You are on page 1of 412

ANALYSIS AND MODELING OF HIGH-RESOLUTION

MULTICOMPONENT SEISMIC REFLECTION DATA

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree Doctor of Philosophy in the

Graduate School of the Ohio State University

By

Erich D. Guy, B.S., M.S.

*****

The Ohio State University

2002

Dissertation Committee:

Dr. Jeffrey Daniels, Adviser

Dr. Matthew Saltzman

Dr. Franklin Schwartz

Dr. William Wolfe


UMI Number: 3110978

Copyright 2002 by
Guy, Erich D.

All rights reserved.

________________________________________________________

UMI Microform 3110978

Copyright 2004 ProQuest Information and Learning Company.


All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.
____________________________________________________________

ProQuest Information and Learning Company


300 North Zeeb Road
PO Box 1346
Ann Arbor, MI 48106-1346
© Copyright by

Erich D. Guy

2002
ABSTRACT

Seismic reflection has been applied during the past two decades towards the
solution of shallow earth problems, with single component P-wave methods traditionally
employed. The facts that (1) seismic body-wave types are sensitive to different physical
properties, (2) seismic sources radiate polarized waves, and (3) seismic receivers are
sensitive to the polarization of scattered body-waves and coherent noise, mean that it is
important to consider the potential usefulness of recording and analyzing different wave-
types and data components prior to conducting a high-resolution reflection survey.
Detailed studies regarding the applicability of multicomponent seismic reflection
techniques for the near-surface have been uncommon. In this dissertation, important
aspects of elastic-wave propagation relevant to high-resolution multicomponent
surveying have been analyzed experimentally and numerically, and methodologies have
been tested and developed that will improve near-surface imaging and characterization.
Factors affecting the ability and relative effectiveness of common-mode P- and S-
wave reflection surveys for mapping features in the near-surface are described and
illustrated through analyses of experimental field data and modeling. It is demonstrated
through comparisons of known subsurface conditions and processed stacked sections,
that combined P- and S-wave common-mode reflection information can allow a geologic
sequence to be imaged more effectively than by using solely P- or S-wave reflection
information. Mine subsidence-related project objectives in the field study area for this
research would not have been met using traditionally acquired P-P component data, but
study area project objectives were able to be met using S-wave reflection information.
Near-surface mode-converted seismic reflection imaging potential was tested
experimentally and evaluated through modeling. Modeling results demonstrate that

ii
potential advantages of near-surface mode-conversion imaging can be realized in theory.
Analyses of acquired multicomponent data however demonstrate that mode-conversion
imaging could not be accomplished in the field study area. Mode-converted reflections
having arrival times and moveouts similar to those that were predicted were not observed
in field data, due to the low amplitudes of mode-converted events and the presence of
random and coherent noise in field data. Analysis methods are presented that can be used
for assessing converted-wave imaging potential in future near-surface reflection studies.
Factors affecting the ability of SH-wave reflection measurements for allowing
near-surface interfaces and discontinuities to be effectively imaged are described. The
necessity of considering resolution issues, velocity model construction methods, and
imaging approaches during SH-reflection data analysis to ensure accurate subsurface
interpretation is demonstrated using numerical models and field data processing
illustrations. A SH-wave reflection data analysis workflow is presented that provides a
methodology for delineating areas of the subsurface where subsidence processes have
been active. The effectiveness of the workflow is demonstrated, and it is shown that SH-
wave reflection surveys can be used to assess mine-subsidence problems along roadways.
Equations that define an incidence angle (angle of intromission) for which the
SH-wave reflection coefficient is zero, and describe its conditions of occurrence are
presented. The SH-wave intromission angle and its potential application have not
previously been discussed in seismology literature.
One appendix of this dissertation presents a computer program developed for the
application of equations describing elastic-wave scattering from planar interfaces. The
programmed equations have a wide range of applications in seismology, and such code
has not previously existed in the public domain. In another appendix, it is shown through
analyses of acquired cross-hole radar data, that EM-wave velocity and amplitude
information can be used to infer near-surface media distribution and mine-related
subsidence activity between boreholes. Data acquisition considerations and data analysis
workflows developed for cross-hole radar measurements are also presented.

iii
Dedicated to my parents Dave and Gerry

iv
ACKNOWLEDGMENTS

I thank my advisor Dr. Jeffrey Daniels for guidance, suggestions, and support. I

also thank my committee members Drs. Matthew Saltzman, Franklin Schwartz, and

William Wolfe for reviews and suggestions, and Dr. Richard Nolen-Hoeksema

(University of Michigan) for research assistance and detailed reviews of chapters 6 and 7.

The seismic reflection and borehole radar data presented in this thesis were

processed using ProMAX software, and the elastic-wave modeling was conducted using

Sierra software; the Landmark Graphics Corporation generously provided both. The

Federal Highway Administration and the Ohio Department of Transportation provided

funding for the acquisition of the data that were used in this research. Special thanks go

to Thomas Lefchik and Rick Ruegsegger from these two organizations respectively.

I acknowledge Dr. Jeffrey Daniels, Dr. Richard Nolen-Hoeksema, John Clark

(Bay Geophysical), Robert Hinson (SeisTech, Inc.), Cam Walker (Walker Marine), Dr.

Don Steeples (University of Kansas), Jennifer Holt, and Zach Daniels for their

involvement/organization of seismic reflection data acquisition, and Zach Daniels for his

involvement in borehole radar data acquisition. I thank Chris Hall and Takeshi Hirano

from BBC&M Engineering, Inc., for providing me with I-70 project-related information.

I learned from many great professors at OSU, including: Drs. Michael Barton,

Tarunjit Butalia, Jeffrey Daniels, David Elliot, Hallan Noltimier, Doug Pride, Franklin
v
Schwartz, Rodney Tettenhorst, and William Wolfe. I also learned from being part a

research group at OSU composed of many fine people, including: Mohammad

Asgharzadeh, Dr. Christina Chan, Jim Conroy, Jennifer Holt, Dr. Selma Kadioglu, Dr.

Changryol Kim, Dr. Jens Munk, Pedro Paramo, and Dr. Stanley Radzevicius. Special

thanks go to: Mohammad Asgharzadeh and Pedro Paramo for seismic data processing

and interpretation discussions, Jim Conroy for I-70 study area and programming

discussions, Dr. Jens Munk for frequent geophysical discussions and advice, and Dr.

Stanley Radzevicius for numerous enjoyable field GPR research sessions and discussions

which improved the work in chapter 7 and appendix B.

I appreciate computer-related assistance that I received from Doug Findlay, Frank

Oh, Chris Toth, and Joe Walton (OSU Geology Dept. computing staff). Special thanks

go to Brent Curtiss (OSU Geology Dept. System Administrator) for his frequent technical

assistance and software installations and maintenance over the last several years.

I am grateful for summer research funding provided by Mark Vendl and the U.S.

Environmental Protection Agency, for conference travel funding provided on numerous

occasions by the Friends of Orton Hall (OSU Geology Dept.), and for a presidential

fellowship provided during my final year by The Ohio State University Graduate School.

I am thankful for valuable advice provided by Dave Guy on countless occasions.

DISCLAIMER

Any statements made in this dissertation concerning the size, nature and location

of objects in the study area subsurface, are an interpretation made by the author (who

assumes no liability). Such statements are subjective, and may be interpreted differently.
vi
VITA

1996…………………………………………… B.S. Geological Sciences, Biological


Science Minor, Mount Union College,
Alliance, Ohio
1998……………………………………………. M.S. Geological Sciences, GIS Focus,
Bowling Green State University,
Bowling Green, Ohio
1998 - Present………………………………….. The Ohio State University, Columbus,
Ohio

PUBLICATIONS

Journal Papers

Guy, E.D., Daniels, J.J., Nolen-Hoeksema, R.C., and Lefchik, T., Accepted, High-
Resolution SH-Wave Seismic Reflection Investigations Near a Coal Mine-Related
Roadway Collapse Feature, Journal of Applied Geophysics, Elsevier Science.

Guy, E.D., Radzevicius, S.J., and Conroy, J.P., Accepted, MATLAB Programs for
Application of Equations Describing Elastic and Electromagnetic Wave Scattering from
Planar Interfaces, Computers & Geosciences, International Association for Mathematical
Geology.

Guy, E.D., and Radzevicius, S.J., 2001, Recognition of Borehole Radar Cable-Related
Effects Using Variable Offset Sounding, Subsurface Sensing Technologies and
Applications: An International Journal, Kluwer Academic/Plenum Publishers, vol. 2, no.
2, pp. 127-139.

Guy, E.D., and Levine, N.S., 2001, GIS Modeling and Analysis of Ohio's CO2 Budget:
Mitigating Ohio's CO2 Emissions Through Reforestation, The Ohio Journal of Science,
The Ohio Academy of Science, vol. 101, no. 3/4, pp. 34-41.

vii
Radzevicius, S.J., Guy, E.D., and Daniels, J.J., 2000, Pitfalls in GPR Data Interpretation:
Differentiating Stratigraphy and Buried Objects from Periodic Antenna and Target
Effects, Geophysical Research Letters, American Geophysical Union, vol. 27, no. 20, pp.
3393-3396.

Guy, E.D., Daniels, J.J., Holt, J., Radzevicius, S.J., and Vendl, M.A., 2000,
Electromagnetic Induction and GPR Measurements for Creosote Contaminant
Investigation, Journal of Environmental and Engineering Geophysics, The Environmental
and Engineering Geophysical Society, vol. 5, no. 2, pp. 11-19.

Kim, C., Daniels, J.J., Guy, E.D., Radzevicius, S.J., and Holt, J., 2000, Residual
Hydrocarbons in a Water Saturated Medium: A Detection Strategy Using Ground
Penetrating Radar, Environmental Geosciences, American Association of Petroleum
Geologists, vol. 7, no. 4, pp. 169-176.

Guy, E.D., Daniels, J.J., Radzevicius, S.J., and Vendl, M.A., 1999, Demonstration of
Using Crossed Dipole GPR Antennae for Site Characterization, Geophysical Research
Letters, American Geophysical Union, vol. 26, no. 22, pp. 3421-3424.

Conference Papers

Guy, E.D., Daniels, J.J., and Daniels, Z., Accepted, Cross-Hole Radar Effectiveness for
Mine-Related Subsidence Investigations, the International 2003 NO-DIG Conference,
International and North American Societies for Trenchless Technology.

Daniels, J.J., Vendl, M.A., Holt, J.J., Guy, E.D., Accepted, Combining Multiple
Geophysical Data Sets into a Single 3D Image, 2003 Symposium on the Application of
Geophysics to Environmental and Engineering Problems, The Environmental and
Engineering Geophysical Society.

Guy, E.D., and Radzevicius, S.J., 2001, Non-Geologic Events in Single- and Cross-hole
Radar Data, Society of Exploration Geophysicists Annual Meeting Expanded Abstracts,
Society of Exploration Geophysicists, 4 p.

Kim, C., Daniels, J.J., Holt, J., and Guy, E.D., 2000, A Physical Model Experiment of the
GPR Response Over Gasoline, in: Proceedings of the Symposium on the Application of
Geophysics to Environmental and Engineering Problems, The Environmental and
Engineering Geophysical Society, pp. 303-310.

Frank, G., Guy, E.D., and Daniels, J.J., 2000, Use of Borehole Ground Penetrating Radar
in Determining the Risk Associated with Boulder Occurrence, in: Proceedings of the
North American NO-DIG Conference, North American Society for Trenchless
Technology, pp. 37-48.

viii
Radzevicius, S.J., Daniels, J.J., Guy, E.D., and Vendl, M.A., 2000, Significance of
Crossed-Dipole Antennas for High Noise Environments, in: Proceedings of the
Symposium on the Application of Geophysics to Environmental and Engineering
Problems, The Environmental and Engineering Geophysical Society, pp. 407-413.

Abstracts/Presentations

Guy, E.D., and Daniels, J.J., 2002, Comparison of P-Wave and S-Wave Reflection
Surveying Effectiveness for Detection of Mine-Related Subsidence Activity Beneath a
Heavily Traveled Roadway, EOS Transactions, American Geophysical Union, vol. 83,
no. 47.

Guy, E.D., Daniels, J.J. (speaker), Nolen-Hoeksema, R.C., and Lefchik, T., 2002, High-
Resolution SH-Wave Seismic Reflection Investigations Near a Coal Mine-Related
Roadway Collapse Feature, NATO Hydrogeophysics Advanced Study Institute.

Daniels, J.J., Guy, E.D., and Holt, J.J., 2002, Physical Model Tests of the Sensitivity of
GPR to LNAPL’s in the Vadose Zone, NATO Hydrogeophysics Advanced Study
Institute.

Paramo, P., Daniels, J.J., Asgharzadeh, M.F., and Guy, E.D., 2002, Processing and
Interpretation of Seismic Reflection Data Acquired in Northwest Ohio Over the East
Continental Rift, Rocky Mountain Section Meeting, American Association of Petroleum
Geologists.

Asgharzadeh, M.F., Daniels, J.J., Paramo, P., and Guy, E.D., 2002, Analysis of P-wave
Reflection Seismic Data Acquired in Northwest Ohio: Implications for Regional
Stratigraphy, Structure, and Hazardous Waste Injection Practices, American Association
of Petroleum Geologists and Society of Exploration Geophysicists joint exposition.

Guy, E.D., Daniels, J.J, Radzevicius, S.J., and Holt, J., 2000, Application of
Electromagnetic Techniques for Organic Contaminant Detection, EOS Transactions,
American Geophysical Union, vol. 81, no.19, p. S234.

Guy, E.D., and Levine, N.S., 2000, Atmospheric Carbon Dioxide Sequestration
Strategies Using Forestation Techniques: a Theoretical Study at the State Level, EOS
Transactions, American Geophysical Union, vol. 81, no.19, p. S92.

Guy, E.D., Daniels, J.J, Kim, C., Holt, J., Radzevicius, S.J., and Vendl, M.A., 2000,
LNAPL Effects on GPR Data, Abstracts with programs, Geological Society of America,
vol. 32, no.4, p. A15.

Guy, E.D., Daniels, J.J, and Radzevicius, S.J., 1999, Shallow Geophysics for
Investigation of a Creosote Contaminated Site in Central Ohio, AAPG Bulletin,
American Association of Petroleum Geologists, vol. 83, no.8, p. 1375.

ix
Radzevicius, S.J., Daniels, J.J., and Guy, E.D., 1999, GPR for Environmental Site
Investigations: the Importance of Polarization, AAPG Bulletin, American Association of
Petroleum Geologists, vol. 83, no.8, pp. 1375-1376.

Guy, E.D., and Levine, N.S., 1998, The Development and Testing of a Carbon Dioxide
Budget Model for the State of Ohio Using GIS, Abstracts with programs, Geological
Society of America, vol. 30, no.2, p. 20.

Guy, E.D. 1998. Development of a Carbon Dioxide Budget Model for the State of Ohio,
Sigma Xi Research Forum Abstracts, Sigma Xi, 2 p.

FIELDS OF STUDY

Major Field: Geological Sciences

Specialization: Engineering and Environmental Geophysics

x
TABLE OF CONTENTS
Page
Abstract ii
Dedication iv
Acknowledgments v
Vita vii
List of Tables xvii
List of Figures xix

Chapters:

1. Introduction 1
1.1 Background 1
1.2 Research Objectives 3
1.3 Significance of Research and Previous Work 4
1.3.1 Shallow Multicomponent Seismic Reflection Data Analysis 5
1.3.1.1 Common-Mode Seismic Reflection Analysis 6
1.3.1.2 Converted-Mode Seismic Reflection Analysis 7
1.3.2 Geophysical Detection of Subsurface Mine-Related
Subsidence Activity 9

1.4 Organization of Document 10

2. Seismic Wave Scattering Theory 15


2.1 Introduction 15
2.2 P- SH- and SV-Waves: Normal Incidence 16
2.3 SH-Waves: Non-Normal Incidence 21
2.4 P- and SV-Waves: Non-Normal Incidence 23

xi
2.4.1 Incident P-Waves 25
2.4.2 Incident SV-Waves 27

3. Experimental Seismic Reflection Data Acquisition 30


3.1 Introduction 30
3.2 Shear-Wave Reflection Survey - 1999 30
3.2.1 Data Acquisition 32
3.2.1.1 Energy Source 34
3.2.2 Data Pre-Processing 35
3.2.2.1 Vibroseis Correlation 35
3.2.2.2 Processing and Display Polarity 38
3.2.2.3 Geometry 39
3.2.2.4 Data Truncation and Trace Editing 40
3.3 Multicomponent Reflection Survey - 2001 41
3.3.1 Data Acquisition 43
3.3.1.1 Energy Source and Configurations 43
3.3.2 Data Pre-Processing 44
3.3.2.1 Multicomponent Data Polarity Considerations 44
3.3.2.2 Geometry 47
3.3.2.3 Data Truncation and Trace Editing 48

4. Factors Affecting Common-Mode Reflection Surveying Effectiveness 51


4.1 Overview 51
4.2 Multicomponent Reflection Data 53
4.3 XX Data Versus YY Data: Factors Affecting Imaging Potential 56
4.3.1 Noise, Optimal Reflection Windows, and Reflection
Coefficients 60

4.3.2 Stacked Signal Quality 66


4.4 YY Data Versus ZZ Data: Factors Affecting Imaging Potential 73

xii
4.4.1 Sensitivity of P- and S-waves to Changes in Overburden
Saturation 76

4.4.1.1 Reflection Coefficients: Unsaturated and Saturated


Overburden Interface 78

4.4.2 Noise and Optimal Reflection Windows 82


4.4.3 Resolution Potential of ZZ Component Data 84
4.4.3.1 P-Wave Versus S-Wave Resolution 88
4.4.4 Data Processing 90
4.4.4.1 Frequency-Wavenumber Filtering 90
4.4.5 P- and S-Wave Stacked Section Imaging 96
4.4.5.1 Insitu Physical Properties Measurement Potential
Using Reflections 99

5. Converted-Mode Reflection Imaging Potential 102


5.1 Overview 102
5.2 Field Multicomponent Reflection Data 104
5.2.1 Common-Mode Reflection Information 107
5.3 Converted-Mode Reflection Imaging Potential 116
5.3.1 Near-Surface Velocity and Density Structure 116
5.3.2 Energies of Mode-Converted Events 118
5.3.2.1 Mode-Conversions at the Unsaturated and
Saturated Overburden Interface 120

5.3.2.2 Mode-Conversions at the Saturated Overburden


and Bedrock Interface 121

5.3.3 Ray Tracing of Mode-Converted Reflections 125


5.3.3.1 Predicted Arrival Times and Moveouts 130
5.3.3.2 Comparison of Predicted and Observed Field Data
Events 132

5.3.4 Elastic-Wave Modeling 133


5.3.4.1 Forward Modeling Approach 134

xiii
5.3.4.2 Pure Source-Polarization Modeling Results 136
5.3.4.2.1 Common-Mode Reflection Synthetics 136
5.3.4.2.2 Converted-Mode Reflection Synthetics 141
5.3.4.3 Mixed Source-Polarization Modeling Results 142
5.3.4.3.1 SV-P and P-P Synthetics 143
5.3.4.3.2 P-SV and SV-SV Synthetics 145

6. SH-Wave Reflection Data Analysis Considerations 149


6.1 Overview 149
6.2 Shear-Wave Reflection Data 150
6.3 Shot Gather Analysis 151
6.3.1 Sources of Noise 152
6.3.2 Overburden and Bedrock Interface Reflection Coefficients 156
6.4 Target Resolution Potential 161
6.4.1 Reflecting Horizon Resolution Potential 161
6.4.2 Discontinuity Resolution Potential 169
6.5 Data Processing 174
6.5.1 Processing Strategy 174
6.5.2 Processing Flow 176
6.5.2.1 Velocity Analysis 179
6.5.3 CMP Stacking Versus Common Offset Imaging
Approaches 184

6.6 Results and Interpretations 185


6.6.1 Line Test-1 187
6.6.2 Line GUE-I70-1 190
6.6.3 Line EBPassYY 198
7. Development of SH-Wave Intromission Angle Theory 205
7.1 Overview 205
7.2 P-Wave Intromission Angle in Acoustic Media Overview 206
7.3 Conditions for SH-Wave Intromission Angle 213

xiv
7.4 Source Code Listing: p_acoustic.m 223

8. Conclusions 226
8.1 Importance of Considering P- and S-Wave Imaging Potential Prior
to Conducting High-Resolution Reflection Surveys 226

8.1.1 Applicability of P- and S-Wave Common-Mode Reflection


Information 227

8.1.2 Potential for Improving Near-Surface Characterization


Using Converted-Mode Reflection Imaging 231

8.2 Factors Affecting SH-Wave Reflection Method Applicability for


Shallow Earth Characterization 233

8.2.1 Shallow SH-Reflection Imaging in Complex and High


Noise Environments 233

8.3 Potential Application of SH-Wave Intromission Angle Equations 236


8.4 Importance of Developed PSHSV Computer Program 237
8.5 Applicability of Cross-Hole Radar for Subsidence Studies: Data
Acquisition and Analysis Considerations 238

8.5.1 Borehole Radar Data Acquisition Considerations 239


8.5.2 Cross-Hole Radar Data Analysis Considerations 240

Appendices:

A. Seismic Reflection Field Test Area 242


A.1 Location 242
A.2 Geologic Setting 242
A.3 Subsidence History 246
A.3.1 Remediation Efforts 250
A.4 Coal Mine-Related Subsidence 254
A.4.1 Underground Coal Mining Methods 254
A.4.2 Partial Extraction Mining Subsidence Mechanisms 255
A.4.3 Interstate 70 Mining and Subsidence Mechanisms 258

xv
B. Program for Application of Equations Describing Elastic Wave Scattering
from Planar Interfaces 261

B.1 Introduction 261


B.1.1 Solutions Obtained from Program Application 262
B.2 Program Description 263
B.2.1 Critical Angles and Phase 264
B.3 Program Functionality 266
B.4 Source Code Listing 280

C. Interpreted SH-Wave Depth Sections for the Entire Field Test Area 288
C.1 Introduction 288
C.2 Results and Interpretations 288

D. Cross-Hole Radar Effectiveness for Investigation of Seismically


Imaged Discontinuities 311

D.1 Overview 311


D.2 Introduction: Borehole Radar Methods 312
D.2.1 Antennas and Measurement Configurations 313
D.3 Cross-Hole Radar Data Acquisition: I-70 Study Area 315
D.3.1 Constant Offset Profile Measurements 319
D.3.2 Multiple Offset Gather Measurements 321
D.4 Cross-Hole Radar Data Processing and Imaging 323
D.4.1 Constant Offset Profile Analysis 323
D.4.2 Multiple Offset Gather Analysis 328
D.5 Cross-Hole Radar Results and Interpretations 331
D.5.1 Eastbound Travel Lane Road Stations 46940 to 46968 332
D.5.2 Eastbound Passing lane Road Stations 48304 to 48379 338
D.5.3 Eastbound Travel Lane Road Stations 48304 to 48395 346
D.5.4 Eastbound Travel Lane Road Stations 48530 to 48640 354

Bibliography 362

xvi
LIST OF TABLES
Table Page

3.1 Field acquisition and recording parameters for lines Test-1 and GUE-
I70-1 acquired during the 1999 seismic reflection survey 33

3.2 Field acquisition and recording parameters for Lines EBTravel, EBPass,
WBPass, and WBTravel, acquired during the 2001 seismic reflection
survey 42

4.1 Information from drill logs near line EBTravel 58

4.2 Data processing flow for Line EBTravel S-wave reflection data (XX
and YY and components), acquired east of the previous roadway
collapse region 69

4.3 Representative subsurface parameters and normal incidence P-wave and


S-wave cumulative reflection coefficients for the I-70 study area
subsurface lithologies and related contrasts in impedance 80

4.4 Velocity model used to generate the synthetic seismogram in Figure


4.12 85

4.5 Quarter-wavelengths calculated for P- and S-waves in the I-70 study


area overburden materials 89

4.6 Data processing flow for lines EBTravel S-wave (YY component) and
P-wave (ZZ component) reflection data 91

5.1 Data-derived I-70 subsurface model layer parameters used for square
root energy coefficient calculations and elastic-wave modeling 117

6.1 Velocity model used to generate synthetic data shown in Figures 6.7 and
6.8 163

6.2 Data processing flow for shear-wave reflection data 175

xvii
7.1 Incident and refracted media (media 1 and 2 respectively) physical
property scenarios used to model the synthetic seismograms and obtain
the solutions presented in Figures 7.1a - 7.1d 207

7.2 Incident and refracted media (media 1 and 2 respectively) physical


property scenarios used to obtain the solutions presented in Figures 7.3
and 7.4 216

B.1 Lithologic parameters used to obtain the solutions plotted as a function


of incidence angle in Figures B.1 - B.3 267

C.1 Locations and apparent dip directions of bedrock horizon faults (mine-
related) indicated on seismic YY component depth sections. Fault
locations (given in feet from the western county line) correspond to road
stations, and are plotted (with apparent dip directions indicated) relative
to the roadway 308

D.1 Cross-hole constant offset profile (COP) and multiple offset gather
(MOG) radar measurements in the I-70 study area (Figure D.1).
Borehole locations correspond to road stations; EBT = eastbound travel
lane, EBP = eastbound passing lane. Road stations are given in feet
from the western county line 317

D.2 Acquisition and recording parameters for cross-hole constant offset


profile (COP) and multiple offset gather (MOG) radar measurements in
the I-70 study area 322

D.3 Data processing and imaging flow for I-70 study area cross-hole
constant offset profile (COP) radar measurements 324

D.4 Data processing and imaging flow for I-70 study area cross-hole
multiple offset gather (MOG) radar measurements 324

xviii
LIST OF FIGURES
Figure Page

2.1 Particle motion associated with wave types (P, SH, and SV). P-wave
particle motion (left) is in the direction of propagation (indicated by
ray path) and within the plane of incidence (defined by source,
reflection point, and receiver). SH-wave particle motion (middle) is
perpendicular to the direction of propagation, and is within a plane that
is perpendicular to the plane of incidence. SV-wave particle motion
(right) is within the plane of incidence, but perpendicular to the
propagation direction 17

2.2 Nomenclature used for incident and scattered wave types and ray
angles with normal, and for calculated coefficient and ratio terms.
P1P2 for example, signifies a refracted P-wave in medium 2 resulting
from an incident P-wave in medium 1. Rays that are shown represent
the possible scattered waves from the planar interface (which is
perpendicular to the plane containing the rays) separating medium 1
and medium 2 (which are isotropic solids) 18

3.1 Photographs of 1999 (top) and 2001 (bottom) seismic reflection data
acquisition in the I-70 study area. See text for details regarding 1999
and 2001 data acquisition geometry and recording parameters 31

3.2 Map view of eastbound lanes of I-70 (road stations 48300 to 48500,
1999 seismic survey stations 100 to 300) showing the location of
reflection lines Test-1 and GUE-I70-1 32

3.3 Line Test-1 shot gathers: (a) correlated using AUX channel –2, and (b)
correlated using a synthetic sweep. The steeply-dipping periodic noise
(with dominant frequency of 65 Hz) evident at near offsets in shot
gathers correlated using AUX channel –2 was suppressed when shot
gathers were correlated using a synthetic sweep. Shot gathers are
unfiltered and have AGC scaling (100 ms window) applied for display
purposes. The x-axes scales of absolute offset from the source are in
feet 37

xix
3.4 CDP fold relative to road stations for 1999 survey lines: (a) Test-1, and
(b) GUE-I70-1 (for both lines the source and geophone intervals were
1 ft, and source locations were offset 2 ft north of geophones) 40

3.5 Map view of eastbound and westbound lanes of I-70 (road stations
48900 to 46694, 2001 seismic survey stations 1000 to 3206) showing
the location of reflection lines EBTravel, EBPass, WBPass, and
WBTravel 41

