent cycles,

but reduced gradually in cycles 2 to 5. Due to the relatively low permeability of the fractured

sample (~5.7 x 10-17

m2), there was a phase shift between the applied pore pressures and the

responding pressures downstream. This phase shift decreased with the increase in the number

of oscillations (Figure 5.2). This is due to the permeability increase due to cumulative

damage. The permeability measured after the completion of the cycles was ~5.0 x 10-16

m2,

which is an increase of one order of magnitude, as compared to the initial permeability of

~1.3 x 10-17

m2.

Page 131: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

113

After the first set of pore pressure cycles, a period of quiescence was introduced by

reducing the pore pressure to 2.5 MPa until the seismic rate fell to a background level. A

second set of pore pressures was then applied, oscillating between 2.5 and 17 MPa, with a

period of 2 minutes. The lower pore pressure amplitude set was introduced, with an objective

to investigate whether pore pressure oscillation can reactivate a fault and generate a

foreshock–main shock–aftershock sequence. It is observed that during this set of pore

pressure oscillation, the fault was reactivated and the seismic rate intensified after the 6th

cycle, when the average peak pore pressure amplitude exceeded the previous maximum

attained (Figure 5.2). During each pore pressure cycle, the corresponding axial strain change

indicated a relative shortening and extension of the sample, which suggests a back-and-forth

sliding movement along the pre-existing fault. This cumulative axial strain change

corresponds to a total of ~0.061 mm vertical movement (comprised of ~0.021 mm extension

movement and ~0.04 mm shortening movement). The amount of movement due to each pore

pressure cycle was small, but hysteretic (i.e., when the pore pressure is reduced to the

minimum amplitude of each cycle, the corresponding axial strain change cannot be

recoverable completely).

Page 132: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

114

Figure 5.1 The seismic rate, pore pressure cycles, and the axial strain change of the F5 experiment. The first 5

cycles simulate an aftershock sequence in which pore pressure was oscillated sinusoidally between 18 and 2.5

MPa, with a period of 2 minutes. The subsequent 16 cycles simulate a foreshock–main shock–aftershock

sequence, when pore pressure was oscillated between 17 and 2.5 MPa. Fault reactivation intensified after the

6th cycle, when the change in strain increased and the main slip 2 occurred at the 11th cycle. Downstream pore

pressure cycles exhibited phase shifts and remained at a transient state in response to the applied upstream pore

pressure cycles (Ying et al., 2009).

Page 133: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

115

Figure 5.2 Phase shift and average pore pressure at each cycle for the F5 experiment. The phase shift decreased

with the increasing oscillation number. The reactivation of faults begins when the average pore pressure

exceeds the previous maximum, i.e., beyond the 6th cycle. A main slip occurs at the 11th cycle, with a reduction

in phase shift.

The increase in average peak pressure was accompanied by a reduction in the phase

shift (Figure 5.2). The largest slip in the foreshock sequence (denoted as main slip 2 in

Figure 5.1) occurred in the 11th cycle, which was denoted with an anomalous reduction in

phase shift. After this slip, the period of phase shift rebounded to the previous trend, with a

reduction in the seismic rate in the aftershock sequence (cycles 12 to 16). The

Page 134: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

116

foreshock–aftershock sequence exhibits a Mogi (1963) Type II model (Figure 2.14), which

reflects the heterogeneity of the sample after the formation of ‘natural’ fractures during the

rupture. This heterogeneity causes an uneven distribution of stress during the transient state

of pore pressure, which is reflected by the irregular distribution of induced seismicity, as

indicated by the AE locations.

It is inferred that the number of pore pressure cycles required for reactivation depends

on the fracture network which controls the mechanical properties of the fractured sample,

which in turn control the development of pore pressures within the system. As the average

pore pressure (average of upstream and downstream) reached the previous maximum average,

the seismic rate began to accelerate and subsequently developed into the reactivation slip (2),

accompanied by a significant axial strain increase (Figure 5.1). The downstream pore

pressure oscillation and the periods of phase shift are good indicators of whether the system

has reached steady-state or remained transient. If the downstream pore pressure peaks

continued to grow, it is likely that faults would be reactivated and further main slips might

occur with an increase in the number of pore pressure cycles. The results suggest that cyclic

pore pressure can induce protracted seismicity during the period in which pore pressure

amplitudes are lower than the previous applied maximum. This effect has been observed in

the Aswan reservoir when the seismicity continued and enhanced between 1982 and 1987,

during which the water level of the reservoir was reduced.

The vertical distances of the hypocentres are measured from the upstream of the

samples (Figure 5.3). The population distribution charts indicate that the seismic events

Page 135: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

117

occurred during the pore pressure oscillation distributed over a wide range; however, there

were fewer events located at the top third of the downstream of the sample (Figure 5.4),

possibly affected by the permeability of the fractured sample.

Figure 5.3 Showing the reference distance for measurement of hypocenter locations.

Figure 5.4 Population distribution of the seismic events occurred during pore pressure oscillations.

125

0

Upstream

Downstream

Page 136: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

118

5.1.1 b-value analysis

The AE samples were grouped into four different periods: i) before the failure of sample,

ii) aftershock sequence of main slip 1, iii) foreshock sequence of main slip 2 and, iv)

aftershock sequence of main slip 2. For each period, log N values were calculated for

location magnitudes ranging from -2.0 to -2.9 at 0.1 intervals. The equation for obtaining

b-values is indicated in Equation 3.8. The b-values were obtained from the slope of the line

measured from the linear descending portion of the frequency-magnitude plots. In the

experiment, there was a slight variation between the b-values. The b-value analyses of the

two aftershock sequences are very similar (< 4% difference). However, the b-value of the

foreshock sequence is ~1.93, which is about 22% lower than that of the aftershock sequence

(Table 5.1, Figure 5.5). This indicates that there are more large-magnitude events occurring

during the foreshock period than during the aftershock period.

Table 5.1 b-values during different periods of the F5 experiment.

Period b-value

Before the failure of sample 2.069

Aftershock sequence of main slip 1 2.385

Foreshock sequence of main slip 2 1.927

Aftershock sequence of main slip 2 2.474

Page 137: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

119

Figure 5.5 Fontainebleau sandstone experiment: a) b-value analysis for the formation of fault; b) b-value

analysis for the aftershock and foreshock sequences.

5.1.2 AE source mechanism

Typical AEs generated during the pore pressure oscillation period were analysed using

focal mechanism solutions. Representative AE data with large magnitudes were chosen for

the analysis. These events have high signal-to-noise ratios which provide high confidence

mechanism solutions. The results indicate that shear or a complex mechanism dominated

during the oscillation pore pressure stage (Figure 5.6).

Page 138: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

120

Figure 5.6 Focal mechanism solutions indicating the failure mechanisms during oscillating pore pressure.

5.2 Experiment F8

The results of the F8 experiment are similar to those of F5 that seismic sequences can be

activated during the period when pore pressure peaks were lowered to 95% of the previous

maximum. However, due to the difference in the mechanical properties of the

naturally-fractured samples, the foreshock development was not as obvious as that showed in

the F5 experiment (Figure 5.1). The main slip caused by the increase in pore pressure

generated significant numbers of AEs (Figure 5.7), which reduced rapidly from about 200

events in the first cycle to about 30 events in the subsequent cycles. The decay is according

to modified Omori law (Utsu, 1961), with p-value about 0.0036. During the foreshock

sequence, the number of AEs gradually increased in the first 4 cycles. After cycle 4, the AE

rate of each cycle was similar and persisted during the aftershock sequence. Starting from the

Page 139: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

121

cycle 4 of the foreshock sequence, the downstream pore pressure reached a ‘steady-state’, i.e.,

the peak pore pressures remained unchanged. It was observed that the seismicity

development was influenced by the downstream pore pressure or the average pore pressure

of the system, instead of the upstream pore pressure. However, in the field, often only the

upstream pore pressures are readily available. The response of the downstream pore pressure

is often unknown, unless monitoring wells are installed in the downstream.

In the F8 experiment the development of the downstream pore pressure was not as rapid

as that of the F5 experiment (Figure 5.1). This, in turn, influenced the strain development, as

well as the seismic rate. This difference was likely caused by the variation in the mechanical

properties between the F5 and F8 samples, when the fracture networks of the two

naturally-fractured samples were not identical.

Page 140: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

122

Figure 5.7 The seismic rate, pore pressure cycles, and axial strain change of the F8 experiment. The first 5

cycles simulated an aftershock sequence in which pore pressure was oscillated sinusoidally between 18 and 2.5

MPa, with a period of 2 minutes. The subsequent 16 cycles simulated a foreshock–main shock–aftershock

sequence when pore pressure was oscillated between 17 and 2.5 MPa. The seismic rate developed slightly in the

first 4 cycles of the foreshock sequence; however, due to the limited increase in downstream pore pressure, the

corresponding strain did not change significantly.

5.3 Control experiment on Fontainebleau sandstone

The strain, seismic rate, and pore pressure change plots are indicated in Figure 5.8. The

plot shows that without the cyclic effect, seismicity due to pore pressure enhancement is

Page 141: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

123

reduced, as compared to the F5 experiment. Aftershocks decay according to the Omori law

(Utsu, 1961) with p-value about 0.0108, about 2.8-3.0 times faster than that of the

aftershocks in the F5 and F8 experiments. This suggests that the rate of reduction in

aftershock sequence was reduced due to oscillating pore pressure. The cumulative number of

seismic events was significantly less than that recorded in the cyclic pore pressure

experiment. The seismicity dissipated in ~6.6 minutes in the control experiment, while the

seismicity continued throughout the 5 cycles of 18 MPa peak pressure in the F5 and F8

cyclic pore pressure experiments. In addition, no reactivation of fault was observed in the

control experiment when the pore pressure is at ~95 % of the previous maximum (i.e., 17

MPa). However, in the control experiment, seismicity was reactivated during the periods

when a significant reduction of pore pressure took place (i.e., reduced to 2.5 MPa). The two

periods of constant pore pressure could be viewed as two trapezoidal pore pressure cycles.

During each decrease and/or increase in pore pressure, seismicity was induced. The

experimental results confirm that cyclic pore pressure has greater influence on seismicity

than a constant step increase of pore pressures.

Page 142: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

124

Figure 5.8 The control experiment indicates pore pressure steps at 18 MPa for 10 minutes and 17 MPa for 32

minutes. The seismicity due to enhancement in pore pressure reduced rapidly. In addition, there was no

reactivation of fault when the pore pressure was at ~95 % of the previous maximum (17 MPa). Seismicity was

reactivated during the reduction of pore pressure (Ying et al., 2009).

Due to the pore pressure step change to 18 MPa for 10 minutes and the subsequent

reduction back to 2.5 MPa, the corresponding shortening of the sample totalled ~0.019 mm,

which is comparable to the total vertical movement measured during the 5 cycles of 18 MPa

peaks in the F5 oscillating pore pressure experiment (Table 5.2). However, during the

subsequent 17 MPa step change, the axial strain change in the control experiment

corresponded to 0.006 mm vertical movement, which is much less than that measured in the

Page 143: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

125

cyclic pore pressure experiment. This suggests that oscillating pore pressure can cause

oscillating movements along the fault, which can continue over a long period of time. The

cumulative wearing of asperities not only reduce the roughness of the fault, but also prolongs

the period of seismicity. This phenomenon can be seen as a ‘slow’ mechanism of oscillating

pore pressure, characterised by a protracted but discontinuous seismicity over a long period,

rather than continuous seismicity within a relatively short period of time, as per a typical

seismic sequence due to a pore pressure step increase. This slow seismicity is also signified

by asperity-controlled stick-slip-like movements. Implications of such phenomenon may be

seen in natural slow earthquakes occur at the intermediate crustal levels of subduction zones

such as SW Japan (Obara, 2002; Ariyoshi et al., 2009) and Cascadia, Chile (Rogers and

Dragert, 2007), in which the subduction zones dip shallowly and are accompanied by fluid

flow and related tremors with relatively low frequencies (typically in 1-5 Hz band).