3.6 Multicomponent recording polarities for 3-C geophones with elements


arranged in a Cartesian system: (a) recording convention consistent
with SEG recommendations, and (b) 2001 multicomponent survey
recording convention. Raw data were recorded during the 2001 survey
with receiver polarities configured such that a tap on the top of a
geophone case yielded an initial positive amplitude, and taps on the
sides of a geophone case in the positive X and Y axis directions
yielded initial negative amplitudes. Source advancement direction is
in the positive X-axis direction, and the positive Y-axis direction has a
rotation angle of +90 degrees from the positive X-axis 46

3.7 Shot gather (P-wave source, Z component) recorded using horizontal,


inline (X component) geophone elements: (left) before, and (right)
after polarity reversal of trailing spread. Absolute values of offset
from the source on the x-axis are in ft 47

3.8 CDP fold relative to road stations for 2001 survey: (a) line EBTravel
inline and transverse shear source configurations (source interval was 1
ft for stations 48900-47558, and 2 feet for stations 47556-46694,
geophone interval was 2 ft), and (b) line EBTravel compressional
source configuration, lines EBPass and WBPass transverse shear
source configuration, and line WBTravel transverse and inline shear
source configurations (source and geophone intervals were 2 ft for all
stations) 50

4.1 Map view of the eastbound lanes of I-70 (road stations 48500 to
48900) showing the location of seismic reflection line EBTravel
relative to the locations of the underground mine workings. Due to the
small map scale and possible errors that exist regarding room and pillar
locations on the map, the spatial relationship between these features
and the I-70 road stations is regarded as only approximately accurate 54

xx
4.2 Nomenclature used to describe multicomponent reflection data, in
terms of the source and receiver orientations and preferential
polarizations. For sources, the X and Y symbols indicate sources
configured to preferentially generate shear particle motion inline and
transverse (crossline) to the seismic line respectively, while Z indicates
a source configured to preferentially generate compressional particle
motion. For receivers, the symbols X and Y indicate horizontal
geophone elements oriented inline and transverse to the line
respectively, while Z indicates a vertical geophone element. The
source and receiver pairs: XX, YY, and ZZ are referred to as the
common-mode components of the nine-component matrix (e.g. XX
means a source and receiver both oriented in the X direction) 55

4.3 Line EBTravel shot gathers: (a) XX component, and (b) YY


component. Gathers are shown: (top) unprocessed, (middle) with a
bandpass filter (50-80-160-200 Hz) and AGC (100 ms window)
applied, and (bottom) interpreted. S-wave reflections from the top-of-
bedrock are indicated, and apparent NMO velocities and approximate
depths are given. The x-axis scales of absolute offset (AOFFSET)
from the source locations (SOU_X; source locations are given in feet
from the western county line, see Figure 4.1) are in feet 57

4.4 Processed (bandpass filter and AGC gain applied) line EBTravel
multicomponent shot gathers: source components X (a), and Y (b).
The S-wave reflections interpreted in Figure 4.3 are superimposed on
the common-mode component (XX and YY) gathers. The x-axis
scales of absolute offset (AOFFSET) from the source locations
(SOU_X; source locations are given in feet from the western county
line, see Figure 4.1) are in feet 59

4.5 Plots showing normalized square root energy coefficients as a function


of incidence angle for a SV-wave (left) and a SH-wave (right) incident
on the overburden (medium 1) and bedrock (medium 2) interface 65

xxi
4.6 Line EBTravel CDP supergathers: XX component (left), and YY
component (right). Gathers are shown before and after NMO
correction using YY component-derived (Table 4.2) S-wave stacking
velocities of: (a) 737 ft/s, (b) 705 ft/s, and (c) 675 ft/s. Arrows next to
dynamic stack functions indicate the top-of-bedrock reflection event.
This reflection is slightly over-corrected on the 48820 (a) XX
component gather (although data still stack reasonably well at the
applied velocity). For supergathers centered at the other two locations
(b and c), this event is corrected similar on both the XX and YY
components. CDP locations (CDP_X) correspond to road stations
(given in feet from the western county line, see Figure 4.1) 67

4.7 Line EBTravel XX and YY component time sections after similar


processing (Table 4.2), and NMO corrections made using the same
velocity model: (a) 100 percent of stacking velocities, (b) 95 percent, and
(c) 105 percent. CDP locations (CDP_X) correspond to road stations
(given in feet from the western county line). Y-axis arrows indicate the
top of the bedrock horizon. The YY component sections have a higher
signal-to-noise ratio, and provide a better stack of the bedrock horizon.
No previous mining occurred below line EBTravel between CDP
locations 48880 and 48660 70
4.8 Line EBTravel shot gathers: (a) ZZ component, and (b) YY
component. Gathers are shown: (top) unprocessed, (middle) with a
bandpass filter (ZZ: 80-120-200-240 Hz, YY: 50-80-160-200 Hz) and
AGC (100 ms window) applied, and (bottom) interpreted. P-wave
reflections from the top-of-saturated-overburden (a) are indicated in
blue, and S-wave reflections from the top-of-bedrock (b) are indicated
in yellow, with apparent NMO velocities and approximate depths
given. The x-axis scales of absolute offset (AOFFSET) from the
source locations (SOU_X; source locations are given in feet from the
western county line, see Figure 4.1) are in feet 74

4.9 Line EBTravel multicomponent (source component Z, receiver


components X, Y, and Z) shot gathers: (a) unprocessed, and (b)
processed (bandpass filter: 80-120-200-240 Hz, and AGC gain with
100 ms window applied). P-wave reflections interpreted in Figure 4.8
are superimposed on the processed common-mode component (ZZ)
gathers. The x-axis scales of absolute offset (AOFFSET) from the
source locations (SOU_X; source locations are given in feet from the
western county line, see Figure 4.1) are in feet 75

xxii
4.10 Line EBTravel ZZ Component CMP gathers: (a) uninterpreted, and (b)
interpreted. Gathers are shown with a bandpass filter (80-120-200-240
Hz) and AGC (100 ms window) applied. P-wave reflections from the
top-of-saturated-overburden are indicated in (d), with apparent NMO
velocities around 1900 ft/sec, and reflector depths of about 32 feet. X-
axis scales of CMP location (CDP_X) and receiver location (REC_X)
are in feet, and correspond to road stations 77

4.11 Plots showing normalized square root energy coefficients as a function


of incidence angle for a P-wave (left), a SV-wave (middle), and a SH-
wave (right) incident on the interface between unsaturated overburden
(medium 1) and saturated overburden (medium 2) 79

4.12 Comparison of synthetic data with line EBTravel field data (ZZ
component). Plots of the velocity model and calculated event arrival
times are shown in (a) and (b) respectively. A synthetic seismogram
generated using the model in (a) with a 150 Hz source is shown
uninterpreted in (c), and interpreted in (d). A shot gather used as a
basis for forward modeling is shown uninterpreted in (e), and
interpreted in (f). A high reflection coefficient at the unsaturated and
saturated overburden (primary) interface, noise, and interference
prevent the interpretation of secondary events in field data (f) 86

4.13 Line EBTravel ZZ component (a) and YY component (b) shot gathers
(source located at road station 48638, east direction to left) and f-k
spectra: (left) gathers without f-k filter applied showing reflections
with zero-offset times of 32 ms (a) and 110 ms (b), and f-k amplitude
spectra showing defined mute polygons, (middle) with polygons
rejected to suppress noise (indicated in boxes 1 and 2), and (right) with
polygons accepted (showing noise rejected through filter application).
Bandpass filters and AGC gain were applied to the gathers before
generating these plots in order to demonstrate f-k filter non-reflection
energy suppression across reflection signal bandwidths. The x-axis
scales of absolute offset from the sources are in feet 93

4.14 Line EBTravel ZZ component (top) and YY component (bottom) time


sections without and with f-k filtering in processing flows. Notice the
suppression of noise and enhancement of reflection signal (within the
circled regions) achieved through application of f-k filters. CDP
location numbers (CDP_X) correspond to road stations (units are in
feet) 95

xxiii
4.15 Line EBTravel ZZ component (a) and YY component (b) time sections
with fold (TR_FOLD) plots. The P-wave reflection event at 28 to 33
ms (a) is the top-of-saturated-overburden on the P-wave depth section
(c). The S-wave reflection event at 105 to 115 ms (b) is the top-of-
bedrock on the S-wave depth section (d). The scales on the bottom x-
axes of (a) and (b) show P-wave and S-wave stacking velocities
respectively (velocity scales are to the right of the sections). CDP
location numbers (CDP_X) correspond to road stations (units are in
feet) 97

5.1 Map view of the eastbound lanes of I-70 (road stations 48600 to
48850) showing the locations of multicomponent seismic reflection
lines (road stations 48630 to 48830) presented in this chapter, relative
to the locations of the underground mine workings. For line EBTravel
9 component (9C) data were acquired, and for line EBPass 3C data
were acquired. Road stations are given in feet from the western county
line 105

5.2 Nomenclature used to describe multicomponent reflection data, in


terms of the source and receiver orientations and preferential
polarizations. For sources, the X and Y symbols indicate sources
configured to preferentially generate shear particle motion inline and
transverse (crossline) to the seismic line respectively, while Z indicates
a source configured to preferentially generate compressional particle
motion. For receivers, the symbols X and Y indicate horizontal
geophone elements oriented inline and transverse to the line
respectively, while Z indicates a vertical geophone element. The
source and receiver pairs: XX, YY, and ZZ are referred to as the
common-mode components of the nine-component matrix (e.g. XX
means a source and receiver both oriented in the X direction) 106

5.3 Line EBTravel 9C shot gathers: (a) unprocessed, and (b) with a
bandpass filter (X and Y sources: 50-80-160-200 Hz; Z source: 80-
120-200-240 Hz) and AGC applied. Common-mode S-wave
reflections (yellow) from the top-of-bedrock are indicated on XX and
YY shots; common-mode P-wave reflections (blue) from the top-of-
saturated-overburden are indicated on ZZ shots. Source locations
(SOU_X) are in feet from the western county line; AOFFSET units are
in feet 108

xxiv
5.4 Line EBTravel 9C depth stacks: (a) XX, XY, and XZ components, (b)
YX, YY, and YZ components, (c) ZX, ZY, and ZZ components. S-
wave source configuration sections (a) and (b) were processed for
common-mode S-wave reflection energy; the S-wave reflection
indicated on the YY component stack is from the top-of-bedrock. P-
wave source configuration sections (c) were processed for common-
mode P-wave reflection energy; the P-wave reflection indicated on the
ZZ component stack is from the top-of-saturated-overburden. CDP
locations (CDP_X) equal road stations (road stations are given in feet
from the western county line) 111

5.5 Line EBPass 3C shot gathers: (a) unprocessed, and (b) with a bandpass
filter (50-80-160-200 Hz) and AGC (100 ms window) applied.
Common-mode S-wave reflections from the top-of-bedrock are
indicated on YY gathers. The x-axis unit of absolute offset is feet.
Source locations (SOU_X) are given in feet from the western county
line 114

5.6 Line EBPass 3C depth stacks processed for common-mode S-wave


reflection energy: (top) YX, (middle) YY, and (bottom) YZ. The S-
wave reflection indicated on the YY component stack is from the top-
of-bedrock. CDP locations (CDP_X) equal road stations (road stations
are given in feet from the western county line) 115

5.7 Cross-section view of the I-70 depth model used for square root energy
calculations and elastic-wave modeling 118

5.8 Square root energy coefficients versus incidence angle using the media
parameters in Table 5.1: (a) downgoing waves at the top of saturated
overburden, (b) downgoing waves at the top of bedrock, and (c)
upgoing waves at the top of saturated overburden. Plots were
generated using the code in Appendix B. 119

5.9 Multiple mode-conversion point possibilities to a SV-wave from an


incident P-wave for an example geologic situation. Point C indicates
mode-conversion at reflection from an interface of interest (i.e. a
primary conversion), while points C’ and C” indicate conversion of
upward and downward traveling P-waves at transmission through an
overlying interface. Reflections represented by the different ray paths
shown will have different amplitudes, arrival times and moveouts
depending upon media compressional and shear impedances.
Horizontal and vertical scales are not implied 123

xxv
5.10 Ray trace plots through the I-70 model: (a) scattered P-waves from an
incident P-wave, (b) scattered mode-converted SV-waves from an
incident P-wave, (c) scattered SV-waves from an incident SV-wave,
and (d) scattered mode-converted P-waves from an incident SV-wave.
Mode-conversion rays that are shown occurred at reflection from
horizons of interest. Sources are located at 0 feet on the x-axis, and the
receiver interval along the x-axis is 2.0 feet 127

5.11 Diagrams demonstrating common-midpoint (CMP) versus common-


conversion-point (CCP) sorting/binning approaches for reflection data.
The CMP approach (a) is often used for sorting and stacking common-
mode reflections, for which the flat interface reflection point lies at the
source-to-receiver midpoint. For mode-converted reflections, the flat
interface reflection point does not lie at the source-receiver midpoint
(b), due to differences in media P- and S-wave velocities. A CCP
approach based on Vp/Vs properly sorts mode-converted reflection
data prior to event moveout correction and stacking (c). Horizontal
and vertical scales are not implied 129

5.12 Arrival times of common-mode and mode-converted reflections from


the I-70 model interfaces: (a) P-P reflections, (b) P-SV reflections, (c)
SV-SV reflections, and (d) SV-P reflections. Respective source types
are located at 0.0 feet (trace number 1) on the x-axis. Amplitude
variations with offset, and the ability to resolve/image each of the
reflection events that are plotted in this figure using field reflection
measurements are not implied 131

5.13 Synthetic seismograms generated using the model in Table 5.1 and
Figure 5.7: (a) with 250 Hz sources, (b) with source frequencies based
on I-70 field data component frequency filtering, and (c) with random
noise (amplitudes equal to 30 percent of average spike amplitudes)
added to each modeled component. Synthetics are displayed with
traces normalized and using the polarity convention outlined in
Chapter 3. The x-axis unit of absolute offset is feet 137

5.14 Synthetic seismograms generated using the model in Table 5.1 and
Figure 5.7. Source energy was equally divided between P- and SV-
waves (Z and X components respectively), and vertical receivers (Z
component) were used. The synthetic in (d) had random noise
(amplitudes equal to 30 percent of average spike amplitudes) added.
Synthetics are displayed with traces normalized and using the polarity
convention outlined in Chapter 3. The x-axis unit of absolute offset is
feet 144

xxvi
5.15 Synthetic seismograms generated using the model in Table 5.1 and
Figure 5.7. Source energy was equally divided between P- and SV-
waves (Z and X components respectively), and vertical receivers (Z
component) were used. The synthetics in (b) - (f) had random noise
(amplitudes equal to a percentage of average spike amplitudes) added.
Synthetics are displayed with traces normalized and using the polarity
convention outlined in Chapter 3. The x-axis unit of absolute offset is
feet 146

5.16 Synthetic seismograms generated using the model in Table 5.1 and
Figure 5.7. Source energy was equally divided between P- and SV-
waves (Z and X components respectively), and inline receivers (X
component) were used. The synthetic in (d) had random noise
(amplitudes equal to 30 percent of average spike amplitudes) added.
Synthetics are displayed with traces normalized and using the polarity
convention outlined in Chapter 3. The x-axis unit of absolute offset is
feet 147

6.1 Map view of the eastbound lanes of I-70 (road stations 48300 to
48500) showing the locations of seismic reflection lines (Test-1, GUE-
I70-1, and EBPassYY) relative to the locations of the underground
mine workings. The area of previous roadway failure is also shown,
where a surface collapse feature roughly 10 ft in diameter was centered
in the travel lane at road station 48345. Road stations are given in feet
from the western county line. Due to the small map scale and possible
errors that exist with regards to room and pillar locations, the spatial
relationship between these features and the road stations is regarded as
only approximately accurate 151

6.2 Examples of shot gathers (uninterpreted and interpreted) with zero-


phase Ormsby bandpass filters (< 12 dB/octave slopes) and AGC gain
(100 ms window) applied: (a) line Test-1 gather (source located at road
station 48382.5, eastbound direction is to the left) with 80-180 Hz filter
and gain, (b) line GUE-I70-1 gather (source located at road station
48363.5, eastbound direction is to the right) with 100-180 Hz filter and
gain, and (c) line EBPassYY gather (source located at road station
48464, eastbound direction is to the right) with 80-160 Hz filter and
gain. Road stations are given in feet from the western county line.
Hyperbolic reflection events interpreted on shot gathers correlate as the
top of bedrock. The x-axes scales of absolute offset from the source
are in feet 153

xxvii
6.3 Line GUE-I70-1 shot gather (source located at road station 48361.5,
eastbound direction is to the right) with various zero-phase Ormsby
bandpass filters (< 12 dB/octave slopes) and AGC gain (100 ms
window) applied: (a) unprocessed, (b) unfiltered with gain, (c) 40-100
Hz filter with gain, (d) 80-140 Hz filter with gain, (e) 120-180 Hz filter
with gain, and (f) 160-220 Hz filter with gain. High amplitude
periodic noise evident in shot gathers (predominantly in the range of
50-80 Hz) is related to source decoupling. The x-axis scale of absolute
offset from the source is in feet 155

6.4 Line EBPassYY shot gather (source located at road station 48464,
eastbound direction is to the right) and amplitude spectrum: (a)
unprocessed, and (b) with zero-phase Ormsby bandpass filter (80-160
Hz, < 12 dB/octave slopes) applied. Noise from roadway traffic
(indicated in the circled region) is predominantly low frequency (e.g.
5-25 Hz), but is also evident in the optimum frequency range for
imaging reflection signal. The x-axis scale of absolute offset from the
source is in feet 157

6.5 Line Test-1 shot gather (source located at road station 48397.5,
eastbound direction to the left): (a) processed and uninterpreted, (b)
amplitude spectrum indicating a dominant frequency of about 80 Hz,
and (c) interpreted. An apparent NMO velocity of 670 ft/sec indicates
the hyperbolic reflection is from the top of bedrock at a depth of 39 ft.
An apparent refraction velocity (unreversed) of 2500 ft/sec is also
interpreted. The x-axis scale on the shot gathers of absolute offset
from the source is in feet 158

6.6 Plot showing normalized square root energy coefficients as a function


of incidence angle for a SH-wave incident on the overburden (medium
1) and bedrock (medium 2) interface. Calculations were made using
the velocities determined from analysis of the shot gather in Figure
6.5c, with the computer program presented in Appendix B. Note the
intromission angle (șI) at about 15 degrees (see Appendix C), and the
critical angle (șc) at about 15.5 degrees 160

xxviii
6.7 Uninterpreted synthetic seismograms generated using the model in
Table 6.1 with a center source frequency of 300 Hz (a), and 100 Hz
(c). The three model interfaces are resolved at near and far offsets
with a source frequency of 300 Hz (b), while only the primary event is
easily interpreted with a source frequency of 100 Hz (d). Also shown
is the uninterpreted (e) and interpreted (f) shot gather (from Figure 4.5)
used as a basis for forward modeling. A high reflection coefficient at
the overburden-bedrock interface, near offset noise, a lower signal to
noise ratio, interference, and wavelet ringiness result in only the
overburden and bedrock reflection being interpretable from the field
data. The x-axis scale of absolute offset from the source is in feet 164

6.8 Uninterpreted (left), and with the top of bedrock reflection and
refraction interpreted (right) synthetic seismograms (center source
frequency = 100 Hz) generated using different S-wave velocities for
the coal seam specified than those in Table 6.1. Interference of
reflection energy does not allow lower amplitude events from the coal
top or bottom to be easily interpreted at any of these modeled coal S-
wave velocities 168

6.9 A synthetic section generated using the overburden and bedrock


velocity and depth parameters determined from the field data in Figure
4.7f as a basis. Five faults (fault locations are indicated by arrows on
the x-axis) were modeled, with the amount of vertical offset for each
fault specified as a fraction of the dominant wavelength (Ȝ). As seen
by the modeled results, vertical offset of the bedrock interface must be
at least a quarter of the dominant wavelength to be easily inferred
without relying on diffraction events (which may not be observed in
field data with a lower signal to noise ratio). The overburden and
bedrock boundary is 39 feet deep at CDP 100, and due to continued
downward offset this interface is at a depth of 55.3 ft at CDP 600
(CDP spacing is 0.5 ft) 170

xxix
6.10 A synthetic stacked section generated using the overburden and
bedrock velocity and depth parameters determined from the field data
in Figure 4.7f as a basis. Four bedrock graben structures resulting
from mine-related subsidence activity (graben locations are indicated
by arrows on the x-axis) were modeled, with the width of each
structure specified as a fraction of the Fresnel zone diameter (d). As
seen by the modeled results, reflections appear to be continuous across
a bedrock subsidence feature when the spatial extent of the feature is
much smaller than the size of the Fresnel zone. The top of bedrock
exists at a depth of 39 feet at all CDP locations, except where this
boundary has subsided to a depth of 46 ft in the grabens (CDP spacing
is 0.5 ft) 173

6.11 Synthetic seismograms (100 Hz center source frequency) that


demonstrate the effectiveness of a 30 percent post-NMO stretch mute
in eliminating most refracted energy. The models shown both consist
of a 40 ft thick overburden above bedrock (bedrock velocity = 2500
ft/sec). In example 1 (a) the overburden velocity is 600 ft/sec, and in
example 2 (b) the overburden velocity is 800 ft/sec. The x-axis scale
on the gathers of absolute offset from the source is in feet 178

6.12 Constant velocity stacks generated for line EBPassYY using a velocity
increment of 25 feet per second (stacking velocities from 700-800 feet
per second are shown). Note the change in event coherency within the
circled region at different stacking velocities. CDP location numbers
(CDP_X) correspond to road stations 180

6.13 Velocity spectrum and pick for line EBPassYY CDP supergather: (a)
before NMO correction, (b) after NMO correction showing significant
stretch at large offsets, and (c) after stretch mute application. The
supergather is centered at road station 48415 181

6.14 Velocity spectra and picks for line EBPassYY CDP supergathers
(before and after NMO correction and stretch mute application)
demonstrating lateral variation in optimum stacking velocity along this
line. The supergathers are centered at: (a) road station 48470, (b) road
station 48410, and (c) road station 48320 183

6.15 Line Test-1 data stacked using different ranges of source to receiver
offsets (ranges specified on stacks are in feet). Stacking all offsets
allowed the overburden and bedrock interface (110 to 120 ms) to be
most effectively imaged. Fold (TR_FOLD) plots are shown on stacks,
and CDP location numbers (CDP_X) correspond to road stations 186

xxx
6.16 Line Test-1 stacked section with fold (TR_FOLD) plot (a). The
continuous reflection event at 110 to 120 ms (a), is the top of bedrock
(b). The color bar on the bottom x-axis of (a) shows bedrock horizon
stacking velocities. Uninterpreted (c) and interpreted (d) depth
sections are also shown. CDP location numbers (CDP_X) correspond
to road stations (road stations are given in feet from the western county
line). To the east of CDP 48345 the overburden-bedrock horizon dips
down to the east and stacking velocities decrease, suggesting that the
removal of coal in this region (see text for discussion) may have
influenced the bedrock topography and resulted in a decreased
overburden stiffness 188

6.17 Line GUE-I70-1 stacked section with fold (TR_FOLD) plot (a). The
reflection at 110 to 120ms (a), is the top of bedrock (b). The color bar
on the bottom x-axis of (a) shows bedrock stacking velocities.
Uninterpreted (c) and interpreted (d) depth sections are also shown.
CDP location numbers (CDP_X) correspond to road stations (road
stations are given in feet from the western county line). A bedrock
discontinuity is interpreted at CDP_X 48391, and an area of disrupted
bedrock is interpreted between CDP_X 48380 and 48408 (indicated by
x-axis arrows and dashed lines). The interpreted area of disruption is
based upon on wavelet character and analysis of the shot gathers (x-
axis flags indicate locations) in Figure 6.19 191

6.18 Geologic cross-section constructed from 1999 drill log data acquired
along the southern edge of the eastbound travel lane of I-70 (Modified
from BBC&M Inc., 1999) 194

6.19 Line GUE-I70-1 uninterpreted (a) and interpreted (b) shot gathers with
source locations corresponding to the flag locations on the x-axis of
Figures 6.17b and 6.17d. Source locations (SOU_X) and CDP
locations (CDP_X) correspond to road stations. The reflection at 110
to 120 ms (b) is the top of bedrock. The 48375.5 shot gather indicates
an apparent updip direction to the east at this location. The 48389.5
shot gather supports the interpretation in Figures 6.17b and 6.17d of
discontinuity and offset in the bedrock horizon at approximately CDP
location 48391. Based on shot gather reflection character, an area of
disrupted bedrock is interpreted between CDP’s 48380 and 48408
(indicated on the x-axis of interpreted gathers by arrows) 196

xxxi
6.20 Line EBPassYY stacked section with fold (TR_FOLD) plot (a). The
reflection at 105 to 115 ms (a), is the top of bedrock (b). The color bar
on the bottom x-axis of the uninterpreted section (a) shows bedrock
horizon stacking velocities. Uninterpreted (c) and interpreted (d) depth
sections are also shown. CDP location numbers (CDP_X) correspond
to road stations (road stations are given in feet from the western county
line). The bedrock horizon is continuous across the section, except
between CDP_X 48329 and 48354 (indicated by x-axis arrows), where
discontinuity is interpreted 199

6.21 Geologic cross-section interpreted from 1999 drill log data acquired
along the north edge of the eastbound passing lane of I-70 in the
previous roadway collapse region (Modified from BBC&M Inc., 1999) 203

7.1 Synthetic seismograms (modeled with a center source frequency of


100 Hz) and plots showing plane-wave reflection coefficients
(magnitude) and phase changes (versus incidence angle) resulting from
an incident P-wave in acoustic media, for the scenarios in Table 7.1:
(a) scenario 1, (b) scenario 2, (c) scenario 3, and (d) scenario 4. The
density (units of g/cm3) and velocity (units of m/s) values used for
modeling are shown on each of the outputs in this figure. Absolute
offset values from the source on the x-axis of synthetic data are in feet.
Values of critical distance (xc), critical angle (șc), intromission distance
(xI), and intromission angle (șI) are indicated. Each model consisted of
two layers, with the interface between layers at 60 ft depth. See text
for discussion. Plots were generated using the code listed subsequently 208

7.2 FDTD modeled snapshots of acoustic P-wave scattering from a planar


interface (located at 60 feet depth), for the physical properties scenario
number 4 in Table 7.1: (a) 35 ms, (b) 45 ms, (c) 55 ms, and (d) 65 ms.
The reflection amplitude null and phase change in (d) are associated
with intromission angle occurrence under these modeled media
conditions 214

7.3 Amplitude, square root energy, and energy coefficients, and phase
changes from an incident SH-wave, for scenarios in Table 7.2: (a)
scenario 1, (b) scenario 2, (c) scenario 3, and (d) scenario 4. Density
(ȡ) units are g/cm3, and shear velocity (Vs) units are m/s. See text for
a discussion of plots. Plots generated using the PSHSV code in
Appendix B 217

xxxii
7.4 Synthetic SH-wave reflection seismograms (modeled with a center
source frequency of 100 Hz), for the scenarios in Table 7.2: (a)
scenario 1, and (b) scenario 2. The shear velocity, impedance, and
density relationships used are shown on the figures. Absolute offset
values from the source on the x-axes are in feet. The intromission
distance (xI) in (b) is indicated. Each model consisted of two layers,
with the interface between layers at a depth of 29.5 ft 221

A.1 Location of study area along I-70 in Guernsey County, Ohio, where the
roadway crosses approximately 2200 feet of the underground Murray
Hill No. 2 mine. The study area is east of Cambridge, Ohio, roughly 4
miles east of Route 77 243

A.2 Photograph and map view of the I-70 study area (stations 46700 to
48900) in Guernsey County, Ohio 244

A.3 Location of the I-70 study area relative to the Murray Hill No. 2 coal
mine, superimposed on a USGS topographic map (Modified from
BBC&M Inc., 1998) 247

A.4 Location of the I-70 study area relative to the Murray Hill No. 2 coal
mine room and pillar workings (Modified from ODNR, 1981) 247