Table 5.2 Corresponding vertical movements inferred by axial strain measurements.

Pore

pressure

peaks

F5 cyclic pore pressure experiment

(cumulative relative movement due

to oscillating pore pressure)

Control experiment (cumulative

relative movement due to step

change in pore pressure)

18 MPa ~0.019 mm ~0.019 mm

17 MPa ~0.061 mm ~0.006 mm

As suggested by the comparison between the control experiment (F7) and the F5

experiment, cyclic pore pressure can cause a longer period of progressive sliding and

seismicity, as compared to a step increase in pore pressure. Most of the seismic events due to

cyclic pore pressure occurred over a longer period of time, while for the pore pressure step

increase the seismicity reduced rapidly to its background level within a shorter period of time.

Page 144: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

126

This effect of oscillating pore pressure versus step increase in pore pressure can be illustrated

by Figure 5.9.

Figure 5.9 Schematic diagram indicating the reduction of shear strength with cycles a to d. Left: showing the

reduction is progressive with the number of cycles. The time required for the strength to get to stage d is longer

because the duration of each cycle peak is short and requires several cycles to accumulate the same effect as

that of a step change in pore pressure. Right: showing a schematic result of a step increase in pore pressure

where the reduction in the friction angle occurs within a shorter period.

5.4 Darley Dale sandstone experiment

Figure 5.10 shows the strain and seismic rate during the oscillating pore pressure stage.

There are 4 activation and aftershock sequences in which the pore pressure exceeded the

previous maximum. The pattern of the four seismic sequences corresponds to the Type I

Mogi model (1963), which suggests that the structure of the sample is homogeneous. Figure

5.11 indicates the typical set of the main shock-aftershock sequence; there was no

reactivation of fault when the pore pressure peaks were reduced to ~95% of the previous

maximum. The aftershocks decayed according to the modified Omori law (Utsu, 1961), with

p-value about 0.158, which is a much faster rate than that of the F5 and F8 experiments. The

permeability of the fractured sample prior to pore pressure oscillation was relatively high,

Page 145: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

127

~3.6 x 10-13

m2, which is two orders of magnitude increase as compared to the initial

permeability of ~5.7 x 10-15

m2.

Figure 5.10 Induced seismicity due to oscillatory pore pressures in Daley Dale sandstone. Eight sets of pore

pressure cycles were performed: with pore pressures cycled between a minimum of 2.5 MPa and a maximum of

7.5, 7.0, 10.0, 9.5, 12.5, 12.0, 18.0 and 17.0 MPa, respectively. The 7.5, 10.0, 12.5 and 18.0 MPa peak pore

pressures sets simulated activation of fault, with increase in pore pressure exceeding the previous maximum,

while the rest were at ~95% of the previous maximum, aiming to investigate reactivation of fault (Note: For

clarity, only the peaks and troughs of the cyclic pore pressures were plotted for the sinusoidal pore pressure

cycles). The upstream and downstream pore pressure equilibrated almost instantaneously (Ying et al., 2009).

Page 146: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

128

Figure 5.11 Showing seismic rate, pore pressure cycles and axial strain change of the Darley Dale sandstone

experiment, with pore pressure oscillated between a 18 and 2.5 MPa simulated aftershock sequence. The

subsequent cycles oscillated between 17 and 2.5 MPa, accompanied by insignificant seismicity, which implies

that the pre-existing fault was not reactivated when the peak pore pressure was lower than the previous

maximum.

5.4.1 b-values

There are more seismic events during the peak pore pressure enhancement periods

(aftershock sequence) (sets 1, 3, 5, and 7 in Figure 5.10) than during the peak pore pressure

reduced periods (sets 2, 4, 6, 8 in Figure 5.10). The frequency-magnitude analyses of the 4

aftershock sequences indicate that there were more low magnitude events than high

magnitude events and that the range of magnitude varied narrowly between -2.5 and -2.9.

Page 147: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

129

This magnitude range is lower than those observed during the formation of the fault (between

-2.2 and -2.9) (Table 5.3, Figure 5.12). In addition, the seismic events that occurred in the

Darley Dale sandstone were generally of lower magnitude than those recorded in the

Fontainebleau sandstone experiment. The larger number of low magnitude events than high

magnitude events cause the b-value to be lower than those obtained in the Fontainebleau

sandstone experiment. This difference is due to the material property difference between the

two sandstones. Darley Dale sandstone behaves more ductile while the Fontainebleau

sandstone is more brittle in nature. According to Scholz (1968), it was found that the

b-values are much higher in ductile rock than that observed in brittle rock.

Figure 5.12 b-value analysis of the Darley Dale sandstone experiment: a) During the formation of the fault; b)

during the aftershock sequences.

Page 148: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

130

Table 5.3 b-values during different periods of the DDS7 experiment.

Period b-value

Before the failure of sample 4.48

Aftershock sequence 1 4.50

Aftershock sequence 2 5.20

Aftershock sequence 3 5.44

Aftershock sequence 4 4.06

The population distribution of the seismic events measured along the vertical distance

indicates that most of the events occurred in the middle portion of the sample, i.e., where the

main fault was located (Figure 5.13), which is also influenced by the event accuracy.

Figure 5.13 The distribution of seismic events that occurred during the oscillation of pore pressure.

5.4.2 AE source mechanism

Typical AEs generated during the oscillating pore pressure periods were analysed using

Page 149: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

131

focal mechanism solutions. Representative AE data with large magnitudes were chosen for

the analysis. These events have high signal-to-noise ratios which provide high confidence

solutions. The results show that shear or complex mechanisms are the main mechanisms

during the pore pressure oscillation period (Figure 5.14). The representative mechanisms

occurring during the increase in pore pressure typically have compression (white) in the

central portion of the focal sphere and dilation (shaded) on both sides of the sphere, indicated

by sphere 2, 3, 4 and 6, which are related to crack opening events; while the representative

mechanisms occurred during the decrease in pore pressure typically have dilation (shaded) in

the central portion of the focal sphere and compression (white) on both sides of the sphere,

indicate by sphere 1 and 5 (Figure 5.14), which are related to crack closing events. This was

observed in the Darley Dale experiment due to the relatively simple fault zone and the

absence of phase shifts between the upstream and downstream pore pressures.

Figure 5.14 Focal mechanism solutions of the Darley Dale sandstone experiment, indicating that the dominant

failure mechanism during the cyclic pore pressure stage is shear or complex.

Page 150: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

132

5.5 Comparison between Fontainebleau sandstone and Darley Dale sandstone

experiments

A comparison between the two cyclic pore pressure experiments is shown in Table 5.4.

The results show that the mechanical properties of the sample influence the decay of

aftershocks, fault reactivation, as well as the development of foreshock and main slip events.

When faults are reactivated, the progressive swarms develop into a main slip after a number

of pore pressure oscillations. However, this reactivation requires a faulted sample with a

heterogeneous structure, as well as low permeability, which in turn allows a phase shift to

develop.

Page 151: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

133

Table 5.4 Comparison of the results of the two cyclic pore pressure experiments.

Fontainebleau Sandstone

(relatively low permeability,

~1.3 x 10-17

m2)

Darley Dale Sandstone

(relatively high permeability

~5.7 x 10-15

m2)

Phase Shift

Between Upstream

and Downstream

Pore Pressure

Present, reduction in phase shift with

increase in the number of cycles,

sharp reduction during the

occurrence of main slip

Downstream pore pressure

caught up with the applied

upstream pore pressure almost

instantaneously

Steady-state Pore

Pressure

Oscillations

The downstream pore pressure peaks

gradually increase, indicating that the

system was at transient state

Equilibrium of the upstream and

downstream pore pressure cycles

suggests steady-state oscillation

was achieved

Characteristics of

Aftershock

Sequence

Reduction in seismic rate is relatively

gradual, thus seismicity can continue

for a longer period

Reduction in seismic rate is

relatively rapid

Fault Reactivation Marked by a significant axial strain

change and increase in seismic rate

No distinguishable axial strain

change; occasional increase in

seismic rate

Characteristics of

Foreshock– Main

Shock–Aftershock

Sequence

Foreshocks gradually developed and

when the average pore pressure

exceeds the previous maximum, the

seismic rate increased and developed

until a main slip occurred, followed

by the aftershock sequence

No observable sequence

5.6 Saw-cut experiment

The main purpose of the saw-cut experiment was to identify any trends of event

migration. However, due to the formation of the high permeability continuous fracture, the

upstream and downstream pore pressure equilibrated almost instantaneously and

pore-pressure-induced AEs occurred along the fractured path almost without delays.

Migration trends were therefore not observed. Due to the low stress level, relatively fewer

Page 152: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

134

AE events were recorded during the oscillating pore pressure period. When the pore pressure

oscillated between 3.25 to 3.75 MPa, the saw-cut sample started to creep and underwent

stable sliding. This is signified by the measured strain indicated in Figure 5.15.

Figure 5.15 The F6 experiment indicated stable sliding along the saw-cut at a later stage, when the average pore

pressure was about 3.5 MPa.

Page 153: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

135

5.6.1 Migration trends

Some researchers (Talwani, 1997; Talwani et al., 2007; do Nascimento et al., 2004; Rao

and Singh, 2008) suggest that protracted RIS has a migration trend, as seismic events with

higher magnitudes and deeper hypocentres were observed at a later stage after the reservoir

impoundment. This migration trend may be possible to recognize in the field because the

diffused distance is longer. However, there were no such patterns found in the laboratory

experiments. This can be due to two reasons. The first is the limited size of the sample, in

which the entire sample can be viewed as a single fractured zone and so any pore pressure

diffusion effect can occur within a relatively short period of time, which inhibits the

observation of the migration of events and trends in the event magnitude change. However,

in the field, migration of events can be observed as hypocentres occurred at deeper depth and

further distance from the initial seismic region due to pore-fluid diffusion (do Nascimento et

al., 2004; Rao and Singh, 2008). The second reason is that the acoustic emissions in the

Fontainebleau sandstone experiments were located in the irregular fractured network. The

heterogeneity of the fractured zone and the progressive shearing effect of the cyclic pore

pressure both contribute to flow anisotropy, which also inhibits the observation of migration

trends in the sample.

5.7 Post-experimental analysis

There were two post-experimental analyses carried out to provide additional

information regarding the characteristics of the tested samples that underwent cyclic pore

pressures. These analyses included X-ray micro-computed tomography analysis and optical

microscopy analysis. These analyses were carried out on the Fontainebleau (F5) and Darley

Page 154: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

136

Dale (DDS7) sandstone samples, which were used in the experiments described in Chapter 5.

5.7.1 X-ray micro-computed tomography analysis

In order to understand the structure of the faulted samples, high-resolution X-ray

micro-computed tomography (CT) was carried out for both the Fontainebleau and Darley

Dale sandstone samples. Prior to scanning, the samples were preserved with epoxy to avoid

movement of faults during handling. The samples were then trimmed into a smaller size to

suit the requirement of the high-resolution SkyScan 1172 scanner. The most representative

middle portions of the samples were retained for the micro-CT scan. The images produced

by the scanner correspond to a series of 34.89 µm thick slices throughout the specimens. The

reconstructed images of the Fontainebleau sandstone sample indicate that the faulting system

is comprised of heterogeneous shear fractures and axial splitting, while that of the Darley

Dale sandstone sample consists of a single shear fracture. Figure 5.16 indicates the typical

micro-CT images of the Fontainebleau and Darley Dale sandstones. These images align with

the fact that slightly heterogeneous medium (F5) has a number of foreshocks preceding the

main shock, while homogeneous medium (DDS7) is characterised by seismicity with no

foreshocks (Mogi, 1963). Slightly heterogeneous medium is defined by the presence of

heterogeneous type of fractures within in a homogeneous sample.

Page 155: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

137

a) b)

Figure 5.16 Typical X-ray micro-CT images along the vertical axis. a) Showing a complex faulted system with

shear fractures and axial splitting within the naturally fractured Fontainebleau sandstone sample; b) showing

that a shear fracture was formed in the Darley Dale sandstone.

The 489 slices of xy-plane scanned images of each specimen were changed to binary

colour for the calculation of normalised crack area with depth (z-direction), i.e., normalised

by the smallest crack area among the 489 slices. This normalised area indicates the relative

ease of fluid flow along the vertical axis during the oscillation of pore pressure (Figure 5.17).

The flow areas or crack areas changed significantly along the vertical axis, which suggests

the variation of local permeability and local transient pore pressure development along the

path. This increase in pore pressures reduced the local contact stress and release as an AE or

stress redistribution within the material.

Page 156: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

138

Figure 5.17 Normalised crack area is calculated as crack area/minimum crack area of the plots. This indicates

the change in flow path in the z-direction. a) For Fontainebleau sandstone; b) for Darley Dale sandstone.

5.7.2 Optical microscopy

During initial rupture, the grains along the fracture zone were crushed into smaller sizes.

The subsequent pore fluid increase may transiently reduce grain-contact stresses to zero, and

induce failure along the existing fault (Iverson and LaHusen, 1989). This fluctuation can

propagate outside the shear zone and promotes shear-zone growth. As the pore fluid

oscillation continues, cumulative damage due to progressive shearing can occur in the

fractured zone. This shearing involves the grinding and over-riding of grains along the

fractured surface, which may extend the shear zone outward. The cumulative axial strain

change during the pore pressure oscillations can be used to infer the relative grinding

movement along the fractured zone. As the number of cycles increase, more grains are worn

off or ground into smaller sizes, which exposes more surface area for wetting, and enhances

the lubricating effect. The wearing of asperities eventually reduces the friction coefficient

Page 157: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

139

and, hence, leads to protracted slips. The amount of progressive wearing due to cyclic pore

pressure cannot be distinguished by optical microscopy. However, optical microscopy can

provide an overall indication of the shearing that has occurred in the fractured zone, due to

the initial rupture and subsequent pore pressure oscillations. Optical microscopy analysis was

carried out on the tested Fontainebleau (F5) and Darley Dale (DDS7) sandstones after the

completion of the X-ray micro-CT scan. Three thin sections along the xy-plane, xz-plane,

and yz-plane were taken from each tested sample. The comparison between the thin sections

of the two types of sandstone can provide some indication of the corresponding differences

in seismic responses.

Fontainebleau sandstone is composed of more than 98% quartz, which is a brittle

mineral. Due to the hardness contrast between the quartz grains and the adjacent softer gouge

material, abrasive wearing could be the dominant process during progressive shearing.

During abrasive wearing, the harder material can plough through the softer material (Scholz,

2002). Optical microscopy indicates that large loose-wear particles with angular forms were

embedded in the gouge material (Figure 5.18b), while for the Darley Dale sandstone, which

is more ductile in nature, the wearing could be adhesive, in which junctions shear off part of

the adjoining asperity, resulting in transfer of material from one surface to the other (Scholz,

2002; Rabinowicz, 1965). Deposits of ground material were found in the fractured zone of

the Darley Dale sandstone (Figures 5.19 and 5.20). The optical microscopy of both

sandstones indicates larger grain sizes further away from the shear zone (Figures 5.18 to

5.20). However, the thin sections of Fontainebleau sandstone exhibit two types of

phenomena. It was found that no significant shearing occurred in areas where the crack

Page 158: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

140

widths were relatively wide (Figure 5.18c), while in other areas excessive shearing has

occurred and the grain size was reduced significantly (Figure 5.18b). This might cause

permeability contrast between the different zones. In regions with relatively large cracks,

pore pressure might dissipate at a faster rate than that within the sheared zone.

The shear zone of the Darley Dale sandstone was relatively loosely packed, which

indicates a relatively high permeability and allowed pore pressure dissipation. This

permeability contrast between the two sandstones may explain why almost instantaneous

pore pressure equilibrium was established in the Darley Dale sandstone during the pore

pressure oscillations, while the phase shift between the upstream and downstream pore

pressure persisted in the Fontainebleau sandstone.

Page 159: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

141

(a) (b) (c)

Figure 5.18 Optical microscopic views of the Fontainebleau sandstone that has been subjected to oscillating

pore pressures. a) Larger grain sizes can be found in areas further away from the sheared zone; b) gouge

material found within the shear zone that has undergone grinding and progressive shearing. The grain size of

the gouge material was reduced significantly and was closely packed, influencing the permeability of the

sample. It is noted that larger grains were embedded in the sheared zone; c) areas with relatively larger cracks

that enhance the local permeability. Consequently, less gouge material was found, which implies that less

progressive shearing has occurred in the area.

Page 160: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

142

Figure 5.19 Thin section of the Darley Dale sandstone tested sample. Left: Typical grain size found in the area

further away from the shear zone. Right: Grains in the sheared zone have significantly smaller grain sizes due to

shearing. There is no embedment of larger, angular grains in the sheared zone and the shear zone is not as

closely packed as compared to that of the Fontainebleau sandstone thin sections.

Page 161: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

143

Figure 5.20 Optical microscopy of the Darley Dale sandstone with sheared material in the fractured zone. The

grain size of the sheared material is reduced significantly. The sheared zone is not as closely packed as that of

the Fontainebleau sandstone, suggesting that the permeability of the sheared zone is relatively high and can

permit almost instantaneous equilibrium in pore pressure between the upstream and downstream.

Page 162: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

144

The relative axial strain change implies that oscillating pore pressure can progressively

wear off asperities at the pre-existing fractured surfaces, which in turn reduce the friction

coefficient and the frictional strength of the material and, subsequently, cause the fractured

sample to be more susceptible to sliding or shearing. As the shear displacement due to each

cycle of pore pressure is relatively small, a large number of cycles is required to wear or

smooth the fractured surface. This increase in shear displacement along the existing fractures

can cause the sample to become heterogeneous and anisotropic (Yeo et al., 1998),

subsequently altering the pore network and easing pore pressure development at different

areas. This alteration has no specific trend and, therefore, the seismic events that occurred

during the pore pressure oscillation have no identifiable migration patterns. This oscillating

pore pressure induced protracted seismicity is expected to last for a long period of time and

the magnitude of the seismic events are often relatively small.

5.8 Conclusions

Many researchers have attributed protracted RIS to the pore diffusion effect. This may

explain the phenomenon of delayed RIS and the migration of RIS to deeper crust over time.

However, in many RIS cases, protracted seismic events at shallow depths continued to

dominate after a prolonged period of time. This suggests that effect due to oscillating pore

pressure other than pore diffusion may be taking place to allow continuous shallow crustal

seismicity.

The experimental results indicate that cyclic pore pressure can cause a progressive

shearing effect, which was validated by the AE source mechanism and implied by the small

Page 163: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

145

strain change due to each cycle. The shearing and corresponding seismicity due to cyclic

pore pressure can last for a longer period of time, as compared to that of a step increase in

pore pressure. This shearing and wearing of asperities at the fractured surface can influence

the flow anisotropy and, hence, affect the pore pressure distribution in the subsequent cycles,

which then induce seismicity at different regions along the existing fractured surface.

The results illustrate that initial seismicity can be activated when the peak pore pressure

exceeds the previous maximum experienced by the system. The decline in the number of

aftershocks occurred more rapidly for the Darley Dale sandstone than for the Fontainebleau

sandstone, which was influenced by the mechanical properties of the samples.

The comparison between the control experiment and that of the cyclic pore pressure

experiment on the Fontainebleau sandstone indicates that pore pressure oscillations can

significantly extend the period of seismicity, i.e., can induce protracted seismicity. The

reactivation of fault and protracted seismicity were found in the heterogeneously-faulted

Fontainebleau sandstone, but not in the Darley Dale sandstone experiment. The relatively

low permeability Fontainebleau sandstone exhibited phase shifts and transient-state pore

pressure cycles, which then gradually developed and reactivated the pre-existing faults.

However, no foreshock sequence was developed in the relatively high-permeability Darley

Dale sandstone, which allowed an equilibrium state of pore pressure oscillation to be reached

almost instantaneously. This finding is in line with the observation that homogeneous media

are characterised by seismicity with no foreshocks, while slightly heterogeneous media are

characterised by a number of foreshocks preceding the main shock. The experimental results

Page 164: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

146

suggest that two main factors influence the protracted seismic response of the samples: i) the

presence of pore pressure oscillation, and ii) the mechanical properties and the heterogeneity

of the fault zone.

The effect of cyclic pore pressure loading can reduce the effective stress and cause

sliding between the fracture surfaces. During each sliding, the amount of movement is small,

as implied by the relatively small changes in axial strain. The movement of each cycle is

hysteretic, as shearing along the interlocked asperities at the fractured surfaces is irreversible.

The progressive changes in phase shifts between the upstream and downstream pore

pressures, as well as the transient changes in downstream pore pressure, can provide useful

information for mitigation measures of reservoir-induced seismicity caused by the

exploitation of underground fluid reservoirs and cyclic variation in the water levels of surface

reservoirs. When a cyclic pore pressure system is at equilibrium, the phase shift between the

upstream and downstream would be steady. In addition, the downstream pore pressure cycles

would generate steady-state oscillations. Therefore, by monitoring the phase shifts and

downstream pore pressure variations in the field, the progression of pore-pressure-induced

seismicity and the state of the pore pressure in the substratum can be monitored. This

provides some information about the susceptibility of the area to protracted RIS and, hence,

appropriate mitigation measures can be carried out when necessary. When controlling the

upstream water levels or pore pressures as a means of reducing reservoir-induced seismicity,

it is essential that the extent of the reduction be large enough to lower the downstream pore

pressures and that the rate of the reduction must be gradual to prevent seismicity induced by

Page 165: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

147

rapid reduction. It should be noted that the response of the downstream pore pressure may

take months or longer, depending on the mechanical properties of the substratum, as well as

the natural seasonal variations.

As the wearing damages accumulate with the increase in the number of pore pressure

cycles, the friction coefficient reduces and, subsequently, induces protracted seismicity. The

X-ray micro-CT scans indicate irregular flow paths within the samples. Optical microscopy

shows a contrast in permeability in different areas of the Fontainebleau sandstone specimen,

which suggests that during the experiment the development of pore pressures might be

different in different regions. In some regions where grains were closely packed together,

progressive shearing could be developed. The permeability contrast between the Darley Dale

and Fontainebleau sandstones can explain the difference in their pore pressure and seismic

responses. In the case of high-permeability Darley Dale sandstone, almost instantaneous

equilibrium could be established between the upstream and downstream pore pressures and

no protracted seismicity or reactivation of fault occurred due to this equilibrium state. For the

low permeability Fontainebleau sandstone, phase shifts between the upstream and

downstream pore pressures persisted and, due to the gradual pore pressure development

within the system, fault reactivation and foreshock sequence was able to develop.

Page 166: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

148

Chapter 6 Forecasts of Reservoir-induced Seismicity

The foreshocks of an earthquake can only be recognised after the main shock has

occurred (Gupta, 1992). Recognition of foreshocks soon after their occurrence is one of the

most important issues in earthquake prediction studies. Although many earthquake forecast

methods have been proposed in the past (e.g., Scholz, 2002; Mogi, 1985), many of these

methods suffer from a lack of reliable data in the run-up to failure, i.e., recording foreshock

sequences and how they can be identified as distinct from a ‘regular’ shock before the event

(Helmstetter, 2003). However, the classification of RIS into foreshocks, main shock, and

aftershocks is much easier due to their localisation within the reservoir region and their

correlation with pore pressure changes. As a result, the study of RIS provides an

exceptionally good opportunity to understand the mechanics of natural earthquakes (Bell and

Nur, 1978; Talwani and Acree, 1984/85; Roeloffs, 1988; Talwani, 1997). The ultimate goal

of both field and laboratory studies of RIS is to forecast the acceleration of seismicity and

earthquakes ahead of time (Umino, 2002; Lin, 2009). The identifiable foreshock–aftershock

sequences of RIS can be used for the verification of various forecasting models, thus

providing a step forward in the long-term goal of earthquake prediction (Gupta, 2002).