A.5 Study area map: locations of observed subsidence and roadway


depression relative to mapped mine workings. Road stations are in
feet from the western county line 249

A.6 Photographs of the March 1995 surface collapse pit centered in the
eastbound travel lane of I-70 at road station 48345. The top photograph
is from Gannett Fleming Corddry & Carpenter, and the bottom
photograph is from the Ohio Division of the Federal Highway
Administration 251

A.7 Map of the I-70 study area showing the approximate grouting limits
and grout takes (in cubic yards per borehole). Road stations are in feet
from the western county line 253

xxxiii
A.8 Typical subsidence mechanisms into room and pillar coal mine
workings (scale is not implied): a) sag subsidence (after Whittaker and
Reddish, 1989), and b) pit subsidence (after Waltham, 1989). Sag
subsidence results in the settling of a relatively broad surface area, and
is caused by the failure of coal pillars (when pillars are either crushed
or punched into the underlying material due to overburden weight). Pit
subsidence usually results in the formation of relatively small pit
features at the surface, and is caused by the upward migration of
collapse features that develop due to mine roof failure between coal
pillars. See text for a detailed discussion of subsidence mechanisms 257

B.1 Plots (generated with the PSHSV program) showing displacement


amplitude, square root energy, and energy coefficients, and phase
changes as a function of incidence angle for a: (a) P-wave, (b) SH-
wave, and (c) SV-wave incident on a shale (medium 1) and sand
(medium 2) interface (Table B.1) 268

B.2 Plots (generated with the PSHSV program) showing displacement


amplitude, square root energy, and energy coefficients, and phase
changes as a function of incidence angle for a: (a) P-wave, (b) SH-
wave, and (c) SV-wave incident on a shale (medium 1) and gas sand
(medium 2) interface (Table B.1) 271

B.3 Plots (generated with the PSHSV program) showing displacement


amplitude, square root energy, and energy coefficients, and phase
changes as a function of incidence angle for a: (a) P-wave, (b) SH-
wave, and (c) SV-wave incident on a shale (medium 1) and coal
(medium 2) interface (Table B.1) 274

C.1 Map view of I-70 eastbound and westbound lanes (EBTravel, EBPass,
WBPass, and WBTravel) showing locations of YY component stacked
sections that are presented in this appendix relative to mapped mine
workings, observed roadway depressions and subsidence features, and
westbound lane land bridges for road stations: (a) 46700-47800, and
(b) 47800-48900. Road stations are in feet along the highway from the
western Guernsey County (Ohio) line 290

C.2 Eastbound travel lane (EBTravel) stacked time (uninterpreted) and


depth (with bedrock horizon interpreted) sections (YY component
data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-
48500, and (d) 48500-48900. CDP locations (CDP_X) correspond to
road stations (in feet from the western Guernsey County line) 292

xxxiv
C.3 Eastbound passing lane (EBPass) stacked time (uninterpreted) and
depth (with bedrock horizon interpreted) sections (YY component
data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-
48500, and (d) 48500-48900. CDP locations (CDP_X) correspond to
road stations (in feet along the highway from the western Guernsey
County line) 296

C.4 Westbound passing lane (WBPass) stacked time (uninterpreted) and


depth (with bedrock horizon interpreted) sections (YY component
data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-
48500, and (d) 48500-48900. CDP locations (CDP_X) correspond to
road stations (in feet along the highway from the western Guernsey
County line) 300

C.5 Westbound travel lane (WBTravel) stacked time (uninterpreted) and


depth (with bedrock horizon interpreted) sections (YY component
data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-
48500, and (d) 48500-48900. CDP locations (CDP_X) correspond to
road stations (in feet along the highway from the western Guernsey
County line) 304

C.6 Map view of I-70 eastbound and westbound lanes showing locations
and apparent dip directions of YY component-derived mine-related
faults (Table C.1), relative to mapped mine workings, observed
roadway depressions and subsidence features, and westbound lane land
bridges for road stations: (a) 46700-47800, and (b) 47800-48900.
Road stations are in feet from the western Guernsey County (Ohio)
line 309

D.1 Map of the eastbound lanes of I-70 (road stations 46900 to 48650)
showing the locations of cross-hole constant offset profile (COP) and
multiple offset gather (MOG) radar measurements between borings
(Table D.1), relative to the mapped locations of underground mine
workings. Due to map scale, mine room and pillar locations are
regarded as approximate. Road stations are given in feet from the
western county line 318

D.2 Cross-hole radar measurements in the I-70 study area for an example
case (19.0 m maximum depth): (left) constant offset profile (COP), and
(right) multiple offset gathers (MOG’s). For MOG’s the maximum Tx
and Rx vertical offset was 4.0 m 320

xxxv
D.3 Demonstration of constant offset profile data processing and imaging
flow: (a) field data after dewow correction, (b) after time-zero
correction, trace editing, truncation, bandpass filtering, amplitude
analysis, and trace normalization, (c) after direct arrival picks, and (d)
processed data with amplitude and velocity plots 325

D.4 Demonstration of multiple offset gather data processing and imaging


flow: (a) field data after dewow correction, merging, time-zero
correction, trace editing, truncation, bandpass filtering, and trace
normalization, (b) zoomed in look at MOG’s after direct arrival picks,
(c) plot of calculated velocity distribution between boreholes obtained
through inversion, and (d) velocity distribution plot after image
interpolation 326

D.5 Average EM-wave velocity and average absolute amplitude plots,


radar COP data, drill logs, and EM-wave velocity tomograms for I-70
eastbound travel lane road station range 46940 to 46968 (Table D.1,
Figure D.1): (a) wells GC307 and GC306, (b) wells GC305 and
GC307, and (c) wells B407G and B407H. A mosaic of the velocity
tomograms and well log information are shown along with interpreted
seismic reflection data and a mine map in Figure D.6 333

D.6 I-70 eastbound travel lane road station range 46940 to 46968: (a) EM-
wave velocity mosaic (Figure D.5), (b) S-wave reflection data, (c)
geologic cross-section from logs, and (d) approximate locations of
mine workings. Velocities (a) and drill logs (c) indicate that the SE
edge of the coal pillar beneath the road actually extends farther SE
than mapped in (d), and that the seismically imaged subsidence feature
(b) resulted from bedrock collapse into the mine room located
immediately south of this pillar 335

D.7 Average EM-wave velocity and average absolute amplitude plots,


radar COP data, drill logs, and EM-wave velocity tomograms for I-70
eastbound passing lane road station range 48304 to 48379 (Table D.1,
Figure D.1): (a) wells GC202 and GC201, (b) wells GC203 and
GC202, (c) wells GC204 and GC203, (d) wells GC205 and GC204, (e)
wells GC206 and GC205, and (f) wells GC207 and GC206 (no MOG
data acquired). A mosaic of the velocity tomograms and well log
information are shown along with interpreted seismic reflection data
and a mine map in Figure D.8 339

xxxvi
D.8 I-70 eastbound passing lane road station range 48300 to 48360: (a)
EM-wave velocity mosaic (Figure D.7), (b) S-wave reflection data, (c)
geologic cross-section, and (d) approximate locations of mine
workings. Velocities (a) indicate the mine-related bedrock subsidence
interpreted from seismic data (b) has occurred between the boreholes
and directly beneath the seismic line. Velocities (a) also extend the
west edge of disruption interpreted from seismic data an additional
several meters west 343

D.9 Average EM-wave velocity and average absolute amplitude plots,


radar COP data, drill logs, and EM-wave velocity tomograms for I-70
eastbound travel lane road station range 48304 to 48395 (Table D.1,
Figure D.1): (a) wells GC212 and GC211, (b) wells GC213 and
GC212, (c) wells GC214 and GC213, (d) wells GC215 and GC214, (e)
wells GC216 and GC215 (no MOG data acquired), (f) wells GC217
and GC216 (no MOG data acquired), and (g) wells B412E and GC217.
A mosaic of the velocity tomograms and well log information are
shown along with interpreted seismic reflection data and a mine map in
Figure D.10 347

D.10 I-70 eastbound travel lane road stations 48304 to 48400: (a) EM-wave
velocity mosaic (Figure D.9), (b) S-wave reflection data, (c) geologic
cross-section from logs, and (d) approximate locations of mine
workings. Velocities (a) support seismic data interpretation (b) of an
intact bedrock surface (stations 48304 to 48340), and bedrock and
overburden subsidence-related disruption (stations 48380 to 48408),
indicating this disruption occurred between the boreholes and directly
beneath the seismic line 351

D.11 Average EM-wave velocity and average absolute amplitude plots,


radar COP data, drill logs, and EM-wave velocity tomograms for I-70
eastbound travel lane road station range 48530 to 48640 (Table D.1,
Figure D.1): (a) wells B413H and GC301, (b) wells GC303 and
B413H, (c) wells B413F and GC304 (no MOG data acquired), (d)
wells GC302 and B413F, and (e) wells B413E and GC302 (no MOG
data acquired). A mosaic of the velocity tomograms and well log
information are shown along with interpreted seismic reflection data
and a mine map in Figure D.12 355

xxxvii
D.12 I-70 eastbound travel lane road station range 48530 to 48620: (a) EM-
wave velocity mosaic (Figure D.11), (b) S-wave reflection data, (c)
geologic cross-sections from logs, and (d) approximate locations of
mine workings. Velocities (a) suggest that the coal pillar beneath the
road (stations 48605 to 48615) actually extends farther south than
mapped in (d), and that the seismically imaged subsidence feature (b)
resulted from bedrock collapse into the mine room located just south of
this pillar 358

xxxviii
CHAPTER 1

INTRODUCTION

1.1 Background

Seismic reflection is a geophysical technique that can allow physical properties of

earth materials to be determined and images of the subsurface to be constructed.

Although the analysis of compressional waves (P-waves) reflected from geologic layers

originated in the early 1900’s for petroleum exploration, shear waves (S-waves) have

only recently been employed for such purposes. The recognition of potential advantages

of S-waves has led to a recent interest throughout the petroleum industry in the testing,

and simultaneous acquisition, processing, and interpretation of multiple components of

the seismic wavefield (multicomponent data analysis). Three previous areas of research

(i.e. focused on reflection imaging methods, digital recording, and 3D technologies) have

led to important developments in the capability of seismic reflection, and

multicomponent data analysis is currently regarded by the petroleum industry as the

fourth area of research that will produce significant advances (Freedman, 1999).

As a result of instrument and computer technology advances, seismic reflection

(predominately P-wave methods) has been applied during the past two decades towards

the solution of shallow earth (0 - 100 m deep) engineering and environmental problems

1
(Hunter et al., 1984; Miller and Steeples, 1990; Bachrach and Nur, 1998; van der Veen

and Green, 1998). Potential advantages of multicomponent reflection data testing and

analysis for improving shallow subsurface imaging and characterization potential exist in

theory, and it was therefore concluded from a recent Department of Energy sponsored

workshop that research concerning the applicability of shallow multicomponent seismic

reflection methods is necessary (Steeples et al., 1997). Multicomponent information has

not typically been obtained and analyzed during shallow studies due to increased

complexity and costs associated with data acquisition and analysis (Kendall and Davis,

1996), and few reports documenting shallow multicomponent reflection research exist.

A section of Interstate 70 (I-70) in eastern Ohio collapsed recently (during 1995)

into an underground mine network, and concern regarding the stability of an undermined

2200 ft section of the (four-lane) roadway continues. In order to address the concern for

subsidence problems along I-70, and at the same time test and develop methodologies for

characterizing near-surface materials, experimental, high-resolution shear-wave and

multicomponent reflection data sets were acquired along the roadway during 1999 and

2001 respectively. During 2002, cross-hole radar data were also acquired in the study

area in order to further address these objectives. These data were all acquired as part of a

Federal Highway Administration (FWHA) and Ohio Department of Transportation

(ODOT) sponsored project.

Results that are presented in this dissertation are sometimes discussed within the

context of the I-70 mine-subsidence project. The methods developed and the conclusions

reached through this research are applicable however, to roadway investigations and

near-surface characterization studies with different objectives conducted elsewhere.

2
1.2 Research Objectives

This dissertation documents a portion of a study that is focused on developing

geophysical methods for improving near-surface characterization potential, understanding

coal mine-related subsidence mechanisms, and locating areas with a high risk for future

mine-related surface collapse along I-70 in eastern Ohio. The portion includes my major

contributions to the study, accomplished through research conducted to meet the

following primary objectives:

1) Investigation of the potential of acquired multicomponent reflection

information for allowing the estimation of insitu elastic properties of media.

2) Design of effective and efficient analysis sequences (processing, imaging, and

interpretation) for an acquired large volume of 2D high-resolution

multicomponent seismic reflection data.

3) Demonstration of improvements in near-surface characterization potential

achieved through the testing and analysis of multiple seismic reflection data

components/body wave-types.

4) Investigation of the potential of acquired multicomponent reflection data for

allowing the construction of converted wave images, and discussion of

analysis and processing sequences for addressing converted wave imaging

during near-surface characterization studies.

5) Determination of the potential of seismic reflection data for allowing the

delineation of subsurface areas where mine-related subsidence activity has

occurred, in order to locate surface areas with a high risk for future collapse.

3
6) Characterization of the I-70 subsurface (in terms of mine-related subsidence

activity and relative location risk for future surface failure) using seismic

reflection, through the implementation of developed data analysis sequences.

7) Investigation of the potential of cross-hole radar methods for providing useful

subsurface information for mine-related subsidence studies.

This dissertation also documents research conducted to meet the following

secondary objectives:

1) A study of seismic wave scattering to accomplish the presentation of plane

wave scattering equations using an original and intuitive nomenclature.

2) Development of a program that calculates reflection/refraction and energy

coefficients for elastic waves incident on an elastic media planar interface.

3) Derivations of equations that define an SH-wave intromission angle and

describe its conditions of occurrence.

1.3 Significance of Research and Previous Work

The significance of the primary objectives addressed in this dissertation is briefly

demonstrated in the following sub-sections, which provide an overview of relevant

background information and summaries of previous work. Further elaborations on

certain issues are subsequently provided in this dissertation. The significance of the

secondary objectives addressed in this dissertation is discussed subsequently in the

dissertation.

4
1.3.1 Shallow Multicomponent Seismic Reflection Data Analysis

There is currently a widespread interest in testing and recording multiple

components of the seismic wavefield throughout the petroleum and engineering

industries. Deep earth studies have determined that different components can allow

project objectives to be more effectively accomplished depending upon conditions, and

that in some cases (dependent upon numerous factors) multicomponent reflection data

can provide more geological information than single component data alone. For instance,

S-wave reflection data can offer potential benefits relative to P-wave data (e.g. related to

component noise sensitivity, differences in wave type resolution, or differences in wave

type sensitivity to physical property variations), or when used concurrently with P-wave

data, such as lithologic interpretation and material property estimation (Stewart and

Gaiser, 1999), and hydrocarbon indicator evaluation (Ensley, 1985). Converted wave

analysis has been shown to be potentially useful for obtaining deep earth S-wave

information and quality converted-mode images (Garotta, 2000). Anisotropy evaluation

is also possible through analysis of split S-wave components, which can been useful for

determining fracture/conduit orientations (Crampin, 1985). Although more complicated,

expensive, and time-consuming than recording and analyzing a single P-wave data

component, there are potential benefits of considering additional components of the

seismic wavefield during shallow reflection studies (only single component P-wave data

have traditionally been acquired during near-surface seismic reflection studies).

5
1.3.1.1 Common-Mode Seismic Reflection Analysis

Successful applications of S-wave reflection data towards shallow earth problems

have been demonstrated (Carr et al., 1998; Harris et al., 2000). However, the number of

shallow seismic reflection reports concerning S-waves is small relative to the number

concerning P-waves, and most S-wave studies to date have considered only a single S-

wave component. Further research that documents capabilities and limitations of the S-

wave reflection technique is necessary in order to increase understanding regarding the

practicality of using S-wave reflections for near-surface characterization.

The potential benefits of testing, recording and analyzing P- and S-wave

reflections concurrently for improving shallow earth characterization potential have not

yet been fully recognized or demonstrated. Few reports concerning the concurrent use of

P- and S-wave reflections exist, but existing reports contain promising results.

Hasbrouck (1990) discussed (without published data), and Goforth and Hayward (1992)

discussed sites where saturated overburden and a target bedrock horizon had nearly the

same P-wave velocity (therefore, P-waves imaged the water table at both sites but were

unable to image lithologic boundaries). Clark et al. (1994) modeled P- and S-wave

responses for a shallow geologic sequence. Their synthetic data demonstrated that for

their modeled sequence, only the combined usage of P- and S-waves would allow the

detection of all modeled geologic boundaries. Holzschuh (2002) demonstrated using

field data, that as a result of their lower velocity, it can be possible to achieve higher

resolution reflection information during near surface studies (relative to that obtainable

using P-waves) by generating and recording S-waves.

6
Because P- and S-waves travel at different speeds and respond to changes in

elastic moduli and density differently, it is reasonable to expect that during shallow

reflection surveys, achievable resolution may be quite different for P- and S-waves, P-

and S-wave reflection coefficients may be different, and P- and S-wave reflections may

be recorded under different (i.e. more or less favorable) noise conditions. It is also

possible in theory to estimate insitu elastic properties of near-surface materials using both

common-mode P- and S-wave reflection information, although I am aware of only a few

previous research reports concerning this issue. Although the previously mentioned

studies discussed certain benefits for shallow earth studies that may be obtained through

concurrent P- and S-wave data testing and analysis, possible benefits remain to be

recognized and explained, demonstrated within the context of different applications, and

using field data acquired in different geologic settings.

1.3.1.2 Converted-Mode Seismic Reflection Analysis

The subject of converted wave recording and analysis has received a great deal of

attention in the petroleum industry in recent years (Stewart and Gaiser, 1999). The

majority of petroleum industry converted-mode research efforts have focused on

applications in marine environments. Marine converted wave acquisition is a fairly new

strategy, and prior to August 2000 only 145 surveys of this type had been conducted

worldwide (Stefanic, 2000). S-waves cannot be directly propagated into an underlying

geologic medium from a source on the water surface, but converted wave recognition

technology has made it possible to obtain S-wave reflectivity and travel time information

in marine environments. Previously, S-wave reflectivity could only be estimated in

7
marine environments (without necessary log data) through AVO inversion of P-wave

amplitudes (Castagna, 1993). It has also been shown that quality converted wave images

can be produced in situations where P-wave reflection quality is poor, for instance where

P-waves traveling upward through a gas cloud are highly attenuated (Granli et al., 1999).

Converted wave research (conducted by the petroleum industry) on land has been

promising (Eliasata and Michelena, 1995; Macrides and Kelamis, 2000), despite being

more challenging in certain respects than in marine environments (for instance noise

problems are typically more severe on land). It has been demonstrated that converted

wave analysis in land-based studies can potentially improve deep earth characterization.

However, no reports concerning shallow (0-100 m deep) earth converted wave studies

have appeared in the literature, despite there being potential advantages of shallow

converted wave analysis (e.g. it is possible in theory to obtain S-wave information using

only a P-wave source, or vice versa, and it is possible in theory to estimate insitu elastic

properties of near-surface media through converted wave analysis). At this point the

recording of converted wave components is not typically considered as a valid option

during shallow reflection survey design. This is due in large part to the lack of successful

converted wave investigations (i.e. limited understandings of when converted wave

imaging may work, and how to analyze mode converted reflections). Research that

reports investigations into the potential applicability of converted waves for near-surface

problems is therefore important. Additionally, no standard reference for approaching

converted wave analysis and imaging has been published, and research documenting data

analysis approaches for shallow multicomponent reflection data is necessary.

8
1.3.2 Geophysical Detection of Subsurface Mine-Related Subsidence Activity

When roadways such as I-70 in eastern Ohio are subject to subsidence, extreme

costs can be incurred, the risk for human casualties can be high, and industrial

productivity can be disrupted. In addition to dewatering of aquifers and organic soils,

natural sedimentary compaction, tectonic events, oil and gas reserve extraction, bedrock

dissolution, and thawing permafrost, a primary cause of roadway subsidence is

overburden collapse into underground mines. While non-mining related subsidence is

common in the United States (see previous sentence for examples), the greatest annual

dollar loss as a result of subsidence damage is mine-related (Isphording, 1992). Mine-

related subsidence has accounted for 20 percent of the total subsidence by land area in the

United States, with the majority of this percentage related to underground coal mines

(Galloway et al., 1999).

The recognition of broad regions along roadways where potential mine-related

subsidence risk exists can be accomplished through: 1) the identification of underground

mine locations, 2) consideration of information obtained through drilling, and 3)

assessment of previous subsidence that may have occurred. It is difficult within a

subsidence-prone region however, to identify precise surface locations having a high risk

for future collapse. In order to do so, mechanisms that are responsible for subsidence

must first be assessed, and a thorough study area subsurface investigation must be

conducted. The National Research Council (NRC, 1991) has therefore recommended

that research be conducted which focuses on the development, evaluation, and

application of innovative methods for addressing subsidence risk, so that studies can be

addressed from a cost-effective, damage prevention standpoint.

9
As the physical properties of subsidence features (e.g., displaced strata, fractures,

and voids) can have a contrast in physical properties with surrounding materials,

geophysical measurements offer the potential to locate areas along the ground surface

with a high risk for future subsidence-related failure. Numerous geophysical techniques

have the potential in theory, and have been shown in practice, to be capable of detecting

variations in physical properties associated with subsidence activity, although no

technique will be successful under all conditions (Fisher, 1971; Munk and Sheets, 1997).

Few reports have demonstrated that seismic reflection (P-wave) measurements

offer a means for detecting and imaging features that have developed from subsidence

activity beneath roadways (Steeples et al., 1986; Miller and Xia, 2002). Detection of

subsurface subsidence activity using seismic reflection (and using other geophysical

techniques) is currently regarded as challenging (Steeples et al., 1997), and the conditions

under which seismic reflection methods provide a high potential for success are not well

understood. Reports concerning the potential of cross-hole radar methods for allowing

the detection of subsurface features associated with mine-related subsidence activity are

non-existent to date. Research documenting (and furthering the understanding of) the

applicability of high-resolution seismic reflection and cross-hole radar methods towards

subsidence-related problems along roadways is therefore important.

1.4 Organization of Document

This dissertation is organized as follows:

Chapter 1 – The motivation for this research is discussed, and background

information and an overview of previous work relevant to this research are provided.
10
Chapter 2 – An overview of seismic wave reflection, refraction, and mode

conversion in elastic media is provided. Using an original and intuitive nomenclature,

plane wave scattering equations for calculating reflection/refraction coefficients, square

root energy coefficients, and energy coefficients from P- SH- and SV-waves incident on

planar interfaces between isotropic media are presented and discussed.

Chapter 3 – Details regarding the acquisition and pre-processing of the

experimental seismic reflection data that were used for this dissertation are provided.

Chapter 4 – Results from research conducted to compare the effectiveness of

near-surface P- and S-wave reflection surveys for mapping features in the shallow

subsurface are presented. Factors influencing the imaging potential (e.g., component

noise mode sensitivity, reflection coefficients, resolution, and subsurface conditions) of

common-mode reflection information are described, and illustrations using analyses of

acquired field multicomponent data gathers, processed sections, calculations, and

modeling are provided. The effectiveness of different common-mode reflection wave-

types/data components for allowing the imaging of horizons, the detection of subsurface

horizon discontinuities, and the estimation of media elastic parameters is discussed.

Chapter 5 – Results from research conducted to test and evaluate the potential for

near-surface mode-converted (P-S and S-P) seismic reflection imaging are presented.

The topic is addressed experimentally using high-resolution 9C field reflection data

acquired over relatively flat-lying geology, and numerically/theoretically using Zoeppritz

equation solutions and elastic-wave modeling. Factors influencing mode-conversion

imaging potential, which differ between deep-earth and near-surface multicomponent

11
reflection studies are also discussed. Analysis methods are presented that can be used for

assessing converted-wave imaging potential in future near-surface reflection studies.

Chapter 6 – An analysis (processing, imaging, and interpretation) flow to serve as

a basis for the processing of SH-wave reflection data acquired in the study area is

described. Factors influencing analysis workflow development (acquired data

characteristics and volume, subsurface conditions, and project objectives) are discussed

and demonstrated through modeling, and illustrations of the affects of applied processes

on data are provided. The effectiveness of SH-wave reflection data and developed

analysis procedures for allowing the delineation of subsurface areas where subsidence

processes have been active is demonstrated. The analysis flow is applied to 3 lines of S-

wave (crossline-crossline, SH-SH) reflection data acquired near a previous roadway

collapse in the I-70 study area, and is shown to allow the location of two areas along the

roadway where high risk for future mine-related surface failure exists.

Chapter 7 – Derived equations that define an incidence angle for which the SH-

wave reflection coefficient is zero (angle of intromission) are presented. Conditions (in

terms of media density and shear-velocity relationships) under which a SH-wave

intromission angle will occur in theory are described with example plots, and the

potential application of inferring such an angle from field data is discussed. An overview

of P-wave intromission angle occurrence in acoustic media is presented, and short source

code for calculating and plotting acoustic P-wave reflection coefficient magnitudes and

phase (as a function of incident angle) is listed.

Chapter 8 – Conclusions reached through this research are summarized.

12
Appendix A – Discussions regarding the I-70 study area geology, coal mine-

related subsidence history, and coal mine-related subsidence mechanisms are presented.

Appendix B – A developed computer program (PSHSV) is described and listed,

that calculates and plots (as a function of incident angle) displacement amplitude

coefficients, normalized square root energy ratios, energy coefficients, and phase changes

for plane P- SH- and SV-waves incident on a planar interface between elastic media.

Functionality of the program is demonstrated within the context of a hypothetical gas

reservoir detection problem.

Appendix C – Processed and interpreted S-wave YY component (crossline-

crossline, SH-SH) time and depth sections are presented for the entire I-70 study area

(eastbound and westbound passing and travel lanes). Analyzed YY data are presented, as

this component was determined through research to provide the highest potential of

acquired data components, for allowing areas of the subsurface where subsidence

processes have been active to be located. Numerous locations along the I-70 lanes,

having a relatively high risk for future mine-related surface failure are identified, based

on significant mine-related disruptions of the bedrock horizon.

Appendix D – Results from research conducted to test the ability of cross-hole

radar methods for providing useful information for mine-related subsidence studies are

presented. It is demonstrated through analyses of data acquired near several subsidence

features imaged using seismic reflection, that average EM-wave velocity and amplitude

plots along with processed velocity tomograms, can be used to infer the presence of

mine-related subsidence activity within near-surface media between boreholes. Cross-

hole radar data are also shown to be useful for reducing uncertainty regarding the

13
locations/extent of mine rooms and pillars, and seismically imaged subsidence features.

Data acquisition considerations and data analysis workflows developed for cross-hole

radar measurements are presented.

14
CHAPTER 2

SEISMIC WAVE SCATTERING THEORY

2.1 Introduction

When a wave is incident on an interface separating materials with an impedance

contrast, scattered (reflected and / or refracted / transmitted) waves are generated. Ray

angles of scattered waves are dependent on the angle of incidence and the velocity

contrast across the interface. Amplitude and energy partitioning between scattered

components is strongly dependent upon the angle of incidence and both the velocity and

density contrasts (impedance contrasts) across the discontinuity. In the case of an

interface between two solids, media on either side respond to an incident seismic wave as

if they were welded together (Castagna, 1993). Media on opposite sides of the interface

would otherwise separate from one another, occupy the same space at the same time, or

slide past one another, and these types of behavior do not occur at an interface between

solids as a result of artificially generated incident seismic energy (Telford et al., 1990).

Displacement amplitudes remain continuous at a welded interface, and stress tensor

components corresponding to the traction across the interface also remain continuous

(Rüger, 2002), through reflection, refraction and mode-conversion of incident seismic

waves.

15
2.2 P- SH- and SV-Waves: Normal Incidence

For P-waves with normal incidence, no components of shear are introduced at an

interface, while for S-waves (SH- or SV-type) with normal incidence, no components of

compression are introduced at an interface. Although at large angles of incidence mode-

conversion can become significant for incident P- and SV-waves, for most geologic

conditions at small incidence angles, little energy undergoes mode-conversion, and the

curves describing energy partitioning do not change significantly (Tatham and

McCormack, 1991). In petroleum exploration studies it can often be assumed that waves

reaching the earth’s surface with ray paths that are within 20 degrees of normal were

reflected vertically from a horizontal interface in the earth (Telford et al., 1990).

Assuming plane waves having normal incidence, seismic amplitude and energy

partitioning can be described using equations that consider only the impedance contrast at

an interface. For P-waves with normal incidence, the reflection and refraction

coefficients are given in the following set of equations (nomenclature modified from

Sheriff and Geldart, 1982):

P1P1vert = A1 / A0 = ( ρ 2Vp2 - ρ 1Vp1) / ( ρ 2Vp2 + ρ 1Vp1) =

(Zp2 - Zp1) / (Zp2 + Zp1)

P1P2vert = A2 / A0 = (2 ρ 1Vp1) / ( ρ 1Vp1 + ρ 2Vp2) =

(2Zp1) / (Zp2 + Zp1) (2.1)

where P1P1vert and P1P2vert are the normal incidence P-wave reflection and refraction

coefficients, A0 , A1, and A2 are the amplitudes of maximum particle displacement of the

incident, reflected and refracted waves, Vp1 and Vp2 are the incident and refracted media

16
Figure 2.1. Particle motion associated with wave types (P, SH, and SV). P-wave particle
motion (left) is in the direction of propagation (indicated by ray path) and within the
plane of incidence (defined by source, reflection point, and receiver). SH-wave particle
motion (middle) is perpendicular to the direction of propagation, and is within a plane
that is perpendicular to the plane of incidence. SV-wave particle motion (right) is within
the plane of incidence, but perpendicular to the propagation direction.

P-wave interval velocities, ρ1 and ρ2 are the incident and refracted media densities, and

Zp1 and Zp2 are the incident and refracted media compressional impedances.

Nomenclature used in this chapter and in the PSHSV program (Appendix B) for

incident and scattered wave types and ray angles (Figure 2.1), and for calculated

coefficients / ratios, is such that the first capital letter of terms indicates the incident wave

type, and the second capital letter indicates the scattered wave type. Subscripts

associated with capital letters indicate the medium of the traveling wave. A subscript of

1 indicates the wave is traveling in the incident medium, while a subscript of 2 indicates

the wave is traveling in the refracted medium. For example, P1P2 indicates a refracted P-

wave in medium 2 resulting from an incident P-wave in medium 1 (Figure 2.2).

17
18
Figure 2.2. Nomenclature used for incident and scattered wave types and ray angles with normal, and for calculated coefficient
and ratio terms. P1P2 for example, signifies a refracted P-wave in medium 2 resulting from an incident P-wave in medium 1.
Rays that are shown represent the possible scattered waves from the planar interface (which is perpendicular to the plane
containing the rays) separating medium 1 and medium 2 (which are isotropic solids).
For S-waves (SH- or SV-type) with normal incidence, the reflection and

refraction coefficients are given in the following set of equations (nomenclature modified

from Shearer, 1999):

S1S1vert = A1 / A0 = ( ρ 1Vs1 - ρ 2Vs2) / ( ρ 1Vs1 + ρ 2Vs2) =

(Zs1 - Zs2) / (Zs1 + Zs2)

S1S2vert = A2 / A0 = (2 ρ 1Vs1) / ( ρ 1Vs1 + ρ 2Vs2) =

(2Zs1) / (Zs1 + Zs2) (2.2)

where S1S1vert and S1S2vert are the normal incidence S-wave reflection and refraction

coefficients, Vs1 and Vs2 are the incident and refracted media S-wave interval velocities,

and Zs1 and Zs2 are the incident and refracted media shear impedances.

The normal incidence P- and S-wave reflection coefficient equations are different

with regards to sign convention. For a similar change in sign of shear and compressional

impedances across an interface, the calculated normal incidence P- and S-wave reflection

coefficients will have opposite sign, because P-wave particle motions are measured

relative to the ray direction and S-wave particle motions are not (Shearer, 1999). From

the previous equations it is seen that when shear impedance increases across an interface,

the calculated reflection coefficient at normal incidence will be negative. Conversely,

when compressional impedance increases across an interface, the calculated reflection

coefficient will be positive.

According to the law of conservation of energy (and assuming no energy loss

along raypaths), the sum of the energy contained within the scattered components (energy

carried away from the interface) must equal the incident energy (i.e. equal to 1.0):

19
E0 = E1 + E2 (2.3)

where E0 is the incident energy, E1 is the reflected energy, and E2 is the refracted energy.

For normal incidence, the ratios of reflected energy to incident energy, and

refracted energy to incident energy, are proportional to the displacement amplitude

coefficients squared and weighted by the media impedances (Sheriff and Geldart, 1982).

The ratio of reflected energy to incident energy at normal incidence is therefore

equivalent to the displacement amplitude reflection coefficient squared. This is shown by

the following set of equations, written for reflected waves resulting from both P- and S-

waves with normal incidence:

E1 / E0 = (A1 / A0)2(Zp1 / Zp1) = (P1P1vert)2

E1 / E0 = (A1 / A0)2 (Zs1 / Zs1) = (S1S1vert)2 (2.4)

The ratio of refracted energy to incident energy is not equal to the square of the

normal incidence displacement amplitude refraction coefficient. This is shown by the

following set of equations, written for refracted waves resulting from both P- and S-

waves with normal incidence:

E2 / E0 = (A2 / A0)2(Zp2 / Zp1) = (P1P2vert)2(Zp2 / Zp1)

E2 / E0 = (A2 / A0)2(Zs2 / Zs1) = (S1S2vert)2(Zs2 / Zs1) (2.5)

As seen from the preceding equations, by summing the square of the reflected to

incident displacement amplitude ratio and the square of the refracted to incident

displacement amplitude ratio, a quantity equivalent to the incident energy is not obtained.

It is therefore often desirable to calculate square root energy coefficients. These

quantities are obtained from the preceding equations, and are given in the following set of

equations for both P- and S-waves with normal incidence:

20
ENP1P1vert = (E1 / E0)1/2 = P1P1vert

ENS1S1vert = (E1 / E0)1/2 = S1S1vert

ENP1P2vert = (E2 / E0)1/2 = (P1P2vert)(Zp2 / Zp1)1/2

ENS1S2vert = (E2 / E0)1/2 = (S1S2vert)(Zs2 / Zs1)1/2 (2.6)

where ENP1P1vert and ENS1S1vert are the square root energy coefficients for reflected P-

and S-waves, and ENP1P2vert and ENS1S2vert are the square root energy ratios for refracted

P- and S-waves.

By squaring the normal incidence square root energy coefficients, the normal

incidence energy coefficients are obtained. The sum of the reflected and refracted P-

wave energy coefficients, and the sum of the reflected and refracted S-wave energy

coefficients, will both be equal to 1.0. The energy coefficients resulting from P- and S-

waves at normal incidence are defined in the following set of equations:

EP1P1vert = (ENP1P1vert)2

ES1S1vert = (ENS1S1vert)2

EP1P2vert = (ENP1P2vert)2

ES1S2vert = (ENS1S2vert)2 (2.7)

where EP1P1vert and ES1S1vert are the energy coefficients for reflected P- and S-waves, and

EP1P2vert and ES1S2vert are the energy coefficients for refracted P- and S-waves.

2.3 SH-Waves: Non-Normal Incidence

When an interface between two homogenous and isotropic solids is parallel to the

polarization of an incident SH-wave, only scattered SH-waves are generated (Figure 2.2),

regardless of the incident angle. No mode-conversion to P- or SV-waves will occur, as


21
the necessary compressional and shear stresses are not induced on the point of incidence.

The ray angles of scattered SH-waves are dependent on the angle of incidence and the

shear velocity contrast across the interface. The energy partitioning between scattered

components is dependent on the angle of incidence and both the shear velocity and

density contrasts (shear impedance contrast) across the interface. For SH- waves with

non-normal angles of incidence, the reflection and refraction displacement amplitude

coefficients are given in the following set of equations (nomenclature modified from

Shearer, 1999):

SH1SH1 = (Zs1cosșSH1SH1 - Zs2cosșSH1SH2) /

(Zs1cosșSH1SH1 + Zs2cosșSH1SH2)

SH1SH2 = (2Zs1cosșSH1SH1) / (Zs1cosșSH1SH1 + Zs2cosșSH1SH2) (2.8)

where SH1SH1 and SH1SH2 are the SH-wave displacement amplitude reflection and

refraction coefficients. The reflection and refraction angle cosines depend upon both the

horizontal and vertical components of the slowness vector:

pp = (u2 - Ș2)1/2 = usinș (2.9)

Ș = (u2 - pp2)1/2 = ucosș (2.10)

where u is the slowness, pp is the horizontal slowness, Ș is the vertical slowness, and ș is

the reflected or refracted ray angle from normal. In this chapter, horizontal slowness is

assigned the variable pp for the case of an incident SH- or SV-wave, and the variable p

for the case of an incident P-wave.

In terms of media velocities, the reflection and refraction angle cosines can be

written in the following set of equations as:

cosșSH1SH1 = Vs1(((1 / (Vs12)) - (pp2))1/2)

22
cosșSH1SH2 = Vs2(((1 / (Vs22)) - (pp2))1/2) (2.11)

where șSH1SH1 is the angle of SH-wave reflection, and șSH1SH2 is the angle of SH-wave

refraction.

The SH-wave square root energy coefficient equations for non-normal incidence

are obtained in a similar manner as those that were obtained for the aforementioned

normal incidence case. For the non-normal incidence case however, angle cosines in

addition to media impedances must also be considered in order to obtain the desired

quantities. The SH-wave square root energy coefficients for non-normal incidence are

given in the following set of equations:

ENSH1SH1 = SH1SH1

ENSH1SH2 = (SH1SH2 )((Zs2cosșSH1SH2) / (Zs1cosșSH1SH1))1/2 (2.12)

where ENSH1SH1 and ENSH1SH2 are the normalized square root energy ratios for

reflected and refracted SH-waves.

The incident SH-wave energy coefficients are defined by the following set of

equations:

ESH1SH1 = (ENSH1SH1)2

ESH1SH2 = (ENSH1SH2)2 (2.13)

where ESH1SH1 and ESH1SH2 are the energy coefficients for reflected and refracted SH-

waves resulting from an incident SH-wave.

2.4 P- and SV-Waves: Non-Normal Incidence

For P- or SV-waves with non-normal incidence, mode-conversion occurs at the

point of incidence because the interface is being both compressed and sheared. From an
23
incident P- or SV-wave, there are four different scattered wave types that can be

generated at a welded interface in order to maintain continuity of displacement and

traction (Figure 2.2): 1) a reflected P-wave, 2) a reflected SV-wave, 3) a refracted P-

wave, and 4) a refracted SV-wave. The energy and amplitude partitioning between

scattered components resulting from incident P- and SV-waves can be determined and

described through solutions to the equations (and those equations later derived from the

equations) developed by Knott (1899) and Zoeppritz (1919).

The ray angles of scattered waves resulting from P- and SV-waves with non-

normal incidence are dependent on the angle of incidence and the contrast in

compressional and shear velocities across the interface. The amplitude and energy

partitioning between each of the scattered components depends upon the media

compressional and shear impedances on either side of the interface and the angle of

incidence. The displacement amplitude (Zoeppritz) coefficients of the scattered waves

can be determined by solving simultaneous linear equations with real or complex

coefficients that contain the unknown P- and SV-wave amplitudes (Richter, 1958). It is

difficult to gain physical insight simply from looking at these equations as to exactly how

the variation of a particular parameter will affect solutions. Because exact expressions

for the coefficients are still not very intuitive in this regard, approximations that work

well with certain assumptions for certain conditions, and more easily allow physical

insight into the effects of variations in individual parameters have been formulated

(Bortfeld, 1961; Aki and Richards, 1980; Shuey, 1985).

24
2.4.1 Incident P-Waves

For incident P-waves, the reflection and refraction displacement amplitude

coefficients are explicitly expressed in the following set of equations (nomenclature

modified from Aki and Richards, 1980):

P1P1 = ((b(cosșP1P1 / Vp1) - c(cosșP1P2 / Vp2))F -

(a+d(cosșP1P1 / Vp1)(cosșP1SV2 / Vs2))Hp2) / D

P1SV1 = - (2(cosșP1P1 / Vp1)(ab+cd(cosșP1P2 / Vp2)*

(cosșP1SV2 / Vs2))pVp1 / Vs1) / D

P1P2 = (2ρ1(cosșP1P1 / Vp1)F(Vp1 / Vp2)) / D

P1SV2 = (2ρ1(cosșP1P1 / Vp1)Hp(Vp1 / Vs2)) / D (2.14)

where P1P1 and P1P2 are the reflected and refracted P-wave displacement amplitude

reflection and refraction coefficients, and P1SV1 and P1SV2 are the reflected and refracted

SV-wave displacement amplitude reflection and refraction coefficients. Variables and

cosine-dependent terms used in the preceding P-wave displacement amplitude coefficient

equations are given in the following set of equations:

a = ρ2(1 - 2Vs22p2) - ρ1(1 - 2Vs12p2)

b = ρ2(1 - 2Vs22p2) + 2ρ1Vs12p2

c = ρ1(1 - 2Vs12p2) + 2ρ2Vs22p2

d = 2(ρ2Vs22 - ρ1Vs12)

E = b(cosșP1P1 / Vp1) + c(cosșP1P2 / Vp2)

F = b(cosșP1SV1 / Vs1) + c(cosșP1SV2 / Vs2)

G = a-d(cosșP1P1 / Vp1)(cosșP1SV2 / Vs2)

H = a-d(cosșP1P2 / Vp2)(cosșP1SV1 / Vs1)


25
D = EF + GHp2 (2.15)

Reflection and refraction angle cosines used in the preceding equations can be written in

the following set of equations as:

cosșP1P1 = Vp1(((1 / (Vp12)) - (p2))1/2)

cosșP1SV1 = Vs1(((1 / (Vs12)) - (p2))1/2)

cosșP1P2 = Vp2(((1 / (Vp22)) - (p2))1/2)

cosșP1SV2 = Vs2(((1 / (Vs22)) - (p2))1/2) (2.16)

where șP1P1 is the angle of P-wave reflection, șP1SV1 is the angle of SV-wave reflection,

șP1P2 is the angle of P-wave refraction, șP1SV2 is the angle of SV-wave refraction, and p is

the horizontal slowness.

The incident P-wave square root energy coefficients are given in the following set

of equations:

ENP1P1 = P1P1

ENP1SV1 = (P1SV1)((Vs1cosșP1SV1) / (Vp1cosșP1P1))1/2

ENP1P2 = (P1P2)((ρ2Vp2cosșP1P2) / (ρ1Vp1cosșP1P1))1/2

ENP1SV2 = (P1SV2)((ρ2Vs2cosșP1SV2) / (ρ1Vp1cosșP1P1))1/2 (2.17)

where ENP1P1 and ENP1P2 are the square root energy coefficients for reflected and

refracted P-waves, and ENP1SV1 and ENP1SV2 are the square root energy coefficients for

reflected and refracted SV-waves.

The incident P-wave energy coefficients are defined by the following set of

equations:

EP1P1 = (ENP1P1)2

26
EP1SV1 = (ENP1SV1)2

EP1P2 = (ENP1P2)2

EP1SV2 = (ENP1SV2)2 (2.18)

where EP1P1 and EP1P2 are the energy coefficients for reflected and refracted P-waves,

and EP1SV1 and EP1SV2 are the energy coefficients for reflected and refracted SV-waves.

2.4.2 Incident SV-Waves

For incident SV-waves, the reflection and refraction displacement amplitude

coefficients are explicitly expressed in the following set of equations (nomenclature

modified from Aki and Richards, 1980):

SV1SV1 = - ((bb(cosșSV1SV1 / Vs1) - cc(cosșSV1SV2 / Vs2))EE -

(aa + dd(cosșSV1P2 / Vp2)(cosșSV1SV1 / Vs1))GGpp2) / DD

SV1P1 = - (2(cosșSV1SV1 / Vs1)(aabb + ccdd((cosșSV1P2) / Vp2)*

((cosșSV1SV2) / Vs2))pp(Vs1 / Vp1)) / DD

SV1SV2 = 2ρ1(cosșSV1SV1 / Vs1)EE(Vs1 / Vs2) / DD

SV1P2 = - 2(ρ1(cosșSV1SV1 / Vs1)GGpp(Vs1 / Vp2)) / DD (2.19)

where SV1SV1 and SV1SV2 are the reflected and refracted SV-wave displacement

amplitude reflection and refraction coefficients, and SV1P1 and SV1P2 are the reflected

and refracted SV-wave displacement amplitude reflection and refraction coefficients.

Variables and cosine-dependent terms used in the preceding SV-wave displacement

amplitude coefficient equations, are defined in the following set of equations:

aa = ρ2(1 - 2Vs22pp2) - ρ1(1 - 2Vs12pp2)

bb = ρ2(1 - 2Vs22pp2) + 2ρ1Vs12pp2


27
cc = ρ1(1 - 2Vs12pp2) + 2ρ2Vs22pp2

dd = 2(ρ2Vs22 - ρ1Vs12)

EE = bb(cosșSV1P1 / Vp1) + cc(cosșSV1P2 / Vp2)

FF = bb(cosșSV1SV1 / Vs1) + cc(cosșSV1SV2 / Vs2)

GG = aa-dd(cosșSV1P1 / Vp1)(cosșSV1SV2 / Vs2)

HH = aa-dd(cosșSV1P2 / Vp2)( cosșSV1SV1 / Vs1)

DD = EEFF + GGHHpp2 (2.20)

Reflection and refraction angle cosines used in the preceding equations can be written in

the following set of equations as:

cosșSV1SV1 = Vs1(((1 / (Vs12)) - (pp2))1/2)

cosșSV1P1 = Vp1(((1 / (Vp12)) - (pp2))1/2)

cosșSV1SV2 = Vs2(((1 / (Vs22)) - (pp2))1/2)

cosșSV1P2 = Vp2(((1 / (Vp22)) - (pp2))1/2) (2.21)

where ș SV1SV1 is the angle of SV-wave reflection, șSV1P1 is the angle of P-wave reflection,

șSV1SV2 is the angle of SV-wave refraction, șSV1P2 is the angle of P-wave refraction, and pp

is the horizontal slowness.

The incident SV-wave square root energy coefficients are given in the following

set of equations:

ENSV1SV1 = SV1SV1

ENSV1P1 = (SV1P1)((Vp1cosșSV1P1) / (Vs1cosșSV1SV1))1/2

ENSV1SV2 = (SV1SV2)((ρ2Vs2cosșSV1SV2) / (ρ1Vs1cosșSV1SV1))1/2

ENSV1P2 = (SV1P2)((ρ2Vp2cosșSV1P2) / (ρ1Vs1cosșSV1SV1))1/2 (2.22)

28
where ENSV1SV1 and ENSV1SV2 are the square root energy coefficients for reflected

and refracted SV-waves, and ENSV1P1 and ENSV1P2 are the square root energy

coefficients for reflected and refracted P-waves.

The incident SV-wave energy coefficients are defined by the following set of

equations:

ESV1SV1 = (ENSV1SV1)2

ESV1P1 = (ENSV1P1)2

ESV1SV2 = (ENSV1SV2)2

ESV1P2 = (ENSV1P2)2 (2.23)

where ESV1SV1 and ESV1SV2 are the energy coefficients for reflected and refracted SV-

waves, and ESV1P1 and ESV1P2 are the energy coefficients for reflected and refracted P-

waves.

29
CHAPTER 3

EXPERIMENTAL SEISMIC REFLECTION DATA ACQUISITION

3.1 Introduction

Seismic reflection data that were used in this dissertation were acquired in

cooperation with an Ohio Department of Transportation (ODOT) and Federal Highway

Administration (FWHA) engineering project. Reflection data were acquired in the

Interstate 70 (I-70) study area (Appendix A) during the fall of 1999 and during the

summer of 2001 (Figure 3.1). This chapter provides details regarding the recording

parameters and geometry of the seismic lines from which data are presented in this

dissertation, and a discussion regarding the pre-processing of the data.

3.2 Shear-Wave Reflection Survey - 1999

Shear wave seismic reflection profiles were acquired along a 200 ft section of the

eastbound lanes of I-70 during the fall of 1999. Line Test-1 was acquired parallel to and

60 ft south of the southern edge of the I-70 eastbound lanes, between road stations 48300

and 48400 (100 feet apart), and Line GUE-I70-1 was acquired between the 48300 and

48500 road stations (200 ft apart) along the southern edge of the eastbound lane of I-70

(Figure 3.2). (Note, road stations are in feet along the highway from the western
30
Figure 3.1. Photographs of 1999 (top) and 2001 (bottom) seismic reflection data
acquisition in the I-70 study area. See text for details regarding 1999 and 2001 data
acquisition geometry and recording parameters.

31
Figure 3.2. Map view of eastbound lanes of I-70 (road stations 48300 to 48500, 1999
seismic survey stations 100 to 300) showing the location of reflection lines Test-1 and
GUE-I70-1.

Guernsey county line. Highway engineers designate stationing as hundreds of feet plus

the remaining feet, for example, 483+00. In the text and the figures of this chapter these

numbers are combined, for example, 48300). Seismic survey stations 100 and 300

corresponded to road stations 48300 and 48500 respectively.

3.2.1 Data Acquisition

Detailed information regarding the acquisition and recording parameters for lines

Test-1 and GUE-I70-1 (acquired during 1999) is presented in Table 3.1. Data were

acquired using a shear wave vibratory source configured to generate preferential shear

particle motion transverse to the seismic lines (sometimes called crossline, or SH), and
32
Description Parameters
Spread type Inline CDP split-spread
Micro-Vib by Bay Geophysical Associates Inc., configured to
Energy source
generate preferential shear particle motion transverse to lines
Source interval 1 ft
2 ft north of lines (located on soil for line Test-1, and on I-70
Source offset
berm pavement for lines GUE-I70-1)
Line Test-1 = road stations 48397.5-48325.5 (73 source
Source locations locations), line GUE-I70-1 = road stations 48295.5-48497.5 (203
source locations)
Sweep type Linear (start taper = 0.08 sec, end taper = 0.06 sec)
Sweep frequencies 50-400 Hz
Sweep length 4 sec
Record length 0.75 sec (4.75 sec listen time minus 4 sec sweep)
Recording instrument OYO DAS-1, 24 bit A/D resolution
Recording channels 96 (with an additional 4 AUX channels)
Data format Recorded in SEG-D 8048 format, converted to SEG-Y format
Field filter and gain Low cut (3 Hz) to remove system noise, 48 dB constant scaling
Sample interval 0.25 ms
Geospace model SMC-70 (40 Hz), one single-component
Geophones horizontal element geophone oriented transverse to the line at
each station
Geophone interval 1 ft
Line Test-1 = road stations 48400-48305 (all geophones live for
Geophone locations each source location), line GUE-I70-1 = road stations 48301-
48492 (roll along began at source location 48349.5)

Table 3.1. Field acquisition and recording parameters for lines Test-1 and GUE-I70-1
acquired during the 1999 seismic reflection survey.

33
single component (horizontal element) 40 Hz geophones (Geospace model SMC-70)

oriented transverse to the seismic lines. Geophones with a natural frequency of 40 Hz

have been found to respond well for practical purposes when the dominant frequency of

field data is below 250 Hz (Steeples, 1998). Data were recorded using a 0.25 ms

sampling interval (resulting in a 2000 Hz Nyquist frequency) with a 96-channel (with 4

additional AUX channels) OYO Geospace DAS-1 seismograph. The change in surface

elevation across the profiles was less than one vertical foot per one hundred lateral feet.

3.2.1.1 Energy Source

The energy source employed was a non-commercial vibrator (approximate

dimensions: X and Z = 1 ft, Y = 2 ft; approximate weight = 300 pounds) that was

designed to preferentially generate shear waves. Depending on the orientation of the

source, preferential shear particle motion could be generated inline or transverse to the

seismic line. The vibrator consisted of 2 internal masses connected to an exterior box by

springs, and was capable of generating frequencies in the range of 1 Hz to 1 GHz. Linear

sweeps (upsweeps) were used to generate frequencies ranging from 50 to 400 Hz for each

record acquired. Start and end tapers (linear vibrator power control parameters) were

used to facilitate source coupling with the ground at the start and end of the sweep, in an

attempt to minimize recorded source-related noise. The source was configured to

generate preferential shear particle motion transverse to the azimuths of seismic lines

Test-1 and GUE-I70-1.

34
3.2.2 Data Pre-Processing

Reflection data were initially recorded in SEG-D format and subsequently

converted to SEG-Y format. Data pre-processing was conducted using the ProMAX

seismic processing package (Landmark Graphics Corporation) on a workstation in the

Department of Geological Sciences at the Ohio State University. Each SEG-Y field

record was read and output as a ProMAX dataset, and these datasets were then merged to

output a single ProMAX dataset for each line before processing and imaging. The

complete processing and imaging operation flows applied to lines Test-1 and GUE-I70-1

are discussed subsequently in this dissertation.

3.2.2.1 Vibroseis Correlation

Vibrator signals were recorded during data acquisition on four AUX channels:

the pilot sweep from the vibrator electronics (channel 0), source side plate acceleration

(channel -1), source mass number 1 acceleration (channel -2), and source mass number 2

acceleration (channel -3). For each record uncorrelated traces were crosscorrelated with

a sweep signal (either recorded in the field or synthetically generated) to produce

correlated records. Crosscorrelation produces zero-phase wavelets in correlated traces

when the sweep signal matches long wave trains recorded in uncorrelated traces.

A zero-phase wavelet consists of a central peak or trough (depending on display

convention) and two lower amplitude side lobes of opposite sign. A subsurface

impedance contrast (boundary) is located at the central peak or trough of a zero-phase

wavelet, and not at the wavelet onset (as is the case for a minimum-phase wavelet).

Because a zero-phase wavelet is centered on a subsurface boundary, interference with

35
wavelets from additional boundaries below and above a given boundary is possible

(Badley, 1985). For the minimum-phase wavelet case, interference with a wavelet from a

given boundary can only occur from boundaries below.

Periodic noise that was related to the source baseplate decoupling from the ground

was evident in line Test-1 and line GUE-I70-1 records. Noise related to source

decoupling is a common problem associated with vibratory sources (Seriff and Kim,

1970), with the amplitudes of such noise typically greater when the source is located on

relatively hard ground as opposed to softer ground. In addition to using the various

vibrator signals that were recorded in the field, correlation was performed using a

constructed synthetic linear sweep (50-400 Hz) in an attempt to minimize source-related

periodic noise in correlated gathers.

As shown in Figure 3.3a, periodic noise (relatively steeply dipping with a

dominant frequency of 65 Hz) is evident at near offsets in line Test-1 shot gathers when

correlation was performed using channel -2. Correlation using the synthetic sweep

resulted in a significant decrease of source-related noise however (Figure 3.3b), with no

apparent loss or degradation of signal (as evident from the hyperbolic reflection event

seen on shot gathers). Depths calculated using the apparent NMO velocities and zero

offset intercept times (approximately 115 ms) of the hyperbolic reflection event shown in

Figure 3.3, when correlated to drill log data from these locations, indicate that this

reflection event is from the overburden and bedrock boundary (i.e. the top of bedrock).

Correlation using a synthetic sweep did not result in decreased source-related

noise amplitudes for line GUE-I70-1 data, and correlated records for this line with the

highest signal to noise ratio were attained through correlation using AUX channel -2.

36
Figure 3.3. Line Test-1 shot gathers: (a) correlated using AUX channel –2, and (b)
correlated using a synthetic sweep. The steeply-dipping periodic noise (with dominant
frequency of 65 Hz) evident at near offsets in shot gathers correlated using AUX channel
–2 was suppressed when shot gathers were correlated using a synthetic sweep. Shot
gathers are unfiltered and have AGC scaling (100 ms window) applied for display
purposes. The X-axes scales of absolute offset from the source are in feet.