In other disciplines, in particular volcanology (Kilburn and Voight, 1998; Reyes-Dávila

et al., 2002; Kilburn 2003; Lavallée et al., 2008), foreshock sequences are also well defined,

which permits the used of the materials failure forecast method for the prediction of volcanic

eruptions (Voight, 1988; McGuire and Kilburn, 1997; Kilburn and Voight, 1998;

Reyes-Dávila and Cruz-Reyna, 2002; Kilburn, 2003; Lavallée et al., 2008). The repeated

pressurization of a volcanic edifice is cyclical in nature and is governed by the same

Page 167: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

149

mechanical and physical rock processes as those govern tectonic deformation. Prior to a

major volcanic eruption, the daily rate of earthquakes often indicates substantial increase

(Kilburn and Voight, 1998). This increase is accompanied by accelerating rates of ground

deformation (Jackson et al., 1998), and is clearly related to the propagation of a magma

conduit through the volcanic edifice. Similarly, in the case of RIS, prior to a major

earthquake, the daily rate of seismicity often indicates a substantial increase due to crack

propagation and/or coalescence. Therefore, the forecast method used in volcanology may be

applicable to the forecasts of RIS.

6.1 Material failure forecast method

In the case of volcanology, eruptions are commonly preceded by self-accelerating

processes such as earthquake frequency and rates of ground deformation (Scarpa and Tilling,

1996; McGuire and Kilburn, 1997). Such acceleration is described by an empirical relation

known as Voight-Fukuzono (VF) relation, proposed for modelling pre-eruption rates of

volcanic deformation (Voight, 1988, 1989; Main, 1999):

γ)(2

2

dt

dA

dt

dc

Ω=

Ω Equation 7.1

where t is time, Ac is a constant, γ is an exponent that measures the degree of non-linearity

(Voight, 1988 & 1989) and lies between 1 and 2 (Kilburn, 2003), and Ω is related to

precursory strain. This equation has been applied retrospectively to predict the time of

volcanic eruptions (Voight, 1988; Kilburn and Voight, 1998) and to predict earthquake failure

times, often based on the observation of accelerating strain (e.g., Bufe and Varnes, 1993;

Bowman et al., 1998).

Page 168: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

150

When approaching the final stages of ground deformation, crack growth becomes

uncontrolled propagation and is described by the condition γ = 2. This failure forecast

method uses the concept that the time-to-failure or the occurrence of peak seismicity depends

on the cumulative damages on the weak zones and, hence, the inverse of the seismic rate or

the inverse of the peak energy rate is used to forecast the time-to-failure. A plot of inverse

seismic rate or inverse of peak energy rate against time follows a negative linear trend, so

that the time at which the inverse rate is zero corresponds to the uncontrolled crack

propagation when dΩ/dt tends to infinity. This specific time can be obtained by a simple

linear extrapolation of the measured trend to the time axis.

6.2 Forecast of main slip

The material failure forecast model (Kilburn and Voight, 1998; Kilburn, 2003; Lavallée

et al., 2008) was adopted for the forecast of peak seismicity for the foreshock sequence data

obtained from the Fontainebleau sandstone F5 experiment. The data set was fit to Equation

7.1, with Ac equals 0.03 and γ equals 2 (Figure 6.1). By plotting the inverse of peak acoustic

emission rate of each pore pressure cycle with time, the extrapolation of the inverses crossing

the lowest points was drawn (Figure 6.2). A best-fit line crossing a minimum of at least four

lowest points on a moving time window was used based on Kilburn and Voight (1998). When

the extrapolated line intercepts the time axis, the expected time-to-failure or expected peak

seismic rate occurrence was defined. Using this method, a main slip was forecasted at cycle

10, while the actual peak seismicity occurred at cycle 11. A sensitivity analysis was also

conducted by using different number of lowest data points for the extrapolation (Table 6.1).

The analysis indicates that the use of the first four data points for the extrapolation would

Page 169: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

151

have generated a higher forecast error (25%).

Figure 6.1 Fitting Ac equals 0.03 and γ equals 2 into failure forecast model: γ)(2

2

dt

dA

dt

dc

Ω=

Ω .

Table 6.1 Sensitivity analysis for the ‘long-term’ forecast of the F5 experimental data, using different number of

points in moving window and different number of points for extrapolation.

Data points within

moving window

Points used for

extrapolation

Estimated

Error %

4 4 25%

7 4 7.6%

8 4 8.3%

8 5 9.5%

8 6 12.2%

Page 170: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

152

Figure 6.2 Application of failure forecast model to the experimental foreshock sequence obtained from the F5

experiment. The method is based on the extrapolation of the inverse of peak seismic rate. The dots indicate the

inverse of the peak seismic rate of each pore pressure cycle, while the solid line indicates the extrapolation from

the four lowest points. The expected time-to-failure is when the line hits the x-axis, which indicates that peak

seismicity occurred in the 10th pore pressure cycle, while the actual main slip and peak seismicity occurred in

the 11th

pore pressure cycle (Ying et al., 2009).

Using this same data set, a ‘short-term’ forecast based on the seismic rate at cycle 11,

prior to the main slip #2 (Figure 6.3) was also performed. The estimated time-to-failure was

just 2 seconds earlier than the actual main slip. This failure forecast model provides more

accurate forecasts for the ‘short-term’ data than the ‘long-term’ data.

Page 171: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

153

Figure 6.3 Application of failure forecast model to short-term forecast of main slip #2 in the F5 experiment.

Seismic rate recorded during cycle 11, prior to the main slip (Ying et al., 2009).

With the approximate size-frequency scaling relationship that has been established in

the literature for analysing millimetre-scale laboratory experiments as compared to field

observations in kilometre-scale (e.g. Burlini et al., 2007; Benson et al., 2008), and the

promising estimated results of the failure forecast model on experimental data, the next step

would be to apply this forecast method to field data. The field-to-experiment size ratio is

kilometre-to-millimetre, i.e. ~106. While the cyclic pore pressure frequency ratio between the

experiment and the field is also in the order of ~106, i.e. experimental frequency of 10

-2 Hz (1

to 2-minute periods) to seasonal oscillation in the order of ~10-8

Hz (annual period). Two

field cases would be chosen for the purpose, one for long-term forecast and the other for

short-term forecast, as described in sections 6.3 and 6.4, respectively.

Page 172: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

154

6.3 Application to the Koyna reservoir (Protracted seismicity forecast)

One classic example of protracted seismicity related to seasonal water level change is

the Koyna reservoir (India), which is one of the longest known sequences of RIS (Gupta,

2002). The impoundment of the Koyna reservoir started in 1961 and the seismicity of the

region has been monitored and observed since 1963. There was a lack of data during the

early stage of reservoir impoundment. After the devastating M6.3 earthquake occurred on 10

December 1967, a seismological network consisting of seven stations was established from

1967 to 1972.

The Koyna reservoir is situated near the west coast of India in the Deccan volcanic

province that covers one-third of the peninsular shield and is known as the Deccan traps

terrain. These basalts are of Palaeocene age (65 Ma), with numerous fractures and tensional

joints underlain by Archean granite and gneiss (Jain et al., 2004). The Deccan trap rocks have

a low permeability and the migration of water occurs through fractures and faults. As

suggested by Gupta (2005) that the Koyna reservoir RIS is an ideal site for earthquake

forecast studies; the failure forecast model was applied to the Koyna RIS data. The aim was

to forecast the peak seismicity that occurred in December 1967. The starting data of the

analysis is in January 1963, with a moving time window of 30 data points. At least 4 data

points were used for the extrapolation of the linear regression line of the inverse seismic rate

trend. The retrospective forecast based on the extrapolation line suggests that peak seismicity

would be expected in January 1968 (Figure 6.4). However, the actual peak seismicity

occurred in December 1967, which is one month earlier. In this case, the application of the

forecast method to long-term RIS data (in months) provides fairly accurate forecast of the

Page 173: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

155

occurrence of peak seismicity.

Figure 6.4 Long-term forecast of the Koyna RIS: Seasonal water level changes and seismicity of M > 2.0

occurred between 1963 and 1970 at the Koyna reservoir (Talwani, 1995) (using seismic data starting in January

1963, with a moving time window with 30 data points). The 4 red square points were used for the extrapolation

of the forecast model. The model suggests an expected peak seismicity occurred in January 1968, which is one

month later than the actual peak seismicity month (December 1967).

Page 174: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

156

6.4 Application to the Monticello reservoir (Initial peak seismicity forecast)

The second example of field application is related to the peak seismicity forecast due to

initial impoundment. The effect of fluid pressure development due to reservoir impoundment

is similar to the fluid pressure development in a magma chamber toward the final stage of

ground deformation, crack growth and propagation. Prior to a major volcanic eruption, the

daily rate of earthquakes often increases substantially. Similarly, the daily seismicity in RIS

related to initial reservoir impoundment or substantial water level changes is also

accompanied with accelerated rates of seismicity. Thus, the same failure forecast method that

has provided reasonable volcanic eruption forecast (Kilburn and Voight, 1998; Kilburn, 2003;

Lavallée et al., 2008) can be applied to estimate the time-to-peak seismicity in the RIS.

Monticello Reservoir is located in South Carolina, U.S.A. The region consists of a thick

stratified sequence of Cambrian meta-sedimentary and meta-volcanic rocks (Secor, et al.,

1982). The South Carolina seismic network was deployed before reservoir impoundment,

which enabled the detection and accurate location of the seismicity events due to the

reservoir impoundment. Induced seismicity has been observed in the Monticello Reservoir

region since the impoundment of the reservoir in December 1977. This seismicity was

attributed to the undrained elastic response due to impoundment of the reservoir (Talwani,

1997). The failure forecast method was applied to forecast the peak initial RIS. The forecast

indicates that the expected date of peak seismicity is 6th February 1978, while the actual peak

seismicity did occur in February 1978 (Figure 6.5) (Talwani, 1997). However, the exact date

of peak seismic rate is not available in the literature.

Page 175: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

157

Figure 6.5 Short-term forecast of the Monticello RIS for peak initial seismicity due to filling of the Monticello

reservoir, from December 1977 to January 1978 (seismic data started on 25 December 1977, with a moving

time window with 30 data points). Seismicity began about three weeks after the beginning of initial

impoundment (Chen and Talwani, 2001; Talwani, 1997). Based on the limited data, the forecast model suggests

peak seismicity is 6th February 1978 (Ying et al., 2009).

A sensitivity analysis was conducted using different number of data points for the

extrapolation of failure forecast (Table 6.2). The results indicate that although the estimated

time-to-failure was different using different number of points for extrapolation, the failure

forecasts fell within a narrow range between early to mid-February 1978.

Page 176: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

158

Table 6.2 Sensitivity analysis for the forecast of peak seismicity of the RIS at the Monticello reservoir using

different number of data points for the extrapolation.

No. of data points used for extrapolation Estimated time-to-failure/peak seismicity

4

5

6

7

8

9

10

11

12

13

6-Feb-78

10-Feb-78

11-Feb-78

9-Feb-78

11-Feb-78

15-Feb-78

14-Feb-78

16-Feb-78

19-Feb-78

18-Feb-78

6.5 Conclusions

It is observed that the fluid pressure development in initial impoundment in reservoir

has similar effects as that in volcanic eruption, in particular the acceleration of seismicity

toward the final failure or peak seismic rate when crack communication and linkage

processes become significant (Ying et al. 2009). In addition, the repeated pressurization of a

volcanic edifice is cyclical in nature, while the seasonal water level change in protracted RIS

sites are also cyclical in nature. The application of the failure forecast model that has

provided reasonable estimates in volcanic eruption forecast may therefore be applicable to

the forecasts in RIS. This forecast model provides reasonable retrospective estimates to the

laboratory data as well as the field data of Monticello reservoir. This is because the failure

Page 177: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

159

forecast method was developed for short-term forecast close to the final stages prior to

volcanic eruption when ground deformation and crack coalescence starts to be significant,

and when the acceleration of the seismic rate or seismic energy becomes rapid. Although the

short-term field case forecasts showed preliminary success, further investigation is required.

The application of this method to longer-term forecasts also requires further justification. It is

advisable that detailed seismic monitoring of RIS should be carried out starting from the

pre-impoundment stage, so that RIS can be distinguished from the local background

seismicity and reliable RIS data can be obtained for accurate failure forecasts. It is suggested

that the failure forecast model developed by Kilburn and Voight (1998) may provide some

general guideline for peak seismic rate estimates in short-term forecasts of RIS.

Page 178: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

160

Chapter 7 Conclusions and Recommendations

7.1 Conclusions

Fluids play an important role in the triggering of seismicity in various areas such as

initial reservoir impoundment, long-term lake level fluctuation, fluid injections in deep

boreholes, geothermal operations, and volcanic activity. This research studies the mechanical

influence of oscillating pore fluid on naturally-fractured sandstones. With the aid of

non-destructive acoustic emission techniques, the acoustic emissions due to pore pressure

oscillations were recorded and analysed. The experimental results provide insights on the

failure mechanism of RIS:

a) Pore pressure variation can induce initial seismicity as well as protracted seismicity. In

the case of initial seismicity, an increase in pore pressure amplitude that exceeds the previous

maximum experienced by the system can induce rapid responses. In the case of protracted

seismicity, existing faults can still be reactivated when the applied pore pressure amplitude is

lower than the previous maximum. This is because the average pore pressure within the

system may continue to increase during the subsequent cycles despite the lowered oscillation

pore pressure peaks at the upstream.

b) This research suggests a number of factors which are important in protracted RIS and,

can influence the growth of seismicity, namely the mechanical properties of material, the

frequency of oscillation and the rate of increase of pore pressure. When the rate of change of

pore pressure is rapid, the seismic rate increases rapidly; when the rate of change of pore

pressure is gradual, the acceleration in seismic rate is also gradual and the number of seismic

Page 179: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

161

events greatly reduces. When the frequency of pore pressure oscillation is too high, the

system may not respond to the peak amplitude of the applied pore pressure. However, when

the frequency of pore pressure oscillation is too low (i.e., the rate of change of pore pressure

is gradual) the acceleration in seismic rate in response to pore pressure oscillation is also

gradual, and fewer seismic events are expected. Furthermore, it is found that the effect of the

amplitude of the applied pore pressure is influencing the seismic response to a less extent.

These factors imply that in the field if the rate of increase in initial reservoir impoundment

can be controlled at a slower rate, then seismicity is less likely to occur. Similarly, in theory,

the control of protracted seismicity can be achieved by reducing the rate of increase of the

seasonal water level change by discharging.

c) The comparison between the RIS responses of the Darley Dale and Fontainebleau

sandstones suggest that the mechanical properties of the sample can influence the

reactivation of fault and induce protracted seismicity. The factors that affect the protracted

seismic response include:

i) Interconnected porosity and permeability of the fractured sample. Fractured samples

with low porosity and permeability allow the pore pressure to be at a transient state for a long

period of time, which is indicated by transient downstream pore pressure peaks and phase

shifts between the upstream and downstream pore pressures.

ii) The heterogeneity of the fractured sample, which also influences the development of the

foreshock sequence. Foreshocks are not found in homogeneous media (in the case of DDS7

experiment), but precede the main shock in slightly heterogeneous media (in the case of F5

experiment).

Page 180: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

162

d) No recognisable event migration trends were observed due to the heterogeneity of the

faulted system and the relatively short distance between the existing fractures within the

sample. However, in the field when the distances are measured in kilometre-scale,

observations of migration trends are more probable.

e) Monitoring of phase shifts and pore pressure variations at both upstream and

downstream may provide guidelines for effective mitigation measures for reservoir-induced

seismicity.

f) The material failure forecast model, which has been widely used in volcanic eruption

forecast, provides promising retrospective forecasts for both the experimental data, and the

short-term field data (seismic data in days). However, the forecast for long-term field data

(seismic data in months) may not be reliable, and would require further investigation. As the

forecast model was developed based on the exponential trend of seismicity close to failure, a

more accurate fit for short-term forecasting than for long-term forecasting is justifiable.

g) Oscillating pore pressure can induce progressive shearing that prolongs the period of

protracted seismicity. This progressive shearing can gradually smooth the asperities at the

fractured contact surface and reduce the friction coefficient, consequently reducing the shear

strength of the fault and causing protracted failures.

h) Oscillating pore pressure also slows down the pore pressure development in the system

and, hence, extends the period of seismicity, as compared to a step increase in pore pressure.

Page 181: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

163

The progressive wearing of asperities smoothes the fractured surfaces as the number of pore

pressure cycles increase. This is a ‘slow’ mechanism due to pore pressure oscillation. The

damage can accumulate and develop into a main slip when a sufficient number of pore

pressure cycles is allowed.

7.2 Recommendations

Thorough investigation of the sites, including the geology and mechanical properties of

the region prior to the construction of surface reservoirs is advisable. The selection of

aseismic zones as locations for reservoirs is crucial; however, if surface reservoirs are

inevitably located in seismogenic zones, proper surveillance of the region’s seismic activity,

as well as the upstream and downstream pore pressure changes, must be carried out in order

to observe any trends of seismic growth or pore pressure enhancement within the region that

may develop into a main slip. It is recommended that seismic monitoring stations should be

set up prior to any man-made pore pressure fluctuations to monitor the background seismic

level as well. This information is particularly important for the study of foreshock sequences,

which can be used for failure forecasts. Furthermore, the material failure forecast model can

be employed for short-term failure estimates and any episodic fault reactivation.

This research provides valuable insights into the effect of oscillating pore pressure on

induced seismicity. It also addresses the gradual change in shear strength due to oscillating

pore pressure. Future laboratory studies on fluid-induced seismicity may focus on the

relationship between the number of pore pressure cycles and the surface roughness in a

controlled environment. These experiments can be done on replicas of naturally-fractured

Page 182: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

164

rock. Furthermore, according to the field statistics by Talwani et al. (2007), there is a

seismogenic permeability range that plays an important role in RIS. In order to identify the

possibility of such a seismogenic permeability range in a laboratory-controlled environment,

a series of rock samples that cover a wide range of permeability can be tested to stipulate the

range. Further research on the seismogenic permeability range of up-scaled model can also

be done by numerical modelling.

Page 183: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

165

References

Aki, K. and Richards, P. G. 2002. Quantitative Seismology: Theory and Methods, 2nd

Edition.

W. H. Freeman and Co., New York. 726 p.

ASC (Applied Seismology Consultants Ltd.). 2007. InSite Software Manual.

Aki, K. and Richards, P. G. 2002. Quantitative seismology. 2nd

ed. University Science Books.

Allen, C. R. 1978. Evaluation of seismic hazard at the Auburn damsite, California. U.S. Bur.

Reclam. Rep., Denver, Colo. 10 p.

Ariyoshi, K., Hori, T., Ampuero, J. -P., Kaneda, Y., Matsuzawa, T., Hino, R. and Hasegawa,

A. 2009. Influence of interaction between small asperities on various types of slow

earthquakes in a 3-D simulation for a subduction plate boundary. Gondwana Res., 16, pp.

534-544.

Ayling, M. R., Meredith, P. G. and Murrell, S. A. F. 1995. Microcracking during triaxial

deformation of porous rocks monitored by changes in rock physical-properties 1

Elastic-wave propagation measurements on dry rocks, Tectonophysics. 245, pp. 205-221.

Baecher, B. G. and Keeney, R. L. 1982. Statistical examination of reservoir induced

seismicity. Bull. Seismol. Soc. Am. 72, pp. 553-569.

Båth, M. 1965. Lateral inhomogeneous in the upper mantle. Tectonophysics. 2, pp. 483-514.

Baud, P. and Meredith, P. G. 1997. Damage accumulation during triaxial creep of Darley

Dale sandstone from pore volumetry and acoustic emission. Int. J. Rock Mech. & Min Sci.

34, paper 024.

Beeler, N. M., Simpson, R. W., Hickman, S. H. and Lockner, D. A. 2000. Pore fluid pressure,

Page 184: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

166

apparent friction, and Coulomb failure. J. Geophys. Res. 105(11), pp. 25533-25542.

Bell, M. L. and Nur, A. 1978. Strength changes due to reservoir-induced pore pressure and

application to Lake Oroville. J. Geophys. Res. 83, pp. 4469-4483.

Benson P. M., Vinciguerra, S. Meredith, P. G. and Young, R. P. 2008. Laboratory simulation

of volcano seismicity. Science. 322, pp. 249-252.

Bergmann, D., Galanulis, K., Winter, D. 1997. Advanced 3-D fringe projection system. GOM

mbH, Braunschweig.

Bernabé, Y., Mok, U. and Evans, B. 2006. A note on the oscillating flow method for

measuring rock permeability. Int. J. Rock Mech. Min Sci. 43, pp. 311-316.

Bourbié, T. and Zinzner, B. 1985. Hydraulic and acoustic properties as a function of porosity

in Fontainebleau sandstone. J. Geophys. Res. 90 (B13), pp. 11524-11532.

Bowman, D. D., Ouillon, G., Sammis, C. G., Sornette, A. and Sornette, D. 1998. An

observational test of the critical earthquake concept. J. Geophys. Res. 103, pp. 24359-24372.

Brace, W. F. and Byerlee, J. D. 1966. Stick-slip as a mechanism for earthquakes. Science

153(3739) pp. 990-992.

Brace, W. F., Walsh J. B. and Frangos, W. T. 1968. Permeability of granite under higher

pressure. J. Geophys. Res., 73, pp. 2225-2236.

Bufe, C. G. and Varnes, D. J. 1993. Predictive modeling of the seismic cycle of the greater

San Fransisco Bay region. J. Geophys. Res. 98, pp.9871-9883.

Burlin, L., Vinciguerra, S., Di Toro, G., De Natale, G. and Burg, J. P. 2007. Seismicity

preceding volcanic eruptions: new experimental insight. Geology 35, pp. 183.

Page 185: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

167

Byerlee, J. D. 1978. Friction of rocks. PAGEOPH. 116, pp. 615-626.

Carslaw, H. S. and Jaeger, J. C. 1959. Conduction of heat in solids. 2nd

ed. Clarendon Press,

Oxford.

Chen, L. and Talwani, P. 1998. Reservoir-induced seismicity in China. PAGEOPH. 153,

pp.133-149.

Chen, L. 2001. On the mechanics of reservoir-induced seismicity. Ph.D. thesis. University of

South Carolina.

Chen, L. and Talwani, P. 2001. Mechanism of initial seismicity following impoundment of

the Monticello reservoir, South Carolina. Bull. Seis. Soc. Am., 91, pp. 1582-1594.

Collins, D. S., Pettitt, W. S. and Young, R. P. 2002. High-resolution mechanics of a

micro-earthquake sequence. PAGEOPH. 159(1-3) pp. 197-219.

Couples, G. D. and Lewis, H. 2007. Introduction: the relationship between damage and

localization. Geological Society, London, Special Publications, 289 pp. 1-6.

David, C. and Darot, M. Jeannette, D. 1993. Pore structures and transport properties of

sandstone. Transp. Porous Med. 11, pp. 161-177.

De Natale, G. and Zollo, A. 1989. Earthquake focal mechanisms from inversion of first P and

S wave motions. In R. Cassini, G. Nolet and Panza, G. F. (eds.). Digital Seismology and Fine

Modelling of the Lithossphere, Plenum Publishing Co. pp. 319-419.

Detournay, E. and Cheng, A. H.-D. 1993. Fundamentals of poroelasticity. In: Hudson, J. A.

(ed.) Comprehensive Rock Engineering. Pergamon, Oxford, pp. 113-171.