37
Differences in the amount (amplitudes) of source-related noise contained in each record,

and the ability to suppress the effect of such noise through correlation using a synthetic

sweep, were apparently related to differences in the magnitude of source decoupling that

occurred. The magnitude of source decoupling was related to the type of material that the

source baseplate was positioned on. Line Test-1 was acquired with the source located on

relatively soft soil, whereas line GUE-I70-1 was acquired with the source located on the

relatively hard roadway pavement (resulting in poorer source coupling with the ground).

3.2.2.2 Processing and Display Polarity

Polarity means the condition of being positive or negative (Sheriff, 1999), and

instead of describing reflections as being positive or negative the term polarity can be

used. Data shown in Figure 3.3 (and data from the 1999 survey subsequently presented

in this dissertation) were displayed using a reverse polarity convention. Reverse polarity

display means that a zero-phase wavelet without interference in data would have a central

trough (white) and two side lobe peaks (shaded black), and would correspond to an

increase in impedance with depth.

The polarity convention used for display of these data is opposite to that of

Society of Exploration Geophysicists (SEG) recommended standards (Sheriff, 1999),

which suggest that a zero-phase wavelet resulting from a boundary where impedance

increases across the boundary should be represented by a central peak (SEG normal

polarity). Despite the existence of the SEG recommendations, the meaning of the terms

normal and reverse polarity differ depending on geographical location, and polarity

conventions used for processing and display typically depend on the preference of the

38
processor or interpreter, and the study objectives (chosen in this study due to processing

and interpretation preference). What is most important is not the processing and display

polarity convention, but that an interpreter knows the polarity of the displayed data.

3.2.2.3 Geometry

Seismic lines Test-1 and GUE-I70-1 (Figure 3.2) were recorded using a 1 ft

geophone interval and inline and split spread (source locations in between geophone

locations) common depth point (CDP) shooting on a nearly horizontal surface (i.e. less

than 1 ft elevation change per 100 ft). Line Test-1 was acquired with single horizontal

element geophones planted between seismic survey stations 200 and 105 (corresponding

to road stations 48400 and 48305), and all geophones were live for each shot. The source

was initially located at survey station 197.5 for line Test-1, and was advanced and shaken

at 1 ft increments through the line (73 shot gathers with 96 traces in each) with a final

position at survey station 125.5. Line GUE-I70-1 was also acquired with single

horizontal element geophones, and the geophones were planted between survey stations

101 and 292 (corresponding to road stations 48301 and 48492), with roll along starting at

source location 149.5. The source was initially positioned at survey station 95.5 for Line

GUE-I70-1 and was advanced and shaken at 1 ft increments through the line (203 shot

gathers with 96 traces in each) with a final position at survey station 297.5.

CDP fold plots for lines Test-1 and GUE-I70-1 are presented in Figure 3.4. Fold

coverage plots shown are based upon the assumptions that subsurface reflectors are flat,

that no lateral velocity variations exist, and that the reflection point is at the geometric

midpoint between the source and receiver locations (i.e. geometric binning).

39
Figure 3.4. CDP fold relative to road stations for 1999 survey lines: (a) Test-1, and (b)
GUE-I70-1 (for both lines the source and geophone intervals were 1 ft, and source
locations were offset 2 ft north of geophones).

3.2.2.4 Data Truncation and Trace Editing

Preliminary analysis of shot records for both lines indicated that no signal of

interest would be lost by truncating the data from 750 ms to 300 ms. Therefore, data

truncation was performed in order to minimize the computational time required for data

processing and imaging. Noisy or bad traces contained within each record were visually

identified and manually killed, and the polarity of traces with a certain offset range for

shots 107 - 203 of Line GUE-I70-1 were reversed (these traces were recorded with

polarity opposite to that of the survey convention).

40
Figure 3.5. Map view of eastbound and westbound lanes of I-70 (road stations 48900 to
46694, 2001 seismic survey stations 1000 to 3206) showing the location of reflection
lines EBTravel, EBPass, WBPass, and WBTravel.

3.3 Multicomponent Reflection Survey - 2001


Multicomponent seismic reflection profiles were acquired along a 2206 ft section

of the eastbound lanes of I-70 during the summer of 2001 (between road stations 46694

and 48900). Seismic lines that are referred to as EBTravel and EBPass were acquired

along the south and north edges respectively of the eastbound lanes, and seismic lines

that are referred to as WBPass and WBTravel were acquired along the south and north

edges respectively of the westbound lanes (Figure 3.5). Seismic survey station 1000

corresponded to road station 48900, and seismic survey station 3206 corresponded to

road station 46694.

41
Description Parameters
Spread type Inline CDP split-spread
IVI “Minivib II” buggy, capable of generating preferential shear
Energy source particle motion inline and transverse to the line, and preferential
compressional particle motion
Line EBTravel: 3 configurations (3, 3-component (3-C) records
for each source station), lines EBPass and WBPass: transverse
Source configurations shear configuration (1, 3-C record for each source station), line
WBTravel: transverse and inline shear configurations (2, 3-C
records for each source station)
For line EBTravel inline and transverse shear configurations the
source interval was 1 ft for road stations 48900-47558, and 2 ft
Source interval for stations 47556-46694, for line EBTravel compressional
source configuration and line EBPass transverse shear
configuration the source interval was 2 ft for all stations
Source offset Source baseplate offset 6 ft from lines on soil (average offset)
Source locations Road stations 48900-46694
Sweep type Linear (start taper = 0.1 sec, end taper = 0.1 sec)
Sweep frequencies 50-500 Hz
Sweep, record lengths 4 sec, 1 sec (5 sec listen time minus 4 sec sweep)
Geometrics 48-channel StrataView and StataVisor modules
Recording instrument
connected in series, 24 bit A/D resolution
240 total, 80 3-component geophones deployed for each shot (3
Recording channels
channels used geophone, channel 1 of 240 used to record pilot)
Data format Recorded in SEG-2 format, converted to SEG-Y format
No field filters applied, pre-amplifier gain applied as a function
Field filter and gain of absolute offset (channels 0-30 = 0 dB, channels 30-32 = 24
dB, channels 32 - 240 = 36 dB, channel 240 = 48 dB)
Sample interval 0.25 ms
Geospace model GS-20DX (10 Hz), one 3-component
Geophones
geophone with orthogonal elements planted at each station
Geophone interval 2 ft
Road stations 48900-46694, when the source reached the last
Geophone locations
geophone the first 16 phones were leapfrogged to the line end

Table 3.2. Field acquisition and recording parameters for Lines EBTravel, EBPass,
WBPass, and WBTravel, acquired during the 2001 seismic reflection survey.

42
3.3.1 Data Acquisition

Detailed information regarding the 2001 seismic reflection survey acquisition and

recording parameters is presented in Table 3.2. Using a truck-mounted vibratory source

(capable of generating preferential shear or compressional particle motion), data were

acquired with a single 3-component (orthogonal elements) 10 Hz geophone (Geospace

model GS-20DX) planted at each station. Data were recorded using a 0.25 ms sampling

interval with 4, 48-channel Geometrics StrataView seismographs and one Geometrics

StrataVisor seismograph networked to allow simultaneous 240-channel recording. The

change in surface elevation across the profiles was less than one vertical foot per one

hundred lateral feet.

3.3.1.1 Energy Source and Configurations

The vibratory energy source employed was mounted on an Industrial Vehicles

International, Inc. (IVI) “Minivib II” buggy, and was capable (depending on vibrator

piston and baseplate orientation) of generating preferential shear particle motion in the

direction of the line (sometimes called inline, or SV) or transverse to the line (sometimes

called crossline, or SH), and preferential compressional particle motion (sometimes

called vertical, or P). Linear sweeps (upsweeps) were used to generate frequencies

ranging from 50 to 500 Hz for each record acquired. For line EBTravel the 3 possible

source configurations were used and 9-component (9-C) data were acquired using 3-

component (3-C) geophones. For both lines EBPass and WBPass the source was

configured to generate preferential shear particle motion transverse to the line, and 3-C

data were acquired using 3-C geophones. For line WBTravel two source configurations

43
were used (transverse and inline shear configurations), and 6-C data were recorded using

3-C geophones. In subsequent sections of this dissertation, components of a line will be

specified (when referring to data) in terms of orthogonal source and receiver components

in a Cartesian system. For example, the component of line EBPass recorded using a

source and geophone that were both orientated transverse to line, will be referred to as

line EBPassYY (YY refers to the source and geophone arrangements respectively in a

Cartesian system – this will be further explained in a subsequent section of this chapter

and also in subsequent chapters).

3.3.2 Data Pre-Processing

Reflection data were initially recorded in SEG-2 format and converted to SEG-Y

format. Correlated SEG-Y field records were received (from Nolen-Hoeksema, 2001),

and each source-receiver combination for each line was output as a single ProMAX

dataset before processing. The complete processing imaging operation flows applied to

reflection data that were acquired during the 2001 survey are discussed subsequently in

this dissertation.

3.3.2.1 Multicomponent Data Polarity Considerations

Proposed standards for the acquisition of multicomponent reflection data that are

consistent with previous SEG standards for data recording have been presented by

Stewart and Lawton (1999). These standards are based upon recommendations from the

SEG standards subcommittee on multicomponent data (Pruett, 1987), and were proposed

with the petroleum industry in mind, but are also applicable for reflection surveys with

44
near-surface objectives. When 3-C geophones with orthogonal elements (one vertical,

and two horizontal) arranged in a Cartesian system are employed, a right-handed

coordinate system (with X, Y, and Z axes), with the Z-axis pointing downwards is

recommended. The Z-axis corresponds to a vertically oriented geophone element, while

the X- and Y-axes correspond to horizontal geophone elements, oriented inline and

transverse to the seismic line respectively.

Standards for the acquisition of multicomponent reflection data were proposed in

order to facilitate the processing and display of data with normal polarity, as defined by

Sheriff (1999), and to provide polarity consistency between displayed components. The

proposed recommendations state that a tap on the top of a geophone case, and taps on the

sides of the geophone case in the positive X-axis direction (in the direction of source

advancement), and positive Y-axis direction (with a positive rotation angle of +90

degrees from the positive X-axis direction) should all yield consistent initial (positive)

amplitudes (Figure 3.6a). When multicomponent data are recorded following these

conventions, the processing and display of common-mode P-wave and converted-mode

(P-SV-wave) sections with the same polarity (for “normal” geologic environments) is

facilitated (Garotta, 2000). A “normal” geologic environment as stated in the previous

sentence, refers to a subsurface interface across which compressional and shear

impedances both change in the same direction.

In the above situation, a P-wave with near-normal incidence on the interface will

typically result in P- and SV-wave (mode-converted) reflected displacement amplitude

coefficients (real part) with opposite sign at positive source to receiver offsets (see

Chapter 2, Appendix B). With the SEG recommended recording polarities (Figure 3.6a),

45
Figure 3.6. Multicomponent recording polarities for 3-C geophones with elements
arranged in a Cartesian system: (a) recording convention consistent with SEG
recommendations, and (b) 2001 multicomponent survey recording convention. Raw data
were recorded during the 2001 survey with receiver polarities configured such that a tap
on the top of a geophone case yielded an initial positive amplitude, and taps on the sides
of a geophone case in the positive X and Y axis directions yielded initial negative
amplitudes. Source advancement direction is in the positive X-axis direction, and the
positive Y-axis direction has a rotation angle of +90 degrees from the positive X-axis.

both of these upward traveling wave types would produce negative initial geophone

outputs (amplitudes) at positive offsets. The polarity of traces recorded from a P-wave

source at negative source-to-receiver offsets on horizontal, inline geophone elements was

reversed during data pre-processing (Figure 3.7), and this is an established petroleum

industry multicomponent data analysis procedure (Garotta, 2000).

For the 2001 multicomponent seismic reflection survey conducted in the I-70

study area, data were acquired for lines EBTravel, EBPass, WBPass, and WBTravel

using 3-C geophones, with elements arranged in a Cartesian system. A tap on the case

top of the geophone model that was used yielded an initial positive amplitude response,

46
Figure 3.7. Shot gather (P-wave source, Z component) recorded using horizontal, inline
(X component) geophone elements: (left) before, and (right) after polarity reversal of
trailing spread. Absolute values of offset from the source on the x-axis are in ft.

however, taps on the sides of a geophone case in the positive X-axis (west) and Y-axis

(north) directions yielded initial negative amplitudes (Figure 3.6b), which differ from the

proposed multicomponent polarity acquisition standards. In order to achieve polarity

consistency between components that were recorded during the 2001 survey, the polarity

of the Z-axis data components (corresponding to vertical geophone elements) were

reversed during data pre-processing. A reverse polarity convention (according to SEG

recommendations) is used for display of the 2001 data presented in this dissertation.

3.3.2.2 Geometry

Seismic lines EBTravel, EBPass, WBPass, and WBTravel were recorded using a

2 ft geophone interval and inline and split spread CDP shooting on a nearly horizontal

47
surface (i.e. less than 1 ft elevation change per 100 ft). For each of the profiles 80, 3-

component geophones (with orthogonal elements) were deployed at one time along the

line (3 recording channels per geophone, with channel 1 of 240 used to record the pilot

sweep). When the source reached the last geophone, the first 16 geophones were

leapfrogged to the end of the line and source advancement continued down the line.

For line EBTravel shear-wave source configurations (inline and transverse), the

source was initially located at road station 48900. The source was advanced and shaken

at 1 ft increments through road station 47558, and at 2 ft increments in between stations

47556 to 46694 (Figure 3.5). For the line EBTravel vertical source configuration, the

lines EBPass and WBPass transverse shear configuration, and the WBTravel transverse

and inline shear configurations, the source was initially positioned at road station 48900,

and was advanced and shaken at 2 ft increments through the lines with a final position at

road station 46694 (Figure 3.5). CDP fold plots for lines EBTravel, EBPass, WBPass,

and WBTravel are presented in Figure 3.8 (fold coverage plots assume geometric

binning, see above for explanation of the term geometric binning).

3.3.2.3 Data Truncation and Trace Editing

After preliminary analysis of the data, truncation (from 1.0 sec to 300 ms) was

performed in order to minimize the computational time required for data processing and

imaging. Traces recorded using channels 193-240 for low numbered shots (when the

source was positioned at survey stations 1000-1158) on line EBTravel (inline and

transverse shear wave source records) were eliminated from the data because of data

acquisition system problems. This range of traces had improper channel numbers written

48
to the trace headers, events across these traces had inaccurate moveout, and traces had

inconsistent polarity. Despite this range of traces being eliminated, the fold was still high

along this part of the survey line relative to the fold obtained during typical shallow

reflection surveys.

Due to the large number of shot records acquired during the 2001 survey,

statistical processes were used to identify and isolate shots containing noisy or bad traces

in order to efficiently edit traces. A time gate for statistical processes was selected, and

statistical attribute values of average trace energy and trace spikiness (ratio of maximum

magnitude sample to trace signal amplitude) were calculated for the data. In order to

prevent large amplitude spikes from obscuring the rest of the data, a logarithmic function

was applied to the average trace energy attribute. Histograms were plotted for the log of

the average trace energy and trace spikiness attributes. Ranges of anomalously high

attributes on the histograms were selected, and these selections were projected to the shot

domain in order to isolate shot gathers that contained traces with anomalously high

amplitudes or spikiness. Bad traces for each of the isolated shots were then visually

identified and manually eliminated.

49
Figure 3.8. CDP fold relative to road stations for 2001 survey: (a) line EBTravel inline
and transverse shear source configurations (source interval was 1 ft for stations 48900-
47558, and 2 feet for stations 47556-46694, geophone interval was 2 ft), and (b) line
EBTravel compressional source configuration, lines EBPass and WBPass transverse
shear source configuration, and line WBTravel transverse and inline shear source
configurations (source and geophone intervals were 2 ft for all stations).

50
CHAPTER 4

FACTORS AFFECTING COMMON-MODE


REFLECTION SURVEYING EFFECTIVENESS

4.1 Overview

Research was conducted to compare the effectiveness of near-surface

compressional-wave (P-wave) and shear-wave (S-wave) reflection surveys for mapping

geologic and manmade features in the shallow subsurface. P-wave reflection data have

traditionally been acquired during shallow reflection surveys, but the number of reports

concerning shallow S-wave reflection surveys is relatively small (see Chapter 1 for an

overview of previous work). Very few reports concerning the concurrent acquisition and

analysis of P- and S-wave reflection data exist. Objectives of the research were

addressed experimentally by acquiring P- and S-wave reflection data in the I-70 field test

area (Appendix A) and through analyses and modeling.

S-wave reflections from the bedrock and overburden interface were consistently

measured in both the XX component (inline-inline, or SV-SV) and the YY component

(crossline-crossline, or SH-SH) field data. However, noise from surface waves resulted

in the optimum reflection window of XX component data being relatively narrow. Stacks

of traces that were produced using YY component data had a higher signal-to-noise ratio

and imaged the top-of-bedrock better than stacks produced using XX component data.

51
P-wave reflections from the unsaturated and saturated overburden interface were

recorded in ZZ component (vertical-vertical, or P-P) field data, due to a large P-wave

velocity increase across this interface. S-wave reflections from this interface however

were not observed in acquired data components. Arrival times of P-wave reflections and

characteristics of recorded noise modes made it challenging to process and use P-wave

reflections from the top of the saturated overburden. P-wave events from deeper

contrasts in impedance could not be resolved in field data due to the following: 1) surface

wave and air wave noise, 2) a high P-wave reflection coefficient at the top of the

saturated overburden, 3) low P-wave reflection coefficients at deeper interfaces, and 4)

interference effects and poor resolution.

Calculations that were based on P- and S-wave reflection velocities (determined

from NMO) and dominant wavelengths suggest that the vertical resolution of S-waves in

the study area dry overburden was more than 1.7 times the resolution of P-waves, while

the resolution of S-waves in the saturated overburden for the study area was more than

4.7 times that of P-waves. The potential of producing a map showing detailed variations

in elastic properties of subsurface media using acquired P- and S-wave reflection

information was limited by the small number of reflection events from acquired data, and

since the observed P- and S-wave events do not correlate to similar subsurface interfaces.

Comparing the known subsurface conditions with processed stacked sections

indicates that the combined P- and S-wave common-mode reflection information allowed

the near-surface geologic sequence to be imaged more effectively than using solely the S-

wave or P-wave information. Areas of the subsurface where geologic discontinuities

were present were most accurately delineated through the processing and interpretation of

52
YY component data. The mine-subsidence project-related objectives in the study area

(Appendix A) would not have been met using the traditionally acquired P-P component

data. Findings of this study demonstrate the necessity of considering the potential

usefulness of different wave-types and data components prior to conducting a near-

surface reflection survey.

4.2 Multicomponent Reflection Data

Research objectives of this study were addressed through analyses of

experimental multicomponent reflection data acquired during 2001 along a section of the

eastbound travel lane of I-70 (Figure 4.1). The reflection line acquired along this lane

during 2001 is referred to as line EBTravel. Nomenclature in this chapter that is used to

describe multicomponent data, in terms of source and receiver orientations and

preferential polarizations, is illustrated in Figure 4.2. The three seismic source

orientations that were used generated preferential shear particle motion inline and

transverse (crossline) to the seismic line (source components X and Y respectively), and

preferential compressional particle motion (source component Z). A single 3-component

geophone was placed at each receiver station. The geophones contained: 1) horizontal

(orthogonal) elements oriented inline and transverse to the line (receiver components X

and Y respectively), and 2) a vertical element (receiver component Z). This chapter

primarily focuses on the usefulness (in terms of reflection information) of common-mode

component data that were acquired for line EBTravel: the XX component (sometimes

called inline-inline, or SV-SV), the YY component (sometimes called crossline-crossline,

53
54
Figure 4.1. Map view of the eastbound lanes of I-70 (road stations 48500 to 48900) showing the location of the seismic
reflection line EBTravel relative to the locations of the underground mine workings. Due to the small map scale and possible
errors that exist regarding room and pillar locations on the map, the spatial relationship between these features and the I-70 road
stations is regarded as only approximately accurate.
Figure 4.2. Nomenclature used to describe multicomponent reflection data, in terms of
the source and receiver orientations and preferential polarizations. For sources, the X and
Y symbols indicate sources configured to preferentially generate shear particle motion
inline and transverse (crossline) to the seismic line respectively, while Z indicates a
source configured to preferentially generate compressional particle motion. For
receivers, the symbols X and Y indicate horizontal geophone elements oriented inline and
transverse to the line respectively, while Z indicates a vertical geophone element. The
source and receiver pairs: XX, YY, and ZZ are referred to as the common-mode
components of the nine-component matrix (e.g. XX means a source and receiver both
oriented in the X direction).

55
or SH-SH), and the ZZ component (sometimes called vertical-vertical, or P-P). Line

EBTravel is often referred to in the text of this chapter, and on figures as either: line

EBTravelXX, line EBTravelYY, or line EBTravelZZ, in order to indicate the specific

source and receiver type of the data that are shown. Detailed information regarding the

2001 reflection survey acquisition parameters, and discussion regarding data geometry

and pre-processing were presented in Chapter 3.

4.3 XX Data Versus YY Data: Factors Affecting Imaging Potential

Shot gathers (i.e. field records showing all of the traces acquired for a single

source location) for the XX and the YY components acquired at four locations along line

EBTravel are shown in Figure 4.3. The source locations for the gathers correspond to

road stations: 48581, 48638, 48727, and 48789. (Note, road stations are in feet along the

highway from the western Guernsey county line. Highway engineers designate stationing

as hundreds of feet plus the remaining feet, for example, 485+81. In the text and the

figures of this chapter these numbers are combined, for example, 48581.) The gathers are

shown as unprocessed data (top), as data with bandpass filter and AGC applied (middle),

and as interpreted data (bottom). S-wave reflections are indicated on the interpreted data

records at zero-offset arrival times of about 0.11 seconds (110 ms). The character of

reflection events relative to reflection events observed to the east and west, and the depth

estimates using velocities derived from reflections when correlated with the available

drill log data (Table 4.1), indicate that the observed reflections (indicated on Figure 4.3)

are caused by the overburden and bedrock boundary. The average dominant frequencies

56
57
Figure 4.3. Line EBTravel shot gathers: (a) XX component, and (b) YY component. Gathers are shown: (top) unprocessed,
(middle) with a bandpass filter (50-80-160-200 Hz) and AGC (100 ms window) applied, and (bottom) interpreted. S-wave
reflections from the top-of-bedrock are indicated, and apparent NMO velocities and approximate depths are given. The x-axis
scales of absolute offset (AOFFSET) from the source locations (SOU_X; source locations are given in feet from the western
county line, see Figure 4.1) are in feet. See text for discussion.
Distance (in feet) Material
Boring log Drill date Road Depth (in feet) to water Depth (in feet)
north (N) or south encountered beneath
number (month/yr) Station during drilling to bedrock
(S) of seismic line bedrock

GC-208 11/99 48419 41’ N of EBTravel 25-27, 31-43 42 Coal

GC-218 11/99 48421 6’ S of EBTravel 22-28, 36-42 41 Not recorded

GC-209 11/99 48459 39’ N of EBTravel Not recorded 43 Not recorded

P-221A 11/99 48500 6’ S of EBTravel 31-42 42 Coal

B-38 * 12/94 48525 5’ S of EBTravel 27 (24 hour level) Not available Coal

58
B-122 ** 06/95 48570 24’ N of EBTravel 27 42 Grout

B-123 *** 06/96 48600 10’ N of EBTravel Not recorded 41 Grout

* No sampling conducted (from 36-53 feet deep) in the top-of-bedrock depth range.
** Log indicates a “broken zone” (42-44 feet deep) of poorly graded sand with fine gravel infilling between cobbles.
*** Log indicates shale unit is heavily jointed with slickensides, and that grout below bedrock is broken and jointed.

Table 4.1. Information from drill logs near line EBTravel (Figure 4.1).
59
Figure 4.4. Processed (bandpass filter and AGC gain applied) line EBTravel multicomponent shot gathers: source components
X (a), and Y (b). The S-wave reflections interpreted in Figure 4.3 are superimposed on the common-mode component (XX and
YY) gathers. The x-axis scales of absolute offset (AOFFSET) from the source locations (SOU_X; source locations are given in
feet from the western county line, see Figure 4.1) are in feet. See text for discussion.
of the reflections in the field records that were acquired using X- and Y-oriented sources

had similar values.

Processed multicomponent gathers (source and receiver component combinations:

XX, XY, XZ, YX, YY, YZ) with the same source locations as the gathers in Figure 4.3,

are shown in Figures 4.4a (source = X) and 4.4b (source = Y). S-wave reflections from

the top-of-bedrock are indicated on the gathers in Figures 4.3 and 4.4. Gathers in Figures

4.3 and 4.4 have individual trace amplitude scaling applied to enhance the later time

events.

4.3.1 Noise, Optimal Reflection Windows, and Reflection Coefficients

In order to produce an accurate image of the subsurface using seismic reflection

data, reflection energy (referred to as signal) must first be confidently identified.

Random noise and different types of coherent noise (i.e. types non-reflection energy)

recorded by geophones can destructively interfere with signal of interest, and this can

make accurate seismic imaging difficult or even impossible. A main goal of seismic

reflection data processing is to enhance identifiable signal by suppressing noise, and to

generate images of the subsurface containing minimal noise-related artifacts.

There are several types of coherent noise that can complicate or prevent shallow

seismic reflection imaging efforts. A blast of air wave noise is typically generated by

seismic sources, and coherent noise can also be generated by other environmental

sources, such as roadway traffic. Surface waves propagate along or near the ground

surface, and often serve as noise that is detrimental to reflection energy. Two types of

surface waves are commonly a concern during near-surface seismic reflection surveying:

60
Rayleigh waves and Love waves (Sheriff and Geldart, 1982). Rayleigh waves propagate

with a retrograde elliptical particle motion confined to the vertical plane in the direction

of the seismic source, and involve a combination of both compressional and shear waves

(P- and SV-waves). Love waves have a particle motion that is parallel to the surface and

perpendicular to the direction of propagation (SH), with no vertical component of motion.

Rayleigh waves will propagate along a surface regardless of the near-surface velocity

structure. Love waves however, require a near-surface layer of relatively low velocity, as

they result from the interference of multiples of reflected SH-wave energy beyond the

critical angle that are confined to a near surface layer.

Traffic noise on the I-70 study area field records was higher for data recorded

using X- and Z-oriented receivers, than for data recorded using Y-oriented receivers.

Recorded traffic noise was predominantly low frequency (e.g. 5-25 Hz), and was

therefore significantly suppressed on all components through the application of a low-cut

frequency filter. Air wave noise was present in all data that were acquired using X- and

Z-oriented receivers, and could not be suppressed without degrading the reflection signal

quality. Air wave noise propagated along the receiver spread before the return of S-wave

reflections from the bedrock. Processing procedures to attenuate air wave noise were not

necessary to prevent it from severely degrading the S-wave reflection signal in the XX

component data.

Detrimental, high amplitude, dispersive Love wave noise was not observed in

most of the YY component data. This is because the S-wave velocity structure across the

study area was such that near-surface road construction-related fill materials had a higher

velocity than underlying overburden materials at most locations. For YY component data

61
at far source-to-receiver offsets where the top-of-bedrock S-wave event was degraded by

non-reflected energy, the noise was suppressed through f-k filtering and post-NMO

stretch muting.

High amplitude Rayleigh wave noise was observed in the data recorded using X-

and Z-oriented receivers (Figure 4.4). Rayleigh waves propagate in the direction of the

seismic source with retrograde elliptical particle motion in the vertical plane. Rayleigh

waves therefore strongly affect both the X and Z receiver components. The velocity of

Rayleigh waves (Vr) is dependent upon Poisson’s ratio ( σ ), and is close to that of the S-

wave velocity (Vs) with the relationship given as:

Vr = ĮVs (4.1)

where Į is a constant that is dependent upon that value of σ . An equation relating Į and

σ can be found in Davis and Selvadurai (1996). For example σ values of: 0.0, 0.25, and

0.5, Vr would be equivalent to: 87.4 percent of Vs, 92.0 percent of Vs, and 95.5 percent of

Vs respectively. Measurements made using acquired field data indicate a representative

range for values of Vr (linear group velocity) in the study area of 575 ft/s to 650 ft/s.

Rayleigh wave noise tended to mask the top-of-bedrock S-wave reflection that

was recorded in XX component data. The Rayleigh wave noise could not be sufficiently

suppressed through frequency filtering, resulting in the optimum reflection window being

narrower than the window for the YY component data (Figure 4.3).

S-wave energy reflected from the top of the bedrock (from a X-oriented source) is

strongly polarized in the horizontal component (X-oriented receiver component) in the

direction of the line (Figure 4.4a). This polarization effect is shown by high amplitude

reflections in the X-oriented receiver gathers (top of Figure 4.4a), relative to those (when

62
even observed) in Y- and Z-oriented receiver gathers (middle and bottom of Figure 4.4a

respectively). The S-wave energy that was reflected from the top of the bedrock (from a

Y-oriented source) is strongly polarized in the horizontal component in the crossline

direction as shown in Figure 4.4b. Reflections have high amplitudes in Y-oriented

receiver gathers (middle of Figure 4.4b), relative to those (when even observed) in X- and

Z-oriented receiver gathers (top and bottom of Figure 4.4b respectively). The amplitude

and polarization characteristics of the recorded noise modes, and the amplitude and

polarization characteristics of reflected energy, determines whether or not a reflection can

be observed in a given data component. In practice, acquisition geometry is never

perfect, media are not perfectly isotropic and laterally homogeneous, and reflections may

be affected by dipping interfaces. Further, seismic sources and receivers are not perfectly

pure in a polarization sense (i.e. more than one type of wave is typically generated by

sources, and receivers typically are not completely sensitive to only a single wave type).

Miller and Pursey (1955) calculated that a vertically oscillating disk on a half-space

medium (having a Poisson’s ratio of 0.25) radiates 67 percent of the total energy as

Rayleigh waves, 26 percent as SV-waves, and 7 percent as P-waves. It is unlikely when

using a Z-oriented source for example, that the X-oriented component would not contain

any reflected energy.

YY component field records generally contained a higher signal-to-noise ratio,

and a larger optimum reflection window than corresponding XX component records

(Figure 4.3). Therefore, accurate estimates of the average overburden S-wave velocity

can generally be better obtained from the YY component data. The signal-to-noise ratio

63
and the size of the optimum reflection window for the XX and YY component data also

affected the quality of images that could be constructed using these components.

Square root energy coefficients were calculated using the equations presented in

Chapter 2 for the overburden and bedrock interface. Representative P- and S-wave

velocities were calculated from the seismic data for the saturated overburden (directly

above bedrock) of 5150 ft/s and 700 ft/s respectively. An S-wave velocity of 2750 ft/s

for the bedrock, and an approximate P-wave velocity of 5500 ft/s (assuming a Vp/Vs ratio

of 2.0), were also used for calculations. The P-wave velocity of the bedrock could not be

measured directly from the field data. A density contrast of 0.63 g/cm3 was estimated for

the region across the interface between saturated overburden and bedrock for

calculations. This approximation was based on an average density value for wet

overburden materials of 1.92 g/cm3 (Telford et al., 1990), and an average bulk density of

2.55 g/cm3 measured for the study area bedrock unit in Harrison County, Ohio (ODNR,

2002). Results from calculations of the square root energy coefficients using these

velocity and density estimates are plotted in Figure 4.5 as a function of incidence angle.

These calculations are based on assumptions that the interface is planar and that the

media are isotropic.

The plots in Figure 4.5 show that the magnitude of the S-wave reflection

coefficient is high at normal incidence (0.68). The magnitudes of the SV-wave (left) and

the SH-wave (right) reflection coefficients are also seen from a qualitative standpoint to

be fairly similar to each other for most incidence angles. The maximum incidence angle

that is shown roughly corresponds to the likely maximum source-to-receiver offset at

which a top-of-bedrock S-wave reflection would be recorded in acquired field data.

64
65
Figure 4.5. Plots showing normalized square root energy coefficients as a function of incidence angle for a SV-wave (left) and a
SH-wave (right) incident on the overburden (medium 1) and bedrock (medium 2) interface. Calculations were made with the
computer program presented in Appendix B. See text for discussion.
Energies of reflected and refracted mode-converted waves from an incident SV-wave are

relatively low for this geologic situation, and such modes only occur at small (< 8

degrees) incidence angles. The reflected mode-converted P-wave and refracted mode-

converted P-wave critical angles are indicated in Figure 4.5. Based on the plots in Figure

4.5 and analyses of the data presented in this section, it can be concluded that the possible

effects of reflection coefficients on the S-wave imaging potential are likely to be minor

relative to the affects of noise recorded in the S-wave data components.

4.3.2 Stacked Signal Quality

An S-wave velocity field can be more accurately constructed through analysis of

YY component data than from XX component data. In most cases, the average

overburden S-wave velocities measured from both components were similar, when the

velocity was measured at a location using both components. Small differences in the

average S-wave overburden velocities (< 3 percent) were measured from corresponding

XX and YY data in a relatively small number of cases (e.g., see the XX and YY

component shot gathers shown in Figure 4.3 that were acquired with the sources located

at road station 48581).

XX component and YY component supergathers (each consisting of 3 smashed

CMP gathers) from three locations along line EBTravel are shown in Figure 4.6. The

gathers are each shown before and after NMO-correction using S-wave stacking

velocities calculated from YY component data (Table 4.2). For the XX component

supergather centered at road station 48820 (Figure 4.6a), the event from the top-of-

bedrock is slightly overcorrected (applied velocity was too low) using the YY

66
Figure 4.6. Line EBTravel CDP supergathers: XX component (left), and YY component
(right). Gathers are shown before and after NMO correction using YY component-
derived (Table 4.2) S-wave stacking velocities of: (a) 737 ft/s, (b) 705 ft/s, and (c) 675
ft/s. Arrows next to dynamic stack functions indicate the top-of-bedrock reflection event.
This reflection is slightly over-corrected on the 48820 (a) XX component gather
(although data still stack reasonably well at the applied velocity). For supergathers
centered at the other two locations (b and c), this event is corrected similar on both the
XX and YY components. CDP locations (CDP_X) correspond to road stations (given in
feet from the western county line, see Figure 4.1).
67
component-derived velocity. However, traces observed on the dynamic stack function

(plotted to the right of the NMO-corrected XX component gather in Figure 4.6a) indicate

the data still stacked reasonably well at the applied velocity. The top-of-bedrock event is

corrected and stacks give similar results for both components using supergathers that are

centered at the other two locations (Figures 4.6b and 4.6c). Analyses conducted using

XX and YY data, suggested that comparable stacks for the two components could be

obtained using a velocity field derived from YY component data. It was apparent from

these analyses that other factors previously discussed in this section (e.g., optimum

reflection windows and signal-to-noise ratios of acquired data) would have a greater

impact than the small errors in the applied stacking velocities (i.e., < 3 percent) on the

quality of constructed XX and YY component images.

The line EBTravel XX and YY component stacked time sections shown in Figure

4.7 were produced with similar parameters (shown in Table 4.2), and the NMO

corrections were applied using the same (YY component-derived) velocity field. Stacked

XX and YY component data from the line are shown with NMO corrections applied at

100 percent of the stacking velocities in Figure 4.7a, at 95 percent in Figure 4.7b, and at

105 percent in Figure 4.7c. At each velocity-field percentage, the YY component data

have a higher top-of-bedrock reflection signal-to-noise ratio than the XX component data,

and the YY component data show better resolution of the bedrock horizon than the XX

component data. This is observed in an area where no previous mining activity occurred

(between road stations 48880 and 48660), and it is reasonable to expect that a

predominately continuous and coherent bedrock horizon should be present across this

area. The amplitude of the bedrock reflection is high and the event is coherent across

68
Processing step Description
Data reformat From SEG-Y to ProMAX format
Vibroseis correlation Both lines correlated with pilot sweep
Geometry Defined using field notes and loaded to headers
Data truncation Records truncated to 300 ms
Trace editing Bad / noisy traces killed
Trace equalization 150 ms spatially varying window
f-k filter Non-reflection energy/linear noise suppression
CMP sort Sorted from shot gathers to midpoint gathers
Stacking velocity function derived from YY component
Velocity analysis through integrated analysis of shot gathers, constant velocity
stacks, and semblance plots
NMO correction Applied based on optimum stacking velocities
Stretch mute 30 percent
Bandpass filter Zero-phase Ormsby filter: 50-80-160-200 Hz
AGC scaling 100 ms window
CMP ensemble / stack Summed NMO-corrected CMP gathers

Table 4.2. Data processing flow for Line EBTravel S-wave reflection data (XX and YY
and components), acquired east of the previous roadway collapse region.

69
Figure 4.7. Line EBTravel XX and YY component time sections after similar processing
(Table 4.2), and NMO corrections made using the same velocity model: (a) 100 percent of
stacking velocities, (b) 95 percent, and (c) 105 percent. CDP locations (CDP_X) correspond
to road stations (given in feet from the western county line, see Figure 4.1). Y-axis arrows
indicate the top of the bedrock horizon. The YY component sections have a higher signal-to-
noise ratio, and provide a better stack of the bedrock horizon. No previous mining occurred
below line EBTravel between CDP locations 48880 and 48660. (continued)
70
Figure 4.7. (continued)

(b) Line EBTravel XX and YY component sections after similar processing. (continued)

71
Figure 4.7. (continued)

(c) Line EBTravel XX and YY component sections after similar processing.

72
this station range in the YY component stacks relative to the XX component stacks

(Figure 4.7). The YY component data provide a better potential than XX component data

to identify disruptions in the bedrock horizon.

4.4 YY Data Versus ZZ Data: Factors Affecting Imaging Potential

Line EBTravel ZZ component shot gathers with source locations corresponding to

road stations: 48580, 48638, 48726, and 48788 are shown in Figure 4.8a. The shot

gathers are shown as unprocessed (top), with bandpass filter and AGC applied (middle),

and with interpretations (bottom). The YY component shot gathers (unprocessed,

processed, and interpreted) shown in Figure 4.3 are shown again in Figure 4.8b, for the

purpose of comparison with ZZ component gathers. The XX and YY component data for

line EBTravel were recorded approximately 2 weeks prior to the ZZ component data

(Chapter 3). A different source spacing was used for the XX and YY data compared to

the ZZ data.

Multicomponent shot gathers (components: ZX, ZY, and ZZ) with the same

source locations as the gathers shown in Figure 4.8a, are presented in Figures 4.9a

(unprocessed) and 4.9b (processed). The P-wave reflections that were indicated on ZZ

component records on Figure 4.8a are again indicated on Figure 4.9b. The frequency

spectra of multicomponent records acquired using a Z-oriented source were similar in all

cases. A bandpass frequency filter is often effective in improving the visibility of near-

surface P-wave reflections (Steeples et al., 1997). The filter that was applied to these

data was an 80-120-200-240 Hz zero-phase Ormsby filter. The shot gathers for these

data are shown in Figures 4.8 and 4.9, with individual trace amplitude scaling applied.
73
74
Figure 4.8. Line EBTravel shot gathers: (a) ZZ component, and (b) YY component. Gathers are shown: (top) unprocessed,
(middle) with a bandpass filter (ZZ: 80-120-200-240 Hz, YY: 50-80-160-200 Hz) and AGC (100 ms window) applied, and
(bottom) interpreted. P-wave reflections from the top-of-saturated-overburden (a) are indicated in blue, and S-wave reflections
from the top-of-bedrock (b) are indicated in yellow, with apparent NMO velocities and approximate depths given. The x-axis
scales of absolute offset (AOFFSET) from the source locations (SOU_X; source locations are given in feet from the western
county line, see Figure 4.1) are in feet. See text for discussion.
75
Figure 4.9. Line EBTravel multicomponent (source component Z, receiver components X, Y, and Z) shot gathers: (a)
unprocessed, and (b) processed (bandpass filter: 80-120-200-240 Hz, and AGC gain with 100 ms window applied). P-wave
reflections interpreted in Figure 4.8 are superimposed on the processed common-mode component (ZZ) gathers. The x-axis
scales of absolute offset (AOFFSET) from the source locations (SOU_X; source locations are given in feet from the western
county line, see Figure 4.1) are in feet. See text for discussion.
4.4.1 Sensitivity of P- and S-Waves to Changes in Overburden Saturation

Reflected P-wave energy is indicated on the interpreted records in Figure 4.8a at

zero-offset arrival times of about 30 ms. Depth estimates using velocity values that were

derived from reflections when correlated with available drill log information (Table 4.1)

indicate that the observed reflections are from the top-of-saturated-overburden (where dry

or partially saturated unconsolidated overburden materials become fully saturated).

Depth estimates from the events interpreted in Figure 4.8a, agree with Hoffman et al.

(1995), where it was reported that the groundwater within the overburden was generally

within 30 feet (9.1 m) of the ground surface. Depth estimates also agree with acquired

hydrologic well data, which indicated that water levels in overburden materials ranged

from within 20 - 30 feet (6.1 - 9.1 m) of the ground surface along the eastbound lanes

during 1999-2001. No major lithologic boundaries were encountered during drilling in

this depth range.

At certain locations along line EBTravel, high amplitude P-wave reflections and

refractions from the top-of-saturated overburden were clearly observed, and were well

separated in the frequency domain from recorded noise modes (Figure 4.10). However,

at most locations along the line, the arrival times of events and the noise on the data made

it challenging to process the data to show the P-wave reflections from the top-of-

saturated-overburden. P-wave reflections from geologic interfaces below the top-of-

saturated-overburden, were not observed in any of the acquired records (this is

subsequently explained).

Reflection and refraction analyses of data acquired in the study area indicated

average P-wave velocities for unsaturated overburden around 1900 ft/s, and P-wave

76
Figure 4.10. Line EBTravel ZZ component CMP gathers: (a) uninterpreted, and (b)
interpreted. Gathers are shown with a bandpass filter (80-120-200-240 Hz) and AGC
(100 ms window) applied. P-wave reflections from the top-of-saturated-overburden are
indicated in (b), with apparent NMO velocities around 1900 ft/sec, and reflector depths of
about 32 feet. X-axis scales of CMP location (CDP_X) and receiver location (REC_X)
are in feet, and correspond to road stations. See text for discussion.

velocities around 5150 ft/s for the saturated overburden. Such large contrasts in

unconsolidated material P-wave velocities due to water saturation have commonly been

observed during near-surface reflection surveys (Miller and Xia, 1998). P-wave

velocities of fully saturated unconsolidated materials are almost always equal to or

greater than (except in cases of very high porosity, i.e., greater than or approximately

equal to 65 percent) the velocity of water (Bradford, 2002). Acquired S-wave data

(Figure 4.8b) contained S-wave reflections from the top-of-bedrock, but did not contain

events correlating to top of saturated overburden (i.e., changes in overburden saturation

77
could not be inferred using S-waves). This is because P- and S-waves are sensitive to

different physical material properties.

P-wave velocities (and amplitudes) can vary substantially with changes in pore

fluid content (Domenico, 1976), primarily because of the sensitivity of bulk modulus (k)

values to such changes. P-wave velocity (Vp), is dependent upon k, the shear modulus

(µ), and bulk density ( ρ ):

Vp = (k + 4/3µ / ρ )1/2 (4.2)

S-wave velocity (Vs) depends upon µ and ρ :

Vs = (µ / ρ )1/2 (4.3)

The rigidity of ideal gases and liquids is the same (zero). Despite possible changes in Vs

related to changes in density or cohesion (West and Menke, 2000), the saturation of an

unconsolidated material with water generally does not change the S-wave velocity

appreciably relative to the change in P-wave velocity.

4.4.1.1 Reflection Coefficients: Unsaturated and Saturated Overburden Interface

Square root energy coefficients (see Chapter 2 for equations) were calculated for

P- and S-waves incident on an interface between unsaturated and saturated overburden

using data-derived P- and S-wave velocities, and approximate bulk densities. Results are

plotted in Figure 4.11 as a function of incidence angle, and are based on assumptions that

the interface is planar, and that the media are isotropic. Representative study area media

parameters used for the square root energy calculations are presented in Table 4.3. P-

wave velocities for the unsaturated and saturated overburden of 1900 ft/s and 5150 ft/s

respectively, were used for calculations of the parameters shown in Table 4.3. A
78
79
Figure 4.11. Plots showing normalized square root energy coefficients as a function of incidence angle for a P-wave (left), a
SH-wave (middle), and a SV-wave (right) incident on the interface between unsaturated overburden (medium 1) and saturated
overburden (medium 2). Calculations were made with the computer program presented in Appendix B. See text for discussion.
Vp Density P-wave S-wave
Vp Vs Vs Poisson’s
Lithology Vp/Vs Reflection Reflection
(ft/sec) (m/sec) (ft/sec) (m/sec) Ratio (ı) (g/cm3) Coefficient Coefficient

Unsaturated
1900 579 700 213 2.7 0.42 1.55
Overburden
0.54 0.11
Saturated
5150 1570 700 213 7.3 0.49 1.92
Overburden
0.08 0.61

80
Bedrock 5500 1676 2750 838 2.0 0.33 2.55

Table 4.3. Representative subsurface parameters and normal incidence P-wave and S-wave cumulative reflection coefficients
for the I-70 study area subsurface lithologies and related contrasts in impedance. See text for discussion.
variation in S-wave velocity across the unsaturated and saturated overburden interface

was not detectable using seismic data, and an S-wave velocity of 700 ft/s was assumed as

a representative value for the study area overburden in calculations. A contrast of 0.37

g/cm3 was assumed as a density contrast at the interface between the unsaturated and

saturated overburden. This was based on average density values of 1.55 g/cm3 for dry

overburden materials and 1.92 g/cm3 for wet overburden materials (Telford et al., 1990).

The plots in Figure 4.11 demonstrate that for this geologic situation:

1) The magnitude of the P-wave reflection coefficient is much larger for all

angles of incidence than the magnitude of the SV- or SH-wave

reflection coefficients.

2) The energy of the reflected and refracted mode-converted waves from

an incident P- or SV-wave are seen to be relatively small for all angles

of incidence.

For common-mode reflected P- and S-waves, the terms “square root energy coefficient”

and “reflection coefficient” represent the same quantities (Chapter 2). The P-wave

impedance contrast affecting the results in Figure 4.11 resulted from both velocity and

density contrasts, but the S-wave impedance contrast affecting the results in Figure 4.11

resulted only from density contrasts.

The magnitude of the P-wave reflection coefficient at normal incidence is 0.54,

while the magnitude of the S-wave reflection coefficient at normal incidence is 0.11

(Table 4.3, Figure 4.11). Using a bedrock S-wave velocity of 2750 ft/s, an approximated

bedrock P-wave velocity of 5500 ft/s (assuming a Vp/Vs ratio of 2.0), and an average

81
bedrock bulk density of 2.55 g/cm3 (ODNR, 2002), normal incidence (cumulative) square

root energy coefficients at the underlying overburden-bedrock interface are considered.

The cumulative square root energy coefficients of reflected P- and S-waves (from

incident P- and S-waves respectively) from the top-of-bedrock are 0.08 and 0.61

respectively (Table 4.3).

General conclusions can be made based on calculations presented in this section,

as follows:

1) P-wave energy reflected from the top-of-saturated-overburden (from an

incident P-wave) is large relative to the S-wave energy reflected from

this interface (from an incident S-wave).

2) P-wave energy reflected from the top-of-saturated-overburden (from an

incident P-wave) is large relative to the P-wave energy reflected from

the top-of-bedrock (from an incident P-wave).

3) S-wave energy reflected from the top-of-bedrock (from an incident S-

wave) is large relative to the S-wave energy reflected from the top-of-

saturated-overburden (from an incident S-wave).

4) S-wave energy reflected from the top-of-bedrock (from an incident S-

wave) is large relative to the P-wave energy reflected from the top-of-

bedrock (from an incident P-wave).

4.4.2 Noise and Optimal Reflection Windows

The air wave and high amplitude Rayleigh wave noise (in data recorded using X-

and Z-oriented receivers) had arrival times at the near offsets that were similar to those of

82
P-wave reflections from the top-of-saturated-overburden in the ZZ component data. The

noise could not be sufficiently suppressed through frequency filtering without degrading

the reflection signal, since the noise existed within the optimum P-wave reflection signal

frequency range (Figure 4.8). A frequency-wavenumber (f-k) filter that was applied to

the ZZ component data (prior to stacking) suppressed the air wave and the surface wave

noise, but the necessity of attenuating the noise resulted in a narrow optimum window to

enhance the P-wave event. The YY data S-wave event from the top-of-bedrock was not

severely degraded by air wave or surface wave noise. The YY component (top-of-

bedrock) S-wave reflection window was wider than the ZZ component (top-of-saturated-

overburden) P-wave reflection window (Figure 4.8). A larger number of traces

containing reflection energy could therefore be summed and stacked for the YY

component data than for the ZZ component data.

Rayleigh wave noise, air wave noise, and P-wave energy reflected from the top-

of-saturated-overburden (from a Z-oriented source), are strongly polarized in the vertical

component, and to a lesser extent these sources of energy are also present on the

horizontal component in the direction of the line (X-component). An X-directed

component of the P-wave reflections in Figure 4.9b are observed with relatively low

amplitude in X-oriented receiver gathers (top), and are not interpretable in Y-oriented

receiver gathers (middle). The top-of-saturated-overburden could therefore be most

effectively imaged using common-mode P-wave reflections in field gathers acquired

using Z-oriented receivers. Reflected (non-converted mode) S-wave energy correlating

to the top-of-bedrock, is also observed on certain gathers in Figure 4.9b recorded using

X- and Y-oriented receivers (for instance, at about 110 ms in the 48788 ZX gather), and

83
this resulted from S-wave generation by the Z-oriented source. As previously discussed

(see above), seismic sources are not perfectly pure in a polarization sense. A substantial

amount of the total energy distributed by a vertically oscillating seismic source is

generally in the form of S-waves (Miller and Pursey, 1955).

4.4.3 Resolution Potential of ZZ Component Data

The resolution potential of common-mode S-wave reflection data is addressed in

Chapter 6, and the results apply to the S-wave data that are the focus of this chapter. This

section focuses on resolution issues related to common-mode P-wave reflection data.

Synthetic seismograms were generated in order to investigate the resolution potential of

ZZ component data relative to the field study area geology. These synthetic seismograms

were generated using an acoustic finite-difference modeling code within the ProMAX

(Landmark Graphics Corporation) software package (see Chapter 6 for details regarding

the code, and for discussion regarding the selection of modeling parameters).

The P-wave interval velocity model in Table 4.4 was used to generate the

synthetic data presented in this section, and the model was constructed using velocities

that were derived through the analysis of ZZ component shot gathers. The model

contains P-wave velocities for the unsaturated and saturated overburden materials that

were derived from NMO and refraction analyses. P-wave velocities for the bedrock and

a coal seam beneath the model overburden were approximated for modeling purposes.

These quantities were not measurable directly from the data, and were approximated by

assuming a Vp/Vs ratio of 2.0 for both materials. S-wave bedrock velocities obtained

from refraction measurements across the study area ranged from 2500 ft/s to 3000 ft/s,

84
P-wave interval velocity (ft/sec) Lithology Layer thickness (ft)
1900 Unsaturated Overburden 29.5
5150 Saturated Overburden 11.5
5500 Bedrock (shale) 20
4790 Coal (bituminous) 7
5500 Bedrock (shale) 112

Table 4.4. Velocity model used to generate the synthetic seismogram in Figure 4.12.

and a P-wave velocity of 5500 ft/s was specified for the model bedrock. An S-wave

velocity of 2395 ft/s, was measured by Wolfe et al. (1989) for the Lower Freeport Coal (a

bituminous coal located stratigraphically beneath the Upper Freeport Coal). Based upon

this coal S-wave velocity value, a P-wave velocity of 4790 ft/s was specified for the

model coal. Drill log data from the site were used to establish representative thicknesses

for materials in the model that could not be determined from the seismic data. Synthetic

seismograms generated with models containing approximations for material bulk

densities did not change the main conclusions that are demonstrated using the modeling

results presented in this section.

A velocity versus depth plot from the P-wave interval velocity and the layer

thickness model in Table 4.4 is shown in Figure 4.12a. The calculated arrival times of

events from the model interfaces are plotted as a function of offset in Figure 4.12b. An

uninterpreted seismogram generated using a zero-phase Ricker wavelet (with a center

frequency of 150 Hz) as the surface-located source is shown in Figure 4.12c. A 150 Hz

source wavelet resulted in a dominant peak-to-peak reflection frequency of 125 Hz. It is

85
Figure 4.12. Comparison of synthetic data with line EBTravel field data (ZZ component).
Plots of the velocity model (Table 4.4) and calculated event arrival times are shown in (a)
and (b) respectively. A synthetic seismogram generated using the model in (a) with a 150
Hz source is shown uninterpreted in (c), and interpreted in (d). A shot gather used as a
basis for forward modeling is shown uninterpreted in (e), and interpreted in (f). A high
reflection coefficient at the unsaturated and saturated overburden (primary) interface,
noise, and interference prevent the interpretation of secondary events in field data (f).
86
seen from the interpreted synthetic seismogram in Figure 4.12d, that the reflection from

the top-of-saturated-overburden (the primary reflection) dominates the record at all

source-to-receiver offsets. A low amplitude reflection from the top of the coal seam

(having opposite polarity than the primary reflection) is interpretable from the synthetic

data at near offsets, but reflections from the other model interfaces located beneath

saturated overburden are not interpretable.

Unsaturated and saturated overburden P-wave velocities were obtained (after

bandpass filter and AGC application) from the shot gather in Figure 4.12e (uninterpreted)

and in Figure 4.12f (interpreted). The dominant reflection frequency of the synthetic data

and field data shown in Figure 4.12 are comparable, and the interpretations in Figure

4.12d are superimposed on the field data in Figure 4.12f. The observed fit of the same

reflection hyperbola and the same linear refraction (from the top-of-saturated overburden)

to the synthetic and field data supports the initial interpretation of these events and their

velocities in the field data. There is no evidence suggesting the presence of additional

events (from the top-of-bedrock or the coal seam) beneath the top-of-saturated-

overburden in the field data. This results from many factors, including: 1) the high

amplitude surface wave noise present (after optimum bandpass frequency filtering) at

near offsets in the field data, 2) the high P-wave reflection coefficient at the unsaturated

and saturated overburden interface, 3) low reflection coefficients at interfaces located

beneath the top of saturated overburden, and 4) interference and poor resolution.

General conclusions can be made from the analysis of the resolution potential of

ZZ component data discussed in this section, as follows:

1) P-wave energy reflected from the top-of-saturated-overburden dominates

87
ZZ component synthetic data at all source to receiver offsets.

2) A high reflection coefficient at the top-of-saturated-overburden, lower

reflection coefficients at deeper interfaces, noise, interference, and poor

resolution prevent the interpretation of reflection events from below the

top-of-saturated overburden in ZZ component field records.

4.4.3.1 P-Wave Versus S-Wave Resolution

Seismic wavelength (Ȝ) affects vertical and lateral resolution, and is related to

wave velocity (V) and frequency (f): Ȝ = V / f. Using dry overburden average P- and S-

wave velocities of 1900 ft/s and 700 ft/s respectively, and average dominant P- and S-

wave data frequencies of 125 Hz and 80 Hz respectively, then the quarter-wavelengths

(Ȝ/4) of P- and S-waves in dry overburden are 3.8 feet and 2.2 feet respectively (Table

4.5). Using saturated overburden average P- and S-wave velocities of 5150 ft/s and 700

ft/s respectively, and average dominant P- and S-wave frequencies of 125 Hz and 80 Hz

respectively, then Ȝ/4 values for P- and S-waves in saturated overburden are 10.3 feet and

2.2 feet respectively (Table 4.5). Ignoring other factors that influence resolution

potential, these calculations suggest that the resolution that can be achieved using S-

waves in the study area dry overburden is more than 1.7 times that which can be achieved

using P-waves, and that the resolution of S-waves in the study area saturated overburden

is more than 4.7 times that of P-waves (Table 4.5).