Page 186: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

168

Dieterich, J. H. 1979. Modeling of rock friction 1. Experimental results and constitutive

equations. J. Geophys. Res. 84(NB5) pp. 2161-2168.

do Nascimento, A. F. 2002. The role of pore pressure diffusion in a reservoir-induced

seismicity site in NE Brazil. PhD thesis. University of Edinburgh, UK.

do Nascimento, A. F., Cowie, P. A., Lunn, R. J. and Pearce, R. G. 2004. Spatio-temporal

evolution of induced seismicity at Açu reservoir, NE Brazil. Geophys. J. Int. 158, pp.

1041-1052.

do Nascimento, A. F., Cowie, P. A., Lunn, R. J. and Pearce, R. G. 2005a. Numerical modeling

of pore pressure diffusion in a reservoir-induced seismicity site in northeast Brazil, Geophys.

J. Int. 160, pp. 249-262.

do Nascimento, A. F., Lunn, R. J. and Cowie, P. A. 2005b. Modelling the heterogeneous

hydraulic properties of faults using constraints from reservoir-induced seismcity. J. Geophys.

Res. 110, B09201.

Dullien, F. A. L. 1979. Porous media: fluid transport and pore structure. Academic Press. 396

p.

Evans, M. D. 1966. Man made earthquakes in Denver. Geotimes. 10, pp. 11-17.

ESG (Engineering Seismology Group Canada Inc). 2002a. GigaRecorder Manual.

ESG (Engineering Seismology Group Canada Inc). 2002b. Hyperion Software User’s Guide

(Version 8.0).

Faulkner, D. R., Mitchell, T. M., Healy, D. and Heap, M. J. 2006. Slip on ‘weak’ faults by the

rotation of regional stress in the fracture damage zone. Nature. 444, pp. 922-925.

Page 187: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

169

Fischer, G. J. 1992. The determination of permeability and storage capacity: Pore pressure

oscillation method. In B. Evans & T.-F. Wong (eds.). Fault Mechanics and Transport

Properties of Rocks. Academic Press, San Diego. pp. 187-211.

Fitzenz, D. D. and Miller, S. A. 2004. New insights on stress rotations from a forward

regional model of the San Andreas fault system near its Big Bend in southern California, J.

Geophys. Res., 109, B08404, doi: 10.1029/2003JB002890.

Ge, S., Liu, M., Lu, N. Godt, J. W. and Luo, G. 2009. Did the Zipingpu reservoir trigger the

2008 Wenchuan earthquake? Geophys. Res. Lett. 36, L20315, doi:10.1029/2009GL040349.

Gentier, S. S. and Hopkins, D. L. 1997. Mapping fracture aperture as a function of normal

stress using a combination of casting, image analysis and modelling techniques. Int. J. Rock

Mech. Min. Sci. Geomech. Abstr. 34(3-4), pp. 359.

Gibowicz, S. J. and Kijko, A. 1994. An Introduction to Mining Seismology, Academic Press,

399 p.

Gough, D. I. and Gough, W. I. 1970. Load-induced earthquakes at Lake Kariba – ii. Geophys.

J. R. Astron. Soc., 21. pp. 79-101.

Grasso, J. R. and Sornette, D. 1998. Testing self-organized criticality by induced seismicity. J.

Geophys. Res. 103, pp. 29965-29987.

Guéguen, Y. and Palciauskas, V. 1994. Introduction to the Physics of Rocks. Princeton

University Press, 294 p.

Guha, S. K. 2000. Induced Earthquakes. Kluwer Academic Publishers, 314 p.

Gupta, H. K., Rastogi, B. K. and Narain, H. 1972. Common features of the reservoir

associated seismic activities. Bull. Seismol. Soc Am. 62, pp. 481-492.

Page 188: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

170

Gupta, H.K. 1983. Induced seismicity hazard mitigation through water level manipulation at

Koyna, India: A suggestion. Bull. Seismol. Soc Am. 73(2), pp. 679-682.

Gupta, H. K. 1985. The present status of reservoir induced seismicity investigations with

special emphasis on Koyna earthquakes. Tectonophysics. 118, pp. 257-279.

Gupta, H. K. 1992. Reservoir-Induced Earthquakes. Elsevier, Amsterdam, 364 p.

Gupta, H. K. 2002. A review of recent studies of triggered earthquakes by artificial water

reservoirs with special emphasis on earthquake in Koyna, India. Earth-Sci. Reviews. 58, pp.

279-310.

Gupta, H.K. 2005. Artifical water reservoir-triggered earthquakes with special emphasis at

Koyna. Curr. Sci. 88(10), pp. 1628-1631.

Helmstetter, A., Sornette, D and Grasso, J-R. 2003. Mainshocks are aftershocks of

conditional foreshocks: How do foreshock statistical properties emerge from aftershock law.

J. Geophys Res. 108, No. B1, 2046.

Howells, D. A. 1974. The time for a significant change of pore pressure. Eng. Geol. 8, pp.

135-138.

Hsieh, P. A. and Bredehoft, J. D. 1981. A reservoir analysis of the Denver earthquakes: a

case of induced seismicity. J. Geophys. 86, pp. 903-920.

Ian, G. M., Meredith, P. G. and Jones, C. 1989. A reinterpretation of the precursory seismic

b-value anomaly from fracture mechanics. Geophys. J. 96, pp. 131-138.

Iverson, R. M. and LaHusen, R. G. 1989. Dynamic pore pressure fluctuations in rapidly

shearing granular material. Science. 246(4931), pp. 796-799.

Page 189: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

171

Jackson, P., Shepherd, J., Tokumaru, M., Yokobe, A., Jackson, B., Hick, P. and Manoharan, P.

1998. Ground deformation monitoring at Soufriere Hills Volcano, Montserrat. Geophys. Res.

Lett. 25(18), pp. 3409-3412.

Jaeger, J. C., Cook, N. G. W. and Zimmerman, R. W. 2007. Fundamentals of Rock Mechanics.

4th

ed. Blackwell Publishing, 475 p.

Jain, R., Rastogi, B. K. and Sarma, C. S. P. 2004. Precursory changes in source parameters

for the Koyna-Warna (India) earthquakes. Geophys. J. Int. 158, pp. 915-921.

Jones, L. M. and Molnar, P. 1979. Some characteristics of foreshocks and their possible

relationship to earthquake prediction and premonitory slip on faults. J. Geophys. Res. 84, pp.

3596-3608.

Kagan, Y. Y. and Knopoff, L. 1978. Statistical short-term earthquake prediction. Science.

236(4808), pp. 1563-1567.

Kendall, K. and Tabor, D. 1971. An ultrasonic study of the area of contact between stationary

and sliding surface, Proc. Roy. Soc. Lond. Ser. A. 323, pp. 321-340.

Kerr, R. A. and Stone, R. 2009. A human trigger for the great quake of Sichuan? Science. 323,

pp. 322.

Kilburn, C. R. J. and Voight, B. 1998. Slow rock fracture as eruption precursor at Soufriere

Hills volcano, Montserrat. Geophys. Res. Lett. 25(19), pp. 3665-3668.

Kilburn, C. R. J. 2003. Multiscale fracturing as a key to forecasting volcanic eruptions, J.

Volc. Geotherm. Res. 125, pp. 271-289.

King, G. C. P., Stein, R. S. and Lin, J. 1994. Static stress changes and the triggering of

Page 190: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

172

earthquakes. Bull. Seism. Soc. Am. 84, pp. 935-953.

Kisslinger, C. 1976.A review of theories of mechanisms of induced seismicity. Eng. Geol.

10, pp. 85-98.

Kranz, R. L., Saltzman, J. S. and Blacic, J. D. 1990. Hydraulic diffusivity measurements on

laboratory rock samples using an oscillating pore pressure method. Int. J. Rock Mech. Min.

Sci. & Geomech. Abstr. 27(5), pp. 345-352.

Landau, L. D. and Lifshitz, E. M. 1959. Fluid Mechanics. Pergamon Press. pp. 49.

Lavallée, Y., Meredith, P.G., Dingwell, D.B., Hess, K.-U., Wassermann, J., Cordonnier, B.,

Gerik, A. and Kruhl, J.H. 2008. Seismogenic lavas and explosive eruption forecasting.

Nature. 453, pp. 507-510.

Lee, M. K. and Wolf, L.W. 1998 Analysis of fluid pressure propagation in heterogeneous

rocks: implications for hydrologically-induced earthquakes. Geophys. Res. Lett. 25(13), pp.

2329-2332.

Lin, C.H. 2009. Foreshock characteristics in Taiwan: Potential earthquake warning. J. Asian

Earth Sci. 34, pp. 655-662.

Linde, A. L. and Sacks, I. S. 2002. Slow earthquakes and great earthquakes along the Nankai

trough. Earth Plan. Sci. Lett., 203, pp. 265-275.

Lockner, D. 1993. The role of acoustic emission in the study of rock fracture. Int. J. Rock

Mech. Min. Geomech. Abstr. 30(7), pp. 883-899.

Lockner, D. A. and Beeler, N. M. 2002. Rock failure and earthquakes. In W. H. K. Lee, H.

Kanamori, P. C. Jennings and C. Kisslingler (eds.). Int. Handbook of Earthquake and Engg.

Seismology. Academic Press, Amsterdam. pp. 505-537.

Page 191: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

173

Main, I. G. 1999. Applicability of time-to-failure analysis to accelerated strain before

earthquake and volcanic eruptions. Geophys. J. Int. 139, F1-F6.

Majer, E., Baria, R., Stark, M., Smith, B., Oates, S., Bommer, J. and Asanuma, H. 2007.

Induced seismicity associated with Enhanced Geothermal Systems. Geothermics. 36, pp.

185-222.

McGarr, A. and Simpson, D. 1997. Keynote lecture: A broad look at induced and triggered

seismicity, “Rockbursts and seismicity in mines”. In: S.J. Gibowicz and S. Lasocki, (eds.),

Proc. of 4th Int. Symp. On Rockbursts and Seismicity in Mines, Poland. A.A. Balkema,

Rotterdam. pp. 385-396.

McGarr, A. 1999. On relating apparent stress to the stress causing earthquake fault slip. J.

Geophys. Res. 104, pp. 3003-3011.

McGuire, W. J. and Kilburn, C. R. J. 1997. Forecasting volcanic events: some contemporary

issues. Geol. Rundsch. 86, pp. 439-445.

Mees, F. 2003. Applications of X-ray computed tomography in the geosciences. Geological

Society of London.

Miller, S. A., Collettini, C., Chiaraluce, L., Cocco, M. Barchi, M. and Kaus, B. J. P. 2004.

Aftershocks driven by a high-pressure CO2 source at depth. Nature. 427, 724-747.

Miyazawa, M., Venkataraman, A., Snieder, R. and Payne, M. A. 2008. Analysis of

microearthquake data at Cold Lake and its applications to reservoir monitoring, Geophysics.

73(3), O15–O21.

Mogi, K. 1963. Some discussions on aftershocks, foreshocks and earthquake swarms – the

fracture of a semi-infinite body caused by an inner stress origin and its relation to the

Page 192: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

174

earthquake phenomena (third paper). Bull. Earthquake Res. Inst. 41, pp. 615-658.

Mogi, K. 1985. Earthquake Prediction. Academic Press. 355 p.

Nelder, J. A. and Mead, R. 1965, A Simplex-Method for Function Minimization, Computer J.

7(4), pp. 308-313.

Obara, K. 2002. Nonvolcanic deep tremor associated with subduction in southwest Japan.

Science, 296, pp. 1679-1681.

Obara, K. and Hirose, H. 2006. Non-volcanic deep low-frequency tremors accompanying

slow slips in the southwest Japan subduction zone. Tectonophysics, 417, pp. 33-51.

Papazachos, B. C. 1975. Foreshocks and earthquake prediction. Tectonophysics, 28, pp.

213-226.

Pettitt, W. S. 1998. Acoustic emission source studies of microcracking in rock, PhD, Keele

University, UK.