88
Average Average Quarter- Quarter-
Dominant Dominant wavelength wavelength
Average Average Vs
Lithology P-wave S-wave (Ȝ/4) for (Ȝ/4) for (P-waves Ȝ/4) /
Vp (ft/sec) (ft/sec)
Frequency Frequency P-waves S-waves (S-waves Ȝ/4)
(Hz) (Hz) (ft) (ft)

Dry
1900 700 125 80 3.8 2.2 1.7
Overburden

Saturated
5150 700 125 80 10.3 2.2 4.7
Overburden

89
Table 4.5. Quarter -wavelengths calculated for P- and S-waves in the I-70 study area overburden materials.
4.4.4 Data Processing

Common-mode P- and S-wave reflection data processing and imaging operations

applied to each of the components (YY and ZZ) were established based on previous

analyses of the data (see above). The analysis flows applied to each of these data

components are described in Table 4.6. A discussion regarding data pre-processing was

presented in Chapter 3. An event correlating as the top-of-bedrock could consistently be

identified in S-wave records acquired in the region east of the previous roadway collapse.

Processing and imaging operations applied to the YY component data presented in this

chapter were determined for the purpose of enhancing reflections from this impedance

contrast. ZZ component records contained P-wave reflections correlating to depths of the

top-of-saturated-overburden, and processing of ZZ component data focused on enhancing

reflections from this impedance contrast. As previously mentioned, P-wave reflections

from the top-of-saturated overburden were not identifiable on as large of a percentage of

shot gathers, as were S-wave reflections from the top-of-bedrock (due to event arrival

times and recorded noise characteristics). S-wave data required more accurate stacking

velocities than did P-wave data for producing quality stacks. A given deviation from

optimum stacking velocity would represent a much larger percentage of the optimum

stacking velocity in the S-wave data case, since stacking velocities for S-wave data were

much lower than those for P-wave data.

4.4.4.1 Frequency-Wavenumber Filtering

For the common-mode component shot gathers presented in this chapter, high

amplitude, linear, non-reflected energy remained in records after optimum bandpass

90
Processing step Description
Data reformat From SEG-Y to ProMAX format
Vibroseis correlation Both lines correlated with pilot sweep
Geometry Defined using field notes and loaded to headers
Data truncation Records truncated to 300 ms
Trace editing Bad / noisy traces killed
Trace equalization 150 ms spatially varying window
f-k filter Non-reflection energy/linear noise suppression
CMP sort Sorted from shot gathers to midpoint gathers
Integrated analysis of shot gathers, constant velocity stacks,
Velocity analysis and semblance plots for YY component, Integrated analysis of
shot gathers and constant velocity stacks for ZZ component
NMO correction Applied based on optimum stacking velocities
30 percent for line EBTravelYY and 40 percent for line
Stretch mute
EBTravelZZ
Zero-phase Ormsby filters: 50-80-160-200 Hz for YY
Bandpass filter
component, and 80-120-200-240 Hz for ZZ component
AGC scaling 100 ms window
CMP ensemble / stack Summed NMO-corrected CMP gathers

Table 4.6. Data processing flow for line EBTravel S-wave (YY component) and P-wave
(ZZ component) reflection data.

91
frequency filtering. An f-k (frequency-wavenumber) filter was therefore evaluated and

applied to each of the common-mode component shot records (after amplitude balancing)

to suppress coherent, linear non-reflection energy. This processing step was critical for

suppressing high amplitude surface wave noise (having arrival times and frequency

content at near offsets similar to those of P-wave reflections) recorded in ZZ component

data. f-k filtering was not as critical of a process for the XX or YY data, as at the arrival

times of S-wave reflections linear noise was recorded at far offsets, and could therefore

be largely suppressed after NMO corrections by proper stretch muting. However,

evaluation of f-k filter effects on the XX and YY component data showed that an

improvement in stacked signal quality, with minimal artifact generation was obtained.

The effectiveness of f-k filtering in improving records (as determined by noise

suppression and reflection signal enhancement) prior to CMP sorting and stacking is

demonstrated using ZZ and YY component field records from line EBTravel (Figures

4.13a and 4.13b respectively). These data were recorded with the respective source

located at road station 48638, and both records contained identifiable reflections prior to

application of f-k filters. Bandpass filters and AGC were applied to the gathers before

generating the plots in Figure 4.13, in order to demonstrate non-reflected energy

suppression using f-k filters across reflection signal bandwidths.

The shot gathers are shown in Figure 4.13 for data prior to f-k filter application,

along with their 2-D amplitude spectra (left). Superimposed on these amplitude spectra,

are the mute polygons defined to suppress steeply dipping linear noise (of relatively low

apparent velocity) evident in both components. Noise for both components would

degrade reflection signal quality upon stacking if not suppressed. For example, see the

92
93
Figure 4.13. Line EBTravel ZZ component (a) and YY component (b) shot gathers (source located at road station 48638, east
direction to left) and f-k spectra: (left) gathers without f-k filter applied showing reflections with zero-offset times of 32 ms (a)
and 110 ms (b), and f-k amplitude spectra showing defined mute polygons, (middle) with polygons rejected to suppress noise
(indicated in boxes 1 and 2), and (right) with polygons accepted (showing noise rejected through filter application). Bandpass
filters and AGC gain were applied to the gathers before generating these plots in order to demonstrate f-k filter non-reflection
energy suppression across reflection signal bandwidths. The x-axis scales of absolute offset from the sources are in feet.
effect of noise suppression at absolute offsets of 0.0 to 34.0 feet at ~32 ms on the left side

of the ZZ record spread (see box 1 in Figure 4.13a), and at absolute offsets of 70.0 to

90.0 feet at ~110 ms on the left side of the YY record spread (see box 2 in Figure 4.13b).

The concentration of linear, non-reflection energy for both components (within the

defined mute polygons) is seen to be well isolated from the reflected energy (energy

concentrated outside of the defined mute polygons, around the frequency axes) in the

amplitude spectra.

Steeply dipping noise is largely suppressed on the ZZ and YY component shot

gathers after applying the f-k filters (by rejecting the mute polygons as seen on the middle

plots in Figure 4.13). Applying the filters with the f-k polygons so that they accepted

rather than rejected the same f-k ranges (as seen on the right plots in Figure 4.13), shows

that the noise that is rejected through application of the filters, and demonstrates that

minimal reflection signal of interest is lost when the filters are applied. Line EBTravel

ZZ component (top) and YY component (bottom) stacked sections without and with pre-

stack f-k filtering included the processing flows (Table 4.6) are shown in Figure 4.14.

These stacks demonstrate improvements in both P-wave and S-wave stacked reflection

signal quality that were obtained through pre-stack f-k filtering.

General conclusions can be made from the analyses of f-k filtering of ZZ

component and YY component data discussed in this section, as follows:

1) f-k filtering applied to the ZZ component and the YY component data is

effective for suppressing coherent noise and enhancing reflection signal.

2) Improvement in stacked signal quality of the ZZ and the YY component

data can be obtained by f-k filtering with minimal artifact generation.

94
95
Figure 4.14. Line EBTravel ZZ component (top) and YY component (bottom) time sections without and with f-k filtering in
processing flows (Table 4.6). Notice the suppression of noise and enhancement of reflection signal (within the circled regions)
achieved through application of f-k filters. CDP location numbers (CDP_X) correspond to road stations (units are in feet).
4.4.5 P- and S-Wave Stacked Section Imaging

Processed (Table 4.6) and stacked line EBTravel P-wave (ZZ component) and S-

wave (YY component) sections are shown in Figure 4.15. The P-wave reflection event

(with an average dominant frequency of 125 Hz) at 28 to 33 ms in Figure 4.15a correlates

as the top-of-saturated-overburden on the P-wave depth section (Figure 4.15c). The

appearance of this reflection event on the P-wave section has been affected from the

stacking of a certain amount of noise along with reflection signal. This is due to the

frequency content and related slopes and velocities of the noise relative to P-wave

reflections. Noise modes could not be as effectively suppressed in ZZ component data as

they could be in YY data. The S-wave reflection event (with an average dominant

frequency of 80 Hz) at 105 to 115 ms in Figure 4.15b correlates as the top-of-bedrock on

the S-wave depth section (Figure 4.15d). These interpretations are supported by the

results obtained through shot gather analyses, modeling, and drill log data (see above).

A comparison of sections from both (ZZ and YY) components (and consideration

of the results previously presented in this chapter) indicates that there are specific

advantages and disadvantages associated with P- and S-wave reflection data acquired in

the study area, as follows:

1) The top-of-saturated-overburden served as a detectable impedance

contrast for P-waves but not for S-waves.

2) P-wave data cannot be used to image impedance contrasts located below

the saturated overburden, while images of the top-of-bedrock can be

constructed using S-wave reflections.

96
97
Figure 4.15. Line EBTravel ZZ component (a) and YY component (b) time sections with fold (TR_FOLD) plots. The P-wave event at 28
to 33 ms (a) is the top-of-saturated-overburden (blue) on the depth section (c). The S-wave event at 105 to 115 ms (b) is the top-of-bedrock
(yellow) on the depth section (d). The scales on the bottom x-axes of (a) and (b) show P-wave and S-wave stacking velocities respectively
(velocity scales are to the right of the sections). CDP location numbers (CDP_X) correspond to road stations (units are in feet). (continued)
Figure 4.15. (continued)

98
(c) and (d) Line EBTravel ZZ component and YY component depth sections. CDP location numbers (CDP_X) correspond to
road stations (given in feet from the western county line). See previous page for complete caption.
3) S-wave data offer the potential for allowing possible areas of the

subsurface where subsidence processes have been active, to be identified

based on observations of disruptions in the bedrock horizon, whereas P-

wave data do not provide this potential.

4.4.5.1 Insitu Physical Properties Measurement Potential Using Reflections

Seismic reflection measurements offer the potential for determining insitu elastic

properties of soil and rock volumes in a cost-effective manner. Although geotechnical

measurements can be made using boreholes at a sampling interval that is typically finer

than seismic data resolution potential, elastic properties determined from seismic

measurements have the advantage of not being potentially affected by material

disturbances related to sampling or tool insertion.

It is possible in theory to produce a subsurface map showing both lateral and

vertical variations in the Vp/Vs ratio, through the correlation of P- and S-wave reflections.

This has previously been accomplished for petroleum exploration purposes, in deep earth

studies that have acquired multicomponent reflection data (Garotta, 2000). It is also

possible to produce a subsurface map showing variations in Poisson’s ratio ( σ , the ratio

of transverse contraction to the longitudinal extension when a rod is stretched), because

the Vp/Vs ratio and σ can be related. The relationship between Vp/Vs (Vp = P-wave

velocity, and Vs = S-wave velocity) and σ can be described in terms of Vp/Vs as:

Vp/Vs = ((1 - σ ) / (0.5 - σ )) 2 (4.4)

99
and in terms of σ as:

σ = ((Vp 2 / 2Vs 2 ) - 1) / ((Vp 2 / Vs 2 ) - 1) (4.5)

The Vp/Vs ratio (and therefore σ ) of a geologic material is a function of many

factors, including lithology, porosity, cementation, depth, age, temperature and pressure

regime, and interstitial fluids (Tatham and McCormack, 1991). Typical values of Vp/Vs

range from 1.7 to 2.0 for rocks, and are often in the range of 2.0 to 7.0 for shallow

unconsolidated sediments (Hasbrouck and Padget, 1982). Because the elastic constants

of geologic materials are defined as positive numbers, values of σ for geologic materials

are within the range of 0.0 to 0.5 (a maximum value of 0.5 corresponds to fluids with no

shear strength). In petroleum exploration studies, σ values around 0.05 have been

measured for very hard, rigid materials, while σ values between 0.25 and 0.33 have

typically been measured for limestones, sandstones, and many common igneous and

metamorphic rocks (Telford et al., 1990; Dobecki, 1993). For soft, shallow

unconsolidated geologic materials, σ values between 0.45 and 0.49 are common

(Dobecki, 1993), with σ values in the range of 0.496 to 0.498 typical for near-surface

clayey, saturated sediments (Benjumea et al., 2001).

The potential for producing a map showing detailed variations in σ for geologic

materials in the I-70 subsurface, based on acquired data reflection information, is limited.

This limitation results from the small number of reflection events in acquired data, and

due to the fact that P- and S-wave reflections from similar subsurface interfaces cannot be

correlated (see above). Representative I-70 lithology values of σ were estimated (Table

4.3) using velocity information obtained from both reflection and refraction analyses, and

by making some assumptions.


100
Using a P-wave velocity of 1900 ft/s (measured from a reflection from the top of

saturated overburden) and a S-wave velocity of 700 ft/s (measured from a reflection from

the top-of-bedrock), and by assuming that S-wave velocities above and below the top-of-

saturated-overburden are the same, a Vp/Vs ratio of 2.7 and a σ value of 0.42 are

calculated for unsaturated overburden (Table 4.3). By using a refraction-derived P-wave

velocity of 5150 ft/s, and by making the same assumption regarding S-wave velocity as

that which was made in the previous example, a Vp/Vs ratio of 7.3 and a σ value of 0.49

are calculated for saturated overburden (Table 4.3). Bedrock S-wave velocities could be

measured from refractions in field data, however, bedrock P-wave velocities could not be

directly measured from the acquired data, and calculations of a representative σ for

bedrock were therefore dependent upon an assumed Vp/Vs ratio (Table 4.3). Because no

P- or S-wave velocities could be measured for the bituminous coal, there is no potential

to estimate σ of the coal in the study area using the acquired seismic data.

101
CHAPTER 5

CONVERTED-MODE REFLECTION IMAGING POTENTIAL

5.1 Overview

Research was conducted to test and evaluate the potential for effective near-

surface converted-mode (P-S and S-P) seismic reflection imaging. In recent years,

multicomponent surveying and converted-wave imaging technologies have rapidly

developed for petroleum exploration purposes, and numerous benefits of deep-earth

converted-wave acquisition and imaging (predominantly P-S) have been recognized (see

Chapter 1 for an overview of previous work). However, documented research concerning

the practicality of using converted-mode reflection information, for the purpose of

improving near-surface characterization, has been non-existent to date.

The potential for near-surface converted-mode reflection imaging was tested

experimentally and evaluated through numerical modeling. Multicomponent reflection

data were acquired along an un-mined section of the I-70 study area, where the geology

(unsaturated and saturated overburden materials overlying consolidated units) is

representative of a subsurface environment class typically encountered during near-

surface seismic studies. Common-mode P- and S-wave reflection events from known

lithologic boundaries were observed in acquired field records, and media parameters

102
derived from the acquired field data were used as a basis for interface energy partitioning

analyses, event travel-time and moveout predictions, and elastic-wave modeling.

Zoeppritz equation solutions indicated that primary mode-conversions (i.e. events

resulting from mode-conversion at reflection from I-70 subsurface interfaces) have high

cumulative energies relative to other mode-converted events resulting from conversion at

transmission across subsurface interfaces. Events having arrival times and moveouts

similar to those predicted for primary mode-conversions however, were not observed in

acquired multicomponent reflection data. Analyses of synthetic seismograms indicated

that potential (resolution-related) benefits of converted-wave reflection (P-S and/or S-P)

imaging exist in theory (under noise-free conditions) for the I-70 subsurface conditions.

Modeling results also indicated that such benefits were unable to be practically realized

using field measurements, due to the low amplitudes of mode-converted reflections

relative to those of common-mode reflections, and the detrimental affects of random

noise and coherent noise modes recorded in field data.

Several factors affecting seismic mode-conversion imaging potential were found

to be different between previous deep-earth multicomponent studies (in which effective

converted-wave imaging has been possible) and the near-surface I-70 multicomponent

field study (in which effective converted-wave imaging was found to not be possible). At

deep-earth interfaces targeted for petroleum-related converted-wave imaging, there

generally is a comparable amount of mode-converted (P-S) and common-mode (P-P)

reflection energy generated at moderate angles of incidence (Stewart et al., 2002).

Targeted deep-earth converted-wave events are also often well separated in time and

space from other reflections and detrimental modes of coherent-noise. At the I-70
103
subsurface interfaces of interest, it was found that (due to media impedance relationships)

there was relatively little energy partitioned into converted-modes, and that scattered

mode-converted (P-S and S-P) energy was not comparable to scattered common-mode

(P-P and S-S) energy at any recorded incidence angles. Further, arrival-times of modeled

converted-wave reflections from the I-70 interfaces were found to be similar to the

observed arrival-times of coherent noise modes recorded in the I-70 field data. The

results of this research indicate that from a practical standpoint, it will be difficult under

subsurface and environmental conditions that are similar to those of the I-70 study area,

to use mode-converted reflection imaging for improving near-surface characterization.

The analysis methods developed through this research can be used as a basis for assessing

converted-wave imaging potential in other possible future studies conducted elsewhere.

5.2 Field Multicomponent Reflection Data

Experimental multicomponent reflection data were acquired in the I-70 study area

(Appendix A) during 2001, with an objective being to allow an evaluation of the potential

for constructing near-surface mode-converted (P-S and/or S-P) images. Information

regarding the 2001 reflection survey acquisition parameters, data geometry and data pre-

processing were presented in Chapter 3. Nine-component (9C) reflection data were

acquired along the south edge of the eastbound lanes of I-70 (referred to as seismic line

EBTravel), and are presented in this chapter. For line EBTravel (Figure 5.1), three

source configurations were used, which generated: preferential shear particle motion

inline and transverse (crossline) to the seismic line (source components X and Y

104
Figure 5.1. Map view of the eastbound lanes of I-70 (road stations 48600 to 48850)
showing the locations of multicomponent seismic reflection lines (road stations 48630 to
48830) presented in this chapter, relative to the locations of the underground mine
workings. For line EBTravel 9 component (9C) data were acquired, and for line EBPass
3C data were acquired. Road stations are given in feet from the western county line.

respectively), and preferential compressional particle motion (source component Z).

Three-component (3C) reflection data are also presented from line EBPass, which was

acquired (across the roadway and parallel to line EBTravel) along the northern edge of

the eastbound lanes of I-70 (Figure 5.1). For line EBPass, one (Y component) source

configuration was used. For both line EBTravel and line EBPass, a single 3-component

geophone was planted at each station. Each geophone contained two (orthogonal)

horizontal elements oriented inline and transverse to the line (receiver components X and

Y respectively), and a vertical element (receiver component Z). Nomenclature used in

this chapter to describe multicomponent reflection data, in terms of source and receiver

orientations and preferential polarizations, is described in Figure 5.2.

105
Figure 5.2. Nomenclature used to describe multicomponent reflection data, in terms of
the source and receiver orientations and preferential polarizations. For sources, the X and
Y symbols indicate sources configured to preferentially generate shear particle motion
inline and transverse (crossline) to the seismic line respectively, while Z indicates a
source configured to preferentially generate compressional particle motion. For
receivers, the symbols X and Y indicate horizontal geophone elements oriented inline and
transverse to the line respectively, while Z indicates a vertical geophone element. The
source and receiver pairs: XX, YY, and ZZ are referred to as the common-mode
components of the nine-component matrix (e.g. XX means a source and receiver both
oriented in the X direction)

106
5.2.1 Common-Mode Reflection Information

Field shot gathers from the full 9C data set acquired along line EBTravel (Figure

5.1) are shown in Figure 5.3. The gathers are shown unprocessed with individual trace

amplitudes normalized in Figure 5.3a, and after bandpass filter and gain application in

Figure 5.3b. Common-mode S-wave reflections from the overburden and bedrock

interface are indicated on the XX component and the YY component shot gathers in

Figure 5.3b. On the ZZ component gathers in Figure 5.3b, common-mode P-wave

reflections from the unsaturated overburden and saturated overburden interface are

indicated. The reasons why these horizons could be effectively imaged using either P- or

S-wave common-mode reflection information respectively, are explained in Chapter 4.

Common-mode reflection energy was often observed on non-diagonal source and

receiver component pairs. For instance, the S-wave reflection event from the top-of-

bedrock in XX component records was also recorded in the XY component (Figure 5.3b).

Numerous factors could be responsible for this, including: source or geophone orientation

errors, the non-perfect nature of sources and geophones in a polarization sense, cross talk

between geophone elements, cross dip of reflecting horizons with respect to the seismic

line, and/or near-surface media anisotropy. S-wave common-mode reflection energy that

was occasionally observed on multicomponent field records that were acquired using a

vertical-oriented (Z component) source, suggest that the vertical source configuration put

a substantial amount of S-wave energy in to the ground. This observation suggests that it

may be possible (depending upon many factors) to obtain useful subsurface S-wave

information during traditional near-surface P-wave reflection surveys, simply by

deploying 3C rather than 1C (vertical element) geophones.


107
108
Figure 5.3. Line EBTravel 9C shot gathers: (a) unprocessed, and (b) with a bandpass filter (X and Y sources: 50-80-160-200
Hz; Z source: 80-120-200-240 Hz) and AGC applied. Common-mode S-wave reflections (yellow) from the top-of-bedrock are
indicated on XX and YY shots; common-mode P-wave reflections (blue) from the top-of-saturated-overburden are indicated on
ZZ shots. Source locations (SOU_X) are in feet from the western county line; AOFFSET units are in feet. (continued)
Figure 5.3. (continued)

109
Common-mode reflection energy was often observed on non-diagonal source and

receiver components, however matched source and receiver pair data contained the

highest amplitude and cleanest common-mode reflection energy. This was explained in

Chapter 4, and is further demonstrated in this chapter through a comparison of

multicomponent depth section images (Figure 5.4). Data components acquired using X-

and Y-oriented sources were processed for common-mode S-wave reflection energy

(Figures 5.4a and 5.4b), while data components acquired using a Z-oriented source were

processed for common-mode P-wave reflection energy (Figure 5.4c). The processing of

these components followed the methodologies and workflows that were previously

developed and described in Chapter 4.

A comparison of the 9C depth sections in Figure 5.4 demonstrates that the highest

quality images of target horizons (i.e. the top-of-bedrock and the top-of-saturated-

overburden) could be generated through the processing of diagonal data components.

These horizons are interpreted in the YY component section and the ZZ component

section respectively (Figure 5.4). No mining activity previously occurred beneath line

EBTravel to the east of road station 48655 (Figure 5.1), and the YY component depth

section indicates that the bedrock horizon is coherent and continuous to the east of this

road station. Acquired 3C shot gathers (Figure 5.5) and processed depth sections (Figure

5.6) from line EBPass (Figure 5.1) also demonstrate that the highest quality imaging of

the bedrock horizon could be accomplished using diagonal (YY) component S-wave

reflection information. A normal fault at road station 48643 was interpreted from YY

data to be the result of the bedrock subsiding into a mine room (Figure 5.6).

110
111
(a) Line EBTravel 9C depth stacks: XX, XY, and XZ components.

Figure 5.4. Line EBTravel 9C depth stacks: (a) XX, XY, and XZ components, (b) YX, YY, and YZ components, (c) ZX, ZY,
and ZZ components. S-wave source configuration sections (a) and (b) were processed for common-mode S-wave reflection
energy; the S-wave reflection indicated on the YY component stack is from the top-of-bedrock. P-wave source configuration
sections (c) were processed for common-mode P-wave reflection energy; the P-wave reflection indicated on the ZZ component
stack is from the top-of-saturated-overburden. CDP locations (CDP_X) equal road stations (road stations are given in feet from
the western county line). See text for discussion. (continued)
Figure 5.4. (continued)

112
(b) Line EBTravel 9C depth stacks: YX, YY, and YZ components. (continued)
Figure 5.4. (continued)

113
(c) Line EBTravel 9C depth stacks: ZX, ZY, and ZZ components.
114
Figure 5.5. Line EBPass 3C shot gathers: (a) unprocessed, and (b) with a bandpass filter (50-80-160-200 Hz) and AGC (100 ms
window) applied. Common-mode S-wave reflections from the top-of-bedrock are indicated on YY gathers. The x-axis unit of
absolute offset is feet. Source locations (SOU_X) are given in feet from the western county line.
115
Figure 5.6. Line EBPass 3C depth stacks processed for common-mode S-wave reflection energy: (top) YX, (middle) YY, and
(bottom) YZ. The S-wave reflection indicated on the YY component stack is from the top-of-bedrock. CDP locations (CDP_X)
equal road stations (road stations are given in feet from the western county line). See text for discussion.
5.3 Converted-Mode Reflection Imaging Potential

In order to process and image possible mode-converted reflections (Chapter 2)

recorded in seismic data, such events must be confidently identified in field records.

Recognition of mode-converted reflections in field data requires that P- and S-wave

velocity and density structure (as a function of depth) be approximated, in order to

predict through modeling, mode-converted event travel-times and amplitudes.

5.3.1 Near-Surface Velocity and Density Structure

The study area subsurface (in un-mined areas) represents a subsurface

environment class that is typically addressed by shallow reflection surveying (i.e. a

relatively flat-lying geologic sequence comprised of unsaturated and saturated

overburden above consolidated units). Representative depth model velocity and density

parameters for the study area subsurface are presented in Table 5.1, and are shown

graphically in Figure 5.7. It is assumed that plane boundaries separate flat layers within

the model, and that the model layers consist of homogeneous and isotropic media.

S-wave velocities (Vs) for unsaturated overburden, saturated overburden, and

bedrock (shale) in the model (Figure 5.7) are seismic data-derived quantities. Optimum

S-wave stacking velocity differences of less than three percent were determined from XX

and YY component data (Chapter 4). The coal Vs could not be measured from acquired

seismic data, and a Vs estimate for coal was therefore specified based on measurements

by Wolfe et al. (1989) for the Lower Freeport Coal (stratigraphically beneath the Upper

Freeport Coal). Representative P-wave velocities (Vp) for unsaturated and saturated

116
Model Vp Density Layer Layer
Vp Vs Vs Poisson’s
Vp/Vs Thickness Thickness
Layer (ft/sec) (m/sec) (ft/sec) (m/sec) Ratio (ı) (g/cm3) (ft) (m)

Unsaturated
1900 579 700 213 2.7 0.42 1.55 29.5 9.0
Overburden

Saturated
5150 1570 700 213 7.3 0.49 1.92 11.5 3.5
Overburden

Bedrock 5500 1676 2750 838 2.0 0.33 2.55 20 6.1

117
Coal Seam 4790 1460 2395 730 2.0 0.33 2.30 7 2.1

Bedrock 5500 1676 2750 838 2.0 0.33 2.55 32 9.8

Table 5.1. Data-derived I-70 subsurface model layer parameters used for square root energy coefficient calculations and elastic-
wave modeling. See text for discussion.
Figure 5.7. Cross-section view of the I-70 depth model (Table 5.1) used for square root
energy calculations and elastic-wave modeling. See text for discussion.

overburden in the model (Figure 5.7) are seismic data-derived quantities. The Vp values

of bedrock and coal could not be measured from acquired seismic data, and Vp values for

these lithologies were therefore estimated by assuming Vp/Vs ratios of 2.0. The near-

surface velocity structure for P- and S-waves in the study area subsurface is quite

different (Figure 5.7). Density values in the model were based on average values for

unsaturated and saturated unconsolidated materials given in Telford et al. (1990), and on

averages of measured values (ODNR, 2002) for the study area bedrock and coal units in

the adjacent Harrison County, Ohio.

5.3.2 Energies of Mode-Converted Events

Using the depth model parameters in Table 5.1 and Figure 5.7, energy partitioning

among wave types from model layer boundaries (i.e. the relative energies of different
118
Figure 5.8. Square root energy coefficients versus incidence angle using the media
parameters in Table 5.1: (a) downgoing waves at the top of saturated overburden, (b)
downgoing waves at the top of bedrock, and (c) upgoing waves at the top of saturated
overburden. Plots were generated using the code in Appendix B. See text for discussion.