Press, W. H., Teukolsky, S. A., Vetterling, W. T. and Flannery, B. P. 2007. Numerical Recipes:

The Art of Scientific Computing, 3rd ed., Cambridge University Press, Cambridge. 1265 p.

Rabinowicz, E. 1995. Friction and wear of materials. 2nd

ed. John Wiley and Sons, New

York. 336 p.

Rajendran, K. 1992. Poroelastic models and the mechanism of reservoir-induced seismicity,

Ph.D. Thesis, University of South Carolina, 141 p.

Rajendran, K. and Talwani, P. 1992. The role of elastic, undrained and drained responses in

triggering earthquake at Monticello reservoir, South Carolina. Bull. Seism. Soc. Am. 82, pp.

1867-1888.

Page 193: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

175

Rao, B. R. and Singh, C. 2008. Temporal migration of earthquakes in Koyna-Warna (India)

region by pore-fluid diffusion. J. Seismol. 12, pp. 547-556.

Rastogi, B. K. 2003. Studies on Koyna and other reservoir induced earthquakes. In: Proc.

National Workshop on Current Practices and Future Trends in Earthquake Geotechnical

Engineering.

Reyes-Dávila, G. A. and la Cruz-Reyna, S. D. 2002. Experience in the short-term eruption

forecasting at Volcan de Colima, México, and public response to forecasts. J. Volcan. &

Geoth. Res. 117, pp. 121-127.

Rice, J. R. and Cleary, M. P. 1976. Some basic stress diffusion solutions for fluid-saturated

elastic porous media with compressible constituents, Rev. Geophys. 14, pp. 227-241.

Richardson, E. and Marone, C. 2008. What triggers tremor? Science. 319, pp. 166-167.

Roeloffs, E. A. 1988. Fault stability changes induced beneath a reservoir with cyclic

variations in water level. J. Geophys. Res. 93(B3). pp. 2107-2124.

Roeloffs, E. A, Wang, H. F., Cheung, L. S. and Haimson, B. C. 1979. Monotonic and cyclic

pore pressure loading of sawcut sandstone. 20th

U. S. Symposium on Rock Mechanics.

Rogers, G. and Dragert, H. 2007. Episodic tremor and slip on the Cascadia subduction zone:

the chatter of silent slip. Science. 300, pp. 1942-1943.

Rubenstein, J. L., Vidale, J. E., Gomberg, J., Bodin, P., Creager, K. C. and Malone, S. D. 2007.

Non-volcanic tremor driven by large transient shear stresses. Nature. 448, pp. 579-582.

Saar, M. O. and Manga, M. 2003. Seismicity induced by seasonal groundwater recharge at

Mt. Hood, Oregon. Earth & Planetary Sci. Lett. 214, pp. 605-618.

Page 194: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

176

Salvo, L., Cloetens, P., Maire, E., Zabler, S., Blandin, J.J., Buffière, J.Y., Ludwig, W., Boller,

E., Bellet, D., Josserond, C. 2003. X-ray micro-tomography an attractive characterisation

technique in material science. Nucl. Instr. and Meth. in Phys. Res. B. 200, pp. 273-286.

Savage, R. M. and Aronson, S. 2005. Comprehensive Textbook of Intraoperative

Transophageal Enchocardiography. Lippincott Williams & Wilkins, 650 p.

Scarpa, R. and Tilling, R. I. (eds.). 1996. Monitoring and Mitigation of Volcano Hazards,

Springer-Verlag, Berlin, 841 p.

Scholz, C. H. 1968. The frequency-magnitude relation of microfracturing in rock and its

relation to earthquakes. Bull. Seismol. Soc. Am. 58(1) pp. 399-417.

Scholz, C. H. 2002. The Mechanics of Earthquakes and Faulting. 2nd ed. Cambridge, 471 p.

Secor, D. T. Jr., Smith, W. A., Snoke, A. W., Peck, L. S., Pitcher, D. M., Prowell, D. C. and

Simpson, D. H. 1982. Geology of the area of induced seismic activity at Monticello

Reservoir, South Carolina. J. Geophys. Res. 87. B8.

Selim, M. M., Imoto, M. and Hurukawa, N. 2002. Statistical investigation of

reservoir-induced seismicity in Aswan area, Egypt. Earth Planets Space. 54, pp. 349-356.

Shapiro, S. A., Kummerow, J., Dinske, C., Asch, G., Rothert, E., Erziner, J. 2006. Fluid

induced seismicity guided by a continental fault: Injection experiment of 2004/2005 at the

German Deep Drilling Site (KTB). Geophy. Res. Lett. 33, L01309.

Shcherbakov, R. Turcotte, D. L. and Rundel, J. B. 2004. A generalized Omori’s law for

earthquake aftershock decay. Geophys. Res. Lett. 31, L11613.

Page 195: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

177

Shelly, D. R., Beroza, G. C. and Ide, S. 2007. Non-volcanic tremor and low-frequency

earthquake swarms. Nature, 446, pp. 305-307.

Shen, C, Chang, C., Chen, H., Li, T., Hueng, L., Wang, T., Yang, C. and Lo, H. 1974.

Earthquakes induced by reservoir impounding and their effect on the Hsinfengkiang Dam.

Sci. Sin. 17(2), pp. 239-272.

Shen, L. Y. and. Chang, B. Q. 1995. Application of stress–pore pressure coupling theory for

porous media to the Xinfengjiang reservoir earthquakes. PAGEOPH. 145(1), pp. 123-137.

Simpson, D. W. 1976. Seismicity changes associated with reservoir impounding. Eng. Geol.

10 pp. 371-385.

Simpson, D. W., Leith, W. S. and Scholz, C. H. 1988. Two types of reservoir-induced

seismicity. Bull. Seismol. Soc. Am. 78(6), pp. 2025–2040.

Simpson, D. W. and Narasimhan, T. N. 1992. Inhomogeneities in rock properties and their

influence on reservoir induced seismicity. In Knoll, P. (ed.), Induced Seismicity. Balkema,

Rotterdam. pp. 345-359.

Snow, D. T. 1972. Geodynamics of seismic reservoirs. Proc. Sym. Percolation through

fissured rock. Struggart, Ges. Erd and Grundbau. T2J 1-19.

Song, I. and Renner, J. 2006. Analysis of oscillatory fluid flow through rock samples.

Geophys. J. Int. 170, pp. 195-204.

Stuart-Alexander, D. E. and Mark, R. K. 1976. Impoundment-induced seismicity associated

with large reservoirs. U.S. Geol. Surv. Open File Rep., pp. 76-770.

Talwani, P. 1976. Earthquakes associated with Clark Hill Reservoir, South Carolina—a case

of induced seismicity. In: 1st Int. Symp. on Induced Seismicity. Eng. Geol. 1, pp. 239-253.

Page 196: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

178

Talwani, P. 1995. Speculation on the causes of continuing seismicity near Koyna reservoir,

India. PAGEOPH. 145, pp. 167-174.

Talwani, P. 1997. On the nature of reservoir-induced seismicity. PAGEOPH. 150. pp.

473-492.

Talwani, P. and Chen, L. 1998. Are seismogenic faults associated with a characteristic

permeability? Introducing ‘Seismogenic’ permeability, ks. EOS, Trans. AGU. 79, F583.

Talwani, P. 2000. Seismogenic properties of the crust inferred from recent studies of

reservoir-induced seismicity – Application to Koyna. Curr. Sci. 79(9), pp. 1327-1333.

Talwani, P., Chen, L. and Gahalaut, K. 2007. Seismogenic permeability, ks. J. Geophys. Res.

112, B07309.

Talwani, P. and Acree, S. 1984/85. Pore pressure diffusion and the mechanism of

reservoir-induced seismicity. PAGEOPH. 122. pp. 947-965.

Thompson, B. D., Young, R. P. and Lockner, D. A. 2009. Premonitory acoustic emissions and

stick-slip in natural and smooth-faulted Westerly granite, J. Geophys. Res. 114, B02205.

Townend J. and Zoback, M.D. 2000. How faulting keeps the crust strong, Geology. 28(5), pp.

399-402.

Trimmer, D. A. 1981. Design criteria for laboratory measurements of low permeability rocks.

Geophys. Res. Lett. 8(9), pp. 973-975.

Umino, N., Okada, T. and Hasegawa, A. 2002. Foreshock and aftershock sequence of the

1998 M 5.0 Sendai, Northeastern Japan, earthquake and its implications for earthquake

nucleation. Bull. Seismol. Soc. Am., 92(6), pp. 2465-2477.

Page 197: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

179

Utsu, T. 1961. A statistical study on the occurrence of aftershocks. Geophys. Magazine. 30,

pp. 521-605.

Voight, B. 1988. A method for prediction of volcanic eruption. Nature. 332, pp. 125-130.

Voight, B. 1989. A relation to describe rate-dependent material failure. Science. 243, pp.

200-203.

Von Seggern, D. 1980. A Random Stress Model for Seismicity Statistics and Earthquake

Prediction. Geophys. Res. Lett. 7(9), pp. 637-640.

Wang, W. L. 2008. Zheng Ming Magazine. June 15 (in Chinese).

Yamashita, T. 1999. Pore creation due to fault slip in a fluid-permeated fault zone and its

effect on seismicity: generation mechanism of earthquake swarm. PAGEOPH. 155, pp.

625-647.

Yeo, I. W., De Freitas, M. H. and Zimmerman, R. W. 1998. Effect of shear displacement on

the aperture and permeability of a rock fracture. Int. J. Rock Mech. Min. Sci. 35, pp.

1051-1070.

Ying, W. L., Benson, P. M. and Young, R. P. 2009. Laboratory simulation of Fluid-Driven

Seismic Sequences in Shallow Crustal Conditions. Geophys. Res. Lett. 36, L20301,

doi:10.1029/2009GL040230.

Young, R. P. and Bowes, R. 2002. Ultrasonic Giga-recorder for Seismic and Rock Fracture

Mechanics Research.

Young R. P., Collins D. S., Reyes-Montes J. M. and Baker, C. 2004. Quantification and

Interpretation of Seismicity, Int. J. of Rock Mech. 41(8), pp. 1317-1328.

Page 198: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

180

Zhao, G., Chen, H., Ma, S., Zhang D. 1995. Research on Earthquakes Induced by Water

Injection in China. PAGEOPH, 145(1), pp. 60-68.

Zoback, M. D., Zoback, M. L., Mount V. S., Suppe, J., Eaton, J. P., Healy, J. H.,

Oppenheimer, D., Resenberg, P., Jones, L., Raleigh, C. B., Wong, I., Scotti, O. and

Wentworth, C. 1987. New evidence on the state of stress of the San Andreas Fault System.

Science. 238 (4830), pp. 1105-1111.

Page 199: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

181

APPENDIX I – Calibration Charts

This section provides the relationship of transducer output (in volts) and the

corresponding measurement of pressure (in MPa), axial displacement (in mm), and volume

(in ml). By applying linear regression, the gradients and intercepts of each transducer

calibration line can be obtained. The more important parameters are illustrated in Figures A1

to A5, while a comprehensive list of parameters are tabulated in Table A1. In addition, the

calibration of source location is provided.

Page 200: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

182

Figure A1 Calibration of the two LVDTs: A and B

Page 201: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

183

Pa (Pressure)

y = 20.112x + 0.0114

0

5

10

15

20

25

0 0.2 0.4 0.6 0.8 1 1.2

Volts

Pore Pressure A

Pb (Pressure)

y = 20.077x + 0.0279

0

5

10

15

20

25

0 0.2 0.4 0.6 0.8 1 1.2

Volts

Pore Pressure B

Figure A2 Calibration of the pore pressure transducers: Pa and Pb.

Page 202: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

184

Pa (Volume)

y = 7.4947x + 37.532

0

10

20

30

40

50

60

70

-4 -2 0 2 4 6

Volts

Volume A (ml)

Pb (Volume)

y = 7.5014x + 37.529

0

10

20

30

40

50

60

70

-6 -4 -2 0 2 4

Volts

Volume B (ml)

Figure A3 Calibration of the permeameter cylinder volume: Pa and Pb.

Page 203: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

185

Cantilever A

y = 0.5844x + 0.0048

0

1

2

3

4

5

0 2 4 6 8

Volts

Deflection (mm)

Cantilever B

y = 2.3313x - 0.0008

0

1

2

3

4

5

6

7

0 0.5 1 1.5 2 2.5 3

Volts

Deflection (mm)

Figure A4 Calibration of the lateral strain measurement device.

Page 204: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

186

Figure A5 Calibration of the axial load transducers, Z1 and Z2.

Table A1 Summary of calibration parameters. The gradient, m, is the key parameter, representing the voltage

change per measurement change.

m, gradient c, intercept Description

1.5151 1.8967 LVDT A

0.5844 0.0048 Cantilever A

1.3009 2.9033 LVDT B

2.3313 -0.0008 Cantilever B

53.813 -26.38 Confining Pressure

27.581 1.0032 Pressure A

27.572 -1.1146 Pressure B

680.14 1.4159 Z1 Load

23.0675 -0.07 Z1 Displacement

680.14 1.4159 Z2 Load

10.1899 -0.025 Z2 Displacement

6.97 67.97 ISCO Vol A

6.97 67.97 ISCO Vol B

20.112 0.0114 Pa Pressure

7.4947 37.532 Pa Volume

20.077 0.0279 Pb Pressure

7.5014 37.529 Pb Volume

Page 205: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

187

Velocity models and source location calibration:

Different velocity models and anisotropy factors were being used at different stage of

the experiments to account for the development of transverse isotropic structure so that the

accuracy of automatic P-wave picking could be enhanced. Figure A6 and A7 show the

velocity models at the start of the experiment and after the failure of sample for typical

Darley Dale sandstone experiment and Fontainebleau sandstone experiment, respectively.

Page 206: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

188

a)

b)

Figure A6 Calibration of velocity models during different periods of the Darley Dale sandstone experiment. a)

The P-wave velocity change. b) Velocity model at the start of the experiment and after the sample failure.

Page 207: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

189

a)

b)

Figure A7 Calibration of velocity models during different periods of the Fontainebleau sandstone experiment. a)

The P-wave velocity change. b) Velocity model at the start of the experiment and after the sample failure.

Page 208: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

190

The absolute error in source location can be calculated by locating the synthetic AE

generated by each individual transmitter. The following example used a P-wave velocity

model of 5300 m/s and anisotropy of 31% for locating the synthetic AE during F5

experiment.

Figure A8 Locating a synthetic acoustic emission to calibrate source location.

Table A2 Absolute errors in source location.

Sensor ID Actual location of sensor Located sensor

Northing Easting Depth Northing Easting Depth

Absolute error

(mm)

Channel 1 23.47 9.72 -42.33 22.5 9.37 -40.1 2.44

Channel 2 9.72 -23.47 -42.33 8.78 -22.5 -40.1 2.56

Channel 3 25.4 0 -21.17 22.3 1.25 -20 3.08

Channel 4 0 25.4 -21.17 2.12 22.01 -21 2.57

Channel 5 -25.4 0 -21.17 -25.99 -0.57 -24.55 -2.69

Channel 6 0 -25.4 -21.17 3.02 -28 -23.8 -3.81

Channel 7 17.96 -17.96 21.17 20.2 -19.23 21.5 -2.15

Channel 8 17.96 17.96 21.17 17.05 16.78 19.48 2.21

Channel 9 -17.96 17.96 21.17 -16.8 16.8 19.81 2.13

Channel 10 -17.96 -17.96 21.17 -18.99 -19.52 -23.11 -2.65

Channel 11 -23.47 9.72 42.33 -21.01 8.78 40.11 3.24

Channel 12 9.72 23.47 42.33 9.67 23.01 40.23 2.02

The maximum absolute error in source location is 3.81 mm.

Page 209: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

191

APPENDIX II – Sensor location and sensor file

Figure A9 Locations of the lateral sensors

Page 210: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

192

Figure A10 Locations of the platen sensors

Page 211: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

193

Page 212: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

194

APPENDIX III – Details of the F2 experiment

Table A4 F2 experiment: Peak AE hits and the time delayed in response to the peak pore pressure magnitude.

Stage Cycle Range of pore

pressure (MPa)

Peak mean AE hits

per second

Delay in response

(s)

1 71 -

2 3.25 -

3 1.875 - 1

4

2.5 – 7.5

1.875 -

1

2

3 2

4

2.5 – 5

No significant AE has been recorded as the

maximum magnitude of pore pressure is less

than the previous peak value

1 1 -

2 0.875 -

3 0.75 - 3

4

2.5 – 7.5

0.5 -

1 23 11

2 2.4 54

3 1.8 70 4

4

2.5 – 10

1.3 49

1

2

3 5

4

2.5 – 7.5

No significant AE has been recorded as the

maximum magnitude of pore pressure is less

than the previous peak value

1 4.3 19

2 4.6 20

3 4 22 6

4

2.5 – 10

4.3 23

1

2

3 7

4

2.5 – 7.5

No significant AE has been recorded as the

maximum magnitude of pore pressure is less

than the previous peak value

1 3.75 28

2 4.6 22

3 3.56 21 8

4

2.5 – 10

4 20

1

2

3 9

4

2.5 – 7.5

No significant AE has been recorded as the

maximum magnitude of pore pressure is less

than the previous peak value

10 1 2.5 – 7.5

No significant AE has been recorded as the

maximum magnitude of pore pressure is less

than the previous peak value

Most of the AE were generated during the first cycle of a maximum pore pressure. Subsequent pore pressure

cycles also induce seismicity provided that the maximum experierenced pore pressure has been reached;

however, the number of AE reduced substantially. Delayed seismic response is also observed.

Page 213: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

195

APPENDIX IV – Response to Different Frequencies of Oscillation and

Rate of Increase in Pore Pressure

The rate of increase in seasonal water level change of artificial reservoir is about 40 m in 3

months time, which is equivalent to a rate of increase of 3 X 10-6 MPa/min. The experimental

results indicate that when the ratio of the rate of increase in laboratory experiment to that of

the field was in the order of 106, there were AE responses during the experiment.

Table A5 AE response corresponding to different frequencies and rate of increase in pore pressure.

Experiment

ID

Frequency

(Hz)

Rate of Increase

(MPA/min)

Change in Pore

pressure (MPa)

Ratio of rate of

increase (lab : field)

AE

Response

F5 0.2 72 3 2.E+07 N.G.

F5 0.1 36 3 1.E+07 N.G.

F5 0.1 48 4 2.E+07 N.G.

F5 0.1 60 5 2.E+07 N.G.

F5 0.1 72 6 2.E+07 N.G.

F5 0.1 84 7 3.E+07 N.G.

F5 0.1 96 8 3.E+07 N.G.

F5 0.1 108 9 3.E+07 N.G.

F5 0.05 54 9 2.E+07 N.G.

F5 0.0083 15 15 5.E+06 Yes

F2 0.0056 3.3 5 1.E+06 Yes

F2 0.0083 5 5 2.E+06 Yes

F2 0.0056 5 7.5 2.E+06 Yes

F2 0.0111 10 7.5 3.E+06 Yes

F3 0.0083 3.5 3.5 1.E+06 Yes

F3 0.0083 4.5 4.5 1.E+06 Yes

F3 0.0083 5.5 5.5 2.E+06 Yes

F3 0.0083 6.5 6.5 2.E+06 Yes

F3 0.0083 7.5 7.5 2.E+06 Yes

F4 0.0020 0.84 3.5 3.E+05 N.G.

F4 0.0014 0.94 5.5 3.E+05 N.G.

F4 0.0010 0.9 7.5 3.E+05 N.G.

DDS4 - 0.2 1 6.E+04 N.G.

DDS5 - 0.5 1 2.E+05 N.G.

DDS6 - 5 5 2.E+06 Yes

DDS7 0.0167 10 5 3.E+06 Yes

DDS7 0.0167 15 7.5 5.E+06 Yes

DDS7 0.0167 19 9.5 6.E+06 Yes

DDS7 0.0167 29 14.5 9.E+06 Yes

Note: Various oscillation frequencies have been used in Fontainebleau sandstone experiments

and it was found that when the frequency ranged between 0.0056 to 0.0111 Hz, seismicity

was induced; however, frequencies higher than 0.05 Hz or lower than 0.002 Hz did not

induce seismicity. More systematic investigation is required to determine the effective range

of frequencies of oscillation for inducing seismicity for different materials; however, it is

outside the scope of this research.

Page 214: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

196

APPENDIX V – Glossary

Abrasive wearing – occurs when a hard rough surface plough through a softer material.

Adhesive wearing – wearing that occurs with junctions shear off part of the adjoining

asperity, resulting in transfer of material from one surface to the other.

Acoustic Emission – a transient stress wave caused by the sudden release of the impulsive

strain energy of a material (Lockner, 1993), which travels as spherical wavefronts in the

material under stress.

Aftershock – an earthquake that occurs after a main shock.

b-value – the slope of the frequency-magnitude scaling relation i.e. log N(≥ M) = a – bM.

Continuous waveform – a stream of ultrasonic waveforms that is recorded in real-time by

the continuous data acquisition units, by each channel. It allows a complete AE catalogue to

be recorded for processing.

Far field – observation of seismic waveforms at distances from the source much larger than

the dominant wavelengths being generated at the source; i.e., several wavelengths away from

the source (ESG, 2002b).

Foreshock – an earthquake that occurs before a main shock.

Fracture – formation of new surfaces in material by breaking the material bonds, and

present in the form of cracks.

Hit count -- A ‘hit’ that is recorded when the waveform signal voltage exceeds a pre-defined

threshold within a pre-defined time window.

Page 215: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

197

Permeability - describes the ability of a medium to transmit fluid and is greatly influenced

by the porosity of the medium.

Phase shift – an observable repositioning between the applied upstream pore pressure and

the measured downstream pore pressure.

Porosity – the measure of the pore volume within the rock. It is defined as the fraction of

rock volume V that is not occupied by solid matter.

P-wave – a primary or compressional seismic wave generated by elastic energy release in

material.

S-wave – a secondary or shear seismic wave that has a comparatively low velocity rotational

vibrations (waves), which propagate through the rock mass.

Seismic event – instability within a rock mass often caused by rock fracturing or slip on

pre-existing features (ESG, 2002b).

Simplex method – an iterative location method that uses a geometric pattern (Simplex) to

minimise errors between the measured and calculated arrival times, and rapidly reduce the

search volume.

Source location – the point or location where crack initiates.

Sensor – a device (transducer) that converts vibrations of one quantity into those of another

(e.g. pressure to voltage or vice versa) (ESG, 2002a).

Triggered data – AE data that is recorded when the waveform signal voltage exceeds a

pre-defined threshold within a pre-defined time window.

Page 216: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

198

APPENDIX VI – Information of Paper Published in GRL

Ying, W. L., Benson, P. M. and Young, R. P. 2009. Laboratory Simulation of Fluid-Driven

Seismic Sequences in Shallow Crustal Conditions. Geophys. Res. Lett. 36, L20301,

doi:10.1029/2009GL040230.

Link: http://www.agu.org/pubs/crossref/2009/2009GL040230.shtml

Contributions to the paper:

Winnie (Wai-lai) Ying carried out the experiments, analysed the data and wrote the

paper.

Philip M. Benson assisted in preparation of experiments and provided valuable

discussions and comments on the paper.

R. Paul Young received NSERC discovery grant which was used for the development of

the Rock Fracture Dynamics Laboratory at the University of Toronto, in which the

experiments were carried out.

Page 217: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

199

Page 218: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

200

Page 219: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

201

Page 220: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

202

Page 221: LABORATORY SIMULATION OF RESERVOIR-INDUCED SEISMICITY · laboratory simulation of reservoir-induced seismicity by winnie ... laboratory simulation of reservoir-induced seismicity

203