119
mode-converted events) was evaluated. The plots in Figure 5.8 were generated with the

equations in Chapter 2, which were coded in the PSHSV program (Appendix B). The

plots show calculated square root energy coefficients as a function of incidence angle, for

both incident P- and SV-waves. The energy partitioning for numerous different event

raypaths can be evaluated from the plots, with such analyses providing insight into the

potential for imaging various mode-converted reflection events from the model sequence.

5.3.2.1 Mode-Conversions at the Unsaturated and Saturated Overburden Interface

From a downgoing P-wave incident on the unsaturated and saturated overburden

interface (Figure 5.8a), there are mode-conversions (i.e. P-SV reflection and P-SV

refraction) occurring at small angles of incidence. However, the energies of mode-

converted waves are small relative to the energies of scattered common-mode P-waves

waves originating at the interface (i.e. P-P reflection and P-P refraction). For the case of

a downgoing SV-wave incident on the unsaturated and saturated overburden interface

(Figure 5.8a), it is seen that most of the energy is partitioned into the refracted SV-wave

(due to the lack of S-wave velocity contrast), with much smaller percentages of energy

going into the common-mode SV-wave reflection and mode-converted waves (i.e. SV-P

reflection and SV-P refraction). There is a small incidence angle range (just before the

reflected P-wave critical angle) where the converted SV-P reflection has greater energy

than the common-mode SV-SV reflection, but SV-P wave energy is still relatively low.

Based on analyses of energy partitioning at the unsaturated overburden and

saturated overburden interface, conclusions are made, as follows:

120
1) The common-mode P-wave reflection has the highest energy of

reflected wave-types at the unsaturated overburden and saturated

overburden interface.

2) There is relatively little energy in common-mode or converted-mode

reflected waves of other types at the interface.

3) There is relatively little mode-converted energy refracted (transmitted)

across the interface from either incident (P or SV) wave-type.

5.3.2.2 Mode-Conversions at the Saturated Overburden and Bedrock Interface

To evaluate imaging potential of mode-converted reflections from the saturated

overburden and bedrock interface, energy partitioning of a downgoing wave at the top of

this interface must be considered, as must partitioning of a subsequent upgoing wave at

the unsaturated overburden and saturated overburden interface. Square root energy

coefficients resulting from downgoing waves at the top of bedrock are shown in Figure

5.8b, and coefficients resulting from upgoing waves at the top of saturated overburden

are shown in Figure 5.8c.

Many energy conversions and converted-mode reflection events (ray paths) from

the saturated overburden and bedrock interface are possible and can be considered from

the plots in Figure 5.8. For example, a P-wave originating at the ground surface can

travel (through the unsaturated and saturated overburden interface) to the saturated

overburden and bedrock interface as a P-wave, reflect as a SV-wave, and then travel back

up to the ground surface as a SV-wave (i.e. a P-P-SV-SV reflection event). This can be

referred to as a primary converted P-SV reflection, as it is a wave that converts only once,
121
at the deepest reflector (i.e. a downgoing P-wave that converts on reflection at the

horizon of interest into an upcoming SV-wave). Alternatively, a P-wave originating at

the ground surface can travel to the unsaturated and saturated overburden interface as a P-

wave, be transmitted as a SV-wave across this interface, reflect as a SV-wave at the

saturated overburden and bedrock interface, and then be transmitted as a SV-wave

upwards across the saturated and unsaturated overburden interface (i.e. a P-SV-SV-SV

reflection event). Figure 5.9 shows multiple mode-conversion point possibilities to a SV-

wave from an incident P-wave for an example geologic situation. Mode-converted

reflections represented by the different ray paths in Figure 5.9 have different amplitudes,

arrival times, and moveouts depending upon media compressional and shear impedances.

It can be possible depending upon impedance relationships to transmit a

significant amount of mode-converted energy (i.e. equal to or greater than the transmitted

common-mode energy) across an interface. It was shown by Purnell (1992) that this

could occur in a case where basalt overlaid sand, due to the close match between the

basalt Vs and the sand Vp. For the I-70 subsurface model (Figure 5.7), significant mode-

converted energy would not be transmitted across the unsaturated and saturated

overburden interface from a downgoing incident P- or SV-wave (Figure 5.8a). This

means that the highest amplitude mode-converted reflections from the saturated

overburden and bedrock interface would result from incident common-mode energy

transmitted (as common-mode energy) across the unsaturated and saturated overburden

interface. As previously mentioned, a downgoing wave that converts on reflection at the

horizon of interest into an upgoing wave of different type can be referred to as a primary

mode-converted reflection. Similar to the findings of these analyses, it has been shown
122
Figure 5.9. Multiple mode-conversion point possibilities to a SV-wave from an incident
P-wave for an example geologic situation. Point C indicates mode-conversion at
reflection from an interface of interest (i.e. a primary conversion), while points C’ and C”
indicate conversion of upward and downward traveling P-waves at transmission through
an overlying interface. Reflections represented by the different ray paths shown will have
different amplitudes, arrival times and moveouts depending upon media compressional
and shear impedances. Horizontal and vertical scales are not implied.

123
that in deep-earth petroleum studies, transmitted or multiple P-SV conversions generally

have much lower amplitudes than the primary (P-SV) reflection (Stewart et al., 2002).

For the case of a downgoing P-wave incident on the saturated overburden and

bedrock interface (Figure 5.8b), most of the incident energy is partitioned into a refracted

P-wave, due to the small contrast in P-wave velocity across the interface. Although

having low energy relative to a refracted P-wave, P-wave and mode-converted SV-wave

reflections do occur, with the reflected SV-wave energy surpassing the reflected P-wave

energy in magnitude as the incidence angle increases. For the case of a downgoing SV-

wave incident on the saturated overburden and bedrock interface (Figure 5.8b), there is

significant energy in both common-mode reflected and refracted SV-waves (due to a

large contrast in S-wave velocity across the interface). Although energy is partitioned

into the mode-converted P-wave reflection and refraction (from an incident SV-wave) for

angles of incidence that are smaller than the refracted SV-wave critical angle, the

energies of mode-converted P-wave events are small, and are non-zero only at near-

normal incidence angles.

For the case of an upgoing P-wave incident on the saturated and unsaturated

overburden interface (Figure 5.8c), most of the energy is either reflected at or refracted

across the interface as common-mode P-wave energy (due to the small S-wave velocity

contrast across the interface). The energy of mode-converted reflected and refracted SV-

waves is relatively small at the interface for all angles of incidence considered. The case

of an upgoing SV-wave incident on the saturated and unsaturated overburden interface

(Figure 5.8c) is similar, in that most of the energy is either reflected at or refracted across

the interface as common-mode SV-wave energy, with little energy partitioned into mode
124
conversions. These results indicate that there is relatively little potential for imaging

mode-conversion events occurring from upgoing wave interactions at the saturated and

unsaturated overburden interface.

Based on the analyses of mode-converted waves discussed in this section,

conclusions are made, as follows:

1) There is relatively little mode-converted energy refracted through the

unsaturated overburden and saturated overburden interface (from a

downgoing incident wave of either P- or SV-type) that will reach the

saturated overburden and bedrock interface.

2) The common-mode SV-wave reflection has the highest energy of

reflected wave-types considered at the saturated overburden and

bedrock interface.

3) Upgoing P- or SV-waves (either common-mode reflected or

converted-mode reflected at the saturated overburden and bedrock

interface) incident on the saturated and unsaturated overburden

interface will experience little energy partitioning into converted-

modes at this interface.

5.3.3 Ray Tracing of Mode-Converted Reflections

It was shown through energy partitioning analyses (Figure 5.8) in the previous

section, that mode-converted events (i.e. P-SV and SV-P) resulting from mode-

conversion at reflection from I-70 subsurface interfaces of interest have high cumulative

energies, relative to mode-converted events resulting from mode-conversion at


125
transmission across interfaces. Results indicated that mode-converted reflections,

occurring at reflection from a given horizon of interest (i.e. primary mode conversions),

have the highest potential (of the possible mode-conversion reflection ray paths) of being

imaged using field multicomponent seismic reflection measurements.

Common-mode (P-P and SV-SV) and converted-mode (P-SV and SV-P)

reflection ray trace plots through the I-70 subsurface model (Figure 5.7) are shown in

Figure 5.10. The ray tracing was done using the QUIKSHOT+ application (Landmark

Graphics Corporation). Synthetic seismic reflection survey parameters consisted of a

single shot located at 0.0 feet on the ground surface, and 80 geophones located on the

ground surface. The geophones had a 2.0-foot interval and a 90-degree azimuth from the

source. The synthetic survey parameters were designed to mimic those of the actual I-70

field multicomponent seismic reflection survey (see Chapter 3). The maximum source-

to-receiver offset shown in Figure 5.10 is equivalent to the maximum source-to-receiver

offset acquired during the I-70 field survey. The mode-conversion ray path segments

shown in Figures 5.10b and 5.10d resulted from incident wave energy partitioning upon

reflection at horizon interfaces (i.e. primary mode-conversions). For instance, in Figure

5.10b, a downgoing incident P-wave is transmitted through the unsaturated and saturated

overburden as a P-wave, is then reflected as an SV-wave at the saturated overburden and

bedrock interface, and then travels upward to the ground surface as a mode-converted

SV-wave reflection (i.e. a primary P-SV reflection).

The unsaturated and saturated overburden interface reflection point for a

common-mode (P-P or SV-SV) reflection is at the mid-point between the source and

receiver (Figures 5.10a and 5.10c), however, this is not the case for mode-converted
126
127
Figure 5.10. Ray trace plots through the I-70 model (Figure 5.7): (a) scattered P-waves from an incident P-wave, (b) scattered
mode-converted SV-waves from an incident P-wave, (c) scattered SV-waves from an incident SV-wave, and (d) scattered mode-
converted P-waves from an incident SV-wave. Mode-conversion rays that are shown occurred at reflection from horizons of
interest. Sources are located at 0 feet on the x-axis, and the receiver interval along the x-axis is 2.0 feet. See text for discussion.
reflections from this interface. The P-SV reflection mid-point lies closer to the receiver

than the P-P reflection mid-point, and the SV-P reflection mid-point lies closer to the

source (Figures 5.10b and 5.10d), because Vp > Vs for the unsaturated overburden.

Asymmetry in the converted-mode reflection raypath is described by Snell’s law (sin ș /

Vp = sin ȥ / Vs; where ș and ȥ are the P- and SV-wave angles of incidence and reflection

respectively, see Chapter 2). Increasing the Vp/Vs ratio will cause the P-SV conversion

mid-point location to move closer to the receiver, and will cause the SV-P conversion

mid-point to move closer to the source.

Due to raypath asymmetry, when converted-mode reflections can be identified in

field reflection data and targeted for imaging, a common-conversion-point (CCP) sorting

and processing approach (Figure 5.11) must be used prior to stacking. This is different

than the common-midpoint (CMP) approach that is often applied for common-mode P- or

S-wave reflection stacking. A CCP approach takes into account ray path asymmetry,

thus correcting for converted-mode reflection mid-point locations. Approximate

solutions for calculating converted-mode reflection points have been presented for

petroleum exploration purposes by Chung and Corrigan (1985), and failure to use such

sorting and processing procedures can result in a smearing of converted-wave reflection

points over a large area of the subsurface (resulting in loss of resolution and reflection

degradation). Additional factors requiring consideration during converted-wave data pre-

processing were discussed in Chapter 3, and generalized converted-wave data processing

and imaging strategies (developed for petroleum exploration purposes) can be found in

Stewart and Gaiser (1999).

128
Figure 5.11. Diagrams demonstrating common-midpoint (CMP) versus common-
conversion-point (CCP) sorting/binning approaches for reflection data. The CMP
approach (a) is often used for sorting and stacking common-mode reflections, for which
the flat interface reflection point lies at the source-to-receiver midpoint. For mode-
converted reflections, the flat interface reflection point does not lie at the source-receiver
midpoint (b), due to differences in media P- and S-wave velocities. A CCP approach
based on Vp/Vs properly sorts mode-converted reflection data prior to event moveout
correction and stacking (c). Horizontal and vertical scales are not implied.

129
5.3.3.1 Predicted Arrival Times and Moveouts

Common- and converted-mode reflection arrival times as a function of offset

from the I-70 model interfaces (Figure 5.7) are shown in Figure 5.12. Amplitude and

phase variations as a function of offset are not inferable from Figure 5.12. The same

synthetic seismic survey parameters that were used to generate the ray trace plots in

Figure 5.10 were used to generate the reflection travel time plots in Figure 5.12.

General differences in character of the common- and converted-mode reflections

in Figure 5.12 are apparent. Due to differences in Vp and Vs values for media, common-

mode P-wave reflections (Figure 5.12a) arrive earlier in time than converted-mode

reflections (Figures 5.12b and 5.12d) and common-mode SV-wave reflections

(Figure5.12c). The common-mode P-wave reflections have the least separation in time

between them, due to the fact that they have the highest average velocities along

raypaths. Due to geometry and media velocity values, the mode-converted P-SV and

SV-P events have similar arrival times and moveout velocities, but a reversed asymmetry

of raypaths. The SV-wave common-mode reflections have the most separation in time

between them, due to relatively low average velocities along raypaths.

Due to asymmetry along raypaths, mode-converted reflections are non-hyperbolic

at large offset-to-depth ratios. This means that if mode-converted reflections could be

identified in field data, then they would not be properly flattened at large offset-to-depth

ratios using conventional normal moveout (NMO) correction. Prior to stacking, NMO

correction (i.e. flattening) of P-P and S-S (common-mode) reflections (Figures 5.12a

and5.12c) has often been accomplished using a two-term truncation of the series

developed by Taner and Koehler (1969). For offset-to-depth ratios that are less than 1.0,
130
131
Figure 5.12. Arrival times of common-mode and mode-converted reflections from the I-70 model interfaces (Figure 5.7): (a) P-
P reflections, (b) P-SV reflections, (c) SV-SV reflections, and (d) SV-P reflections. Respective source types are located at 0.0
feet (trace number 1) on the x-axis. Amplitude variations with offset, and the ability to resolve/image each of the reflection
events that are plotted in this figure using field reflection measurements are not implied.
results in Larson (1996) indicate that conventional NMO-correction should also be

effective in flattening converted-mode reflections. When the offset-to-depth ratio

exceeds 1.0 however, a higher order approximation, such as that developed by Slotboom

(1990) will more effectively flatten mode-converted reflection events prior to stacking.

Further discussion regarding converted-mode reflection moveout analysis can be found in

Frasier and Winterstein (1990), and in Stewart and Gaiser (2001).

5.3.3.2 Comparison of Predicted and Observed Field Data Events

The predicted event travel-time and moveout plots (Figure 5.12) were compared

with the field multicomponent reflection data acquired in the I-70 study area. The plots

were compared to the shot records shown in Figures 5.3 and 5.5, to the rest of the shot

records processed to generate the stacked sections in Figures 5.4 and 5.6, and to the

majority of the remaining thousands of shots acquired along lines EBTravel and EBPass

(for I-70 seismic reflection survey details see Chapter 3). Comparisons were made with

various bandpass filters, f-k filters, and gains applied to the field data components, and

with the field data components sorted and displayed as various types of gathers (e.g.

common-mid-point, common-receiver, and common-source).

Common-mode P- and S-wave reflections from the top-of-saturated overburden

and the top-of-bedrock respectively, were consistently found to closely agree with

predicted common-mode event travel-times/moveouts (Figure 5.12) from these

interfaces. However, P- and S-wave common-mode events from other interfaces were

not observed in field data (Chapter 4). Converted-mode events having travel-

times/moveouts similar to those in Figure 5.12 were not observed in field reflection
132
records, which aside from the common-mode events mentioned above consisted of

(airwave, random, source-generated, surface-wave, and traffic-related) noise (Figure 5.3).

Because converted-mode reflections could not be confidently and consistently identified

in the field data, they could not be targeted for further processing and imaging, as could

be accomplished using the observed P- and S-wave common-mode events (see above).

Evaluation of the field data components acquired in the I-70 study area, along

with results from analyses (based on field-data derived information) that have been

discussed so far in this chapter, suggest that mode-converted reflections were not

observed in the field multicomponent data due to amplitude- and noise-related issues. In

order to further investigate and better understand why mode-conversion imaging could

not be effectively accomplished in the I-70 near-surface environment (and to provide

useful insight for possible future mode-converted wave imaging attempts elsewhere),

elastic-wave modeling was conducted.

5.3.4 Elastic-Wave Modeling

Numerical elastic-wave modeling can be useful for addressing seismic reflection

data acquisition, processing, amplitude and phase analysis, and interpretation problems.

For this study, such modeling was conducted in order to gain a better understanding of

why converted-mode reflections were not observed in multicomponent seismic field

records acquired in the I-70 study area. The I-70 subsurface (in un-mined areas)

represents a subsurface environment class that is typically encountered during near-

surface seismic reflection studies. The modeling approach and results presented in the

subsequent sections therefore have relevance, and can be applied for the purpose of
133
evaluating near-surface converted-wave imaging potential, prior to data acquisition in

other studies focused on characterizing similar near-surface environments elsewhere.

5.3.4.1 Forward Modeling Approach

Elastic-wave modeling and synthetic seismogram generation were accomplished

using three (Landmark Graphics Corporation) Sierra software applications: 1) MIMIC+

(geophysical model building), 2) QUIKSHOT+ (forward ray tracing), and 3) SLIPR+

(ray trace data reformatting, convolution, and SEGY output). Synthetic data that were

output as SEGY files were then reformatted and displayed using ProMAX seismic

processing software (Landmark Graphics Corporation) for digital image output.

A 2D model that incorporated I-70 subsurface media properties (Table 5.1, Figure

5.7) was built (using MIMIC+). A reflection survey array that was similar to that of the

I-70 line EBTravel field multicomponent reflection survey (Chapter 3) was then

constructed (using QUIKSHOT+).

Full elastic raytracing through the subsurface model (Figure 5.7) was conducted

(using QUIKSHOT+) in order to calculate reflection (common- and converted-mode)

travel-times and offset-dependent reflectivity. Reflection travel-times were determined

using a straight-line propagation between each model layer, following Snell’s law (see

above). The rays that were captured depended upon user-input ray instructions defined at

each layer interface. The employed raytracing algorithm assumed that energy can be

represented by plane wave fronts, and was used such that calculated amplitudes were

dependent upon interface reflection and transmission coefficients (calculated results

obtained were similar to those that would obtained using the PSHSV code in Appendix
134
B). The final amplitudes along a given raypath were a function of source-type and

amplitude, changes in the amplitude of a ray due to interface reflection and transmission

(post-critical complex reflection coefficients were considered), and geometric spreading.

Raytracing did not account for attenuation, dispersion, or absorption of energy with

distance, and events such as airwave, direct-waves, multiples, refracted-waves, source-

generated noise, surface waves, and traffic noise were not considered by the code.

Although many of these noise types contributed to converted-wave reflections not being

observed in the I-70 field records, it will be subsequently demonstrated through

modeling, that even without the presence of such coherent noise types, converted-wave

reflections from each of the I-70 subsurface interfaces would not likely be able to be used

effectively for imaging due to amplitude-, random noise-, and resolution-related issues.

A given raytracing resulted in the output of a spike seismogram file. A spike

seismogram was then reformatted, and convolved with a zero-phase Ricker wavelet to

produce a synthetic seismogram (using the SLIPR+ application). The amplitudes of a

synthetic seismogram were affected by the receiver component that was specified for

convolution (i.e. amplitude decrease resulted when the direction of particle motion was at

an angle to the receiver element axis). A geometric spreading amplitude recovery factor

was applied subsequent to convolution, and trace amplitudes were then normalized to the

maximum seismogram amplitude. Synthetic seismograms were then output, reformatted,

read into ProMAX, and displayed for digital image output. Synthetics were displayed

without filtering or AGC, with individual trace amplitudes normalized, and using the

polarity convention previously outlined in Chapter 3.

135
5.3.4.2 Pure Source-Polarization Modeling Results

Synthetic seismograms generated by elastic raytracing through the I-70

subsurface model (Figure 5.7) are shown in Figure 5.13. The synthetics in Figure 5.13

were produced with the assumption that a given (surface-located) source-type propagated

a single body wave-type (i.e. compressional or shear with inline particle motion) into the

subsurface. Ray instructions at horizon interfaces were specified such that surface-

located receivers recorded common-mode and primary converted-mode reflections.

Seismograms representing four combinations of source and receiver component

pairs are presented in Figure 5.13: ZZ component (P-P), ZX component (P-SV), XX

component (SV-SV), and XZ component (SV-P). Synthetics generated using a 250 Hz

source (a higher dominant frequency than the I-70 field reflection data) are shown in

Figure 5.13a. The synthetics in Figure 5.13b were generated using center source

frequencies that were similar to the processed common-mode dominant reflection

frequencies in I-70 field data components. For the synthetics in Figure 5.13c, random

noise (having amplitudes equal to 30 percent of average spike amplitudes) was added to

each modeled component. Examples of synthetic data having varying levels of random

noise will be subsequently presented.

5.3.4.2.1 Common-Mode Reflection Synthetics

The results and conclusions reached through elastic-wave modeling of common-

mode reflections (Figure 5.13) are similar to those that were reached through finite-

difference modeling (Chapters 4 and 6). It is seen from the 250 Hz source synthetic ZZ

component data (Figure 5.13a), that the P-P reflection from the top-of-saturated-
136
137
Figure 5.13. Synthetic seismograms generated using the model in Table 5.1 and Figure 5.7: (a) with 250 Hz sources, (b) with source
frequencies based on I-70 field data component frequency filtering, and (c) with random noise (amplitudes equal to 30 percent of average
spike amplitudes) added to each modeled component. Synthetics are displayed with traces normalized and using the polarity convention
outlined in Chapter 3. The x-axis unit of absolute offset is feet. See text for discussion. (continued)
Figure 5.13. (continued)

138
(b) Synthetic seismograms generated using the model in Table 5.1 and Figure 5.7. (continued)
Figure 5.13. (continued)

139
(c) Synthetic seismograms generated using the model in Table 5.1 and Figure 5.7.
overburden (the primary P-P reflection, see Figure 5.7) dominates the record. While the

saturated overburden and bedrock interface cannot be resolved at near offsets using this

source frequency, a single event resulting from interference of the bedrock and coal

reflections is seen to be separated in time from the primary P-P reflection. At a center

source frequency of 125 Hz (Figure 5.13b), which is similar to the dominant frequency of

primary P-P reflections in I-70 field data, the low amplitude bedrock and coal

interference event is still observed at near offsets, but only the interface between

unsaturated and saturated overburden is clearly resolved. After adding random noise to

the seismograms in Figure 5.13c, only the primary P-P reflection could be targeted for

processing and imaging. The modeling results discussed in this section agree with

analyses of ZZ component field records that were acquired in the I-70 study area (see

Chapter 4).

Synthetic XX component data that were generated at a dominant source frequency

of 250 Hz are shown in Figure 5.13a. At this source frequency, common-mode SV-

reflections from each of the model interfaces are well resolved at near and far offsets, and

therefore could be targeted for processing and imaging. At a center source frequency of

80 Hz, which is similar to the dominant frequency of the primary S-wave reflections

observed in XX component field data, the synthetic SV-wave reflection from the top-of-

bedrock is seen to dominate the record (Figure 5.13b). A relatively low amplitude S-

wave event from the top-of-saturated overburden (arising from a density contrast) is

separated in time from the top-of-bedrock SV-reflection. However, after adding random

noise (Figure 5.13c), it is seen that only the SV-wave reflection from the top-of-bedrock

could be targeted for processing and imaging. The modeling results discussed in this
140
section agree with analyses of XX component field records that were acquired in the I-70

study area (see Chapter 4).

5.3.4.2.2 Converted-Mode Reflection Synthetics

Elastic-wave modeling shows that from a theoretical standpoint, converted-mode

reflections from the I-70 model sequence (Figure 5.7) offer resolution-related benefits

(under noise-free conditions) relative to common-mode P-wave reflections. It is seen

from the 250 Hz source synthetic ZX component and XZ component data (Figure 5.13a)

that the model interfaces are clearly resolved by converted-mode (P-SV and SV-P)

reflections. These results are contrary to the ZZ component modeling results obtained

using this source frequency, which showed that only the top-of-saturated-overburden was

resolved well using P-P reflections.

The ZX and XZ component synthetic records are shown using lower dominant

source frequencies in Figure 5.13b, which are similar to the dominant frequencies of

common-mode reflections in acquired I-70 field data. Converted-mode (P-SV and SV-P)

reflections are still separated enough in time at the lower frequencies to allow the top two

model interfaces (i.e. the top-of-saturated-overburden and the top-of-bedrock) to be

resolved over a substantial range of offsets under the noise-free conditions. After adding

random noise to each of the ZX and XZ components (Figure 5.13c), the top two model

interfaces are still resolved (the top-of-bedrock events only over a limited offset range)

however, P-SV and SV-P events representing interference of the bedrock and coal

converted-mode reflections are no longer inferable (due to their low amplitude relative to

other converted-mode events).


141
Although having lower amplitudes than P-P reflections (Figure 5.8), converted-

mode (P-SV and SV-P) reflections offer relative resolution benefits under noise-free

conditions, because a portion of the converted-mode raypaths propagate at shear-

velocities, thus separating events better in time. These observations suggest that if the

noise-conditions were such that converted-mode reflections could be identified in actual

field records (which they were not in the I-70 multicomponent field data case), near-

surface characterization benefits could be realized through the processing and imaging of

converted-wave reflections.

5.3.4.3 Mixed Source-Polarization Modeling Results

Analyses of multicomponent data acquired in the I-70 study area indicated that

the employed seismic sources were not pure in a polarization sense, as in addition to

propagating the respective desired body wave-type they also generated other types of

body-waves (see above). It is generally known that when employing a seismic source

designed for P-waves for example, that a certain amount of SV-wave energy also would

be propagated into the earth. Consequently, it can be expected that in addition to

recording converted-mode P-SV wave energy, an X-oriented receiver would also record

common-mode SV-wave reflection energy that was initially propagated by the vertical

source. It has previously been shown during petroleum exploration studies that a vertical

(P-wave) vibrator may also generate enough S-wave energy to allow the direct

construction of a quality common-mode S-wave section, if multicomponent receivers are

used (Fyfe et al., 1993). The potential for constructing such images using

multicomponent receivers, is also dependent upon other factors (such as the noise
142
sensitivity of different receiver orientations), in addition to the magnitudes of body wave

energies that are generated by a source.

For the modeling results presented in this section, the seismic source propagated

equal amounts of compressional (P-wave) and inline shear (SV-wave) energy into the

earth (i.e. the seismic source was equally divided between Z and X components). These

modeling efforts allow an investigation into common- and converted-mode reflection

interference, demonstrate relative differences in amplitudes of common- and converted-

mode reflections, and allow a determination to be made regarding the relative theoretical

imaging potentials of converted-mode (SV-P and P-SV) reflections for the modeled

subsurface conditions (Figure 5.7).

5.3.4.3.1 SV-P and P-P Synthetics

Synthetic seismograms generated using a seismic source that was equally divided

between P- and SV-wave energy, and Z-oriented receivers are presented in Figure 5.14.

A spike seismogram (Figure 5.14a) shows the arrival times and moveouts of common-

mode (P-P) and converted-mode (SV-P) reflections from the subsurface model interfaces

(Figure 5.7). Results from convolving the model reflectivity sequence with a 250 Hz

source wavelet, and with a 125 Hz source wavelet are shown in Figures 5.14b and 5.14c

respectively (traces are each normalized to their maximum amplitude). At both of these

dominant source frequencies, it is seen that the primary P-P reflection from the top-of-

saturated-overburden has the highest amplitudes. The SV-P reflections from the top two

model interfaces are lower amplitude than, and are separated in time (arriving later) from

143
144
Figure 5.14. Synthetic seismograms generated using the model in Table 5.1 and Figure 5.7. Source energy was equally divided between P-
and SV-waves (Z and X components respectively), and vertical receivers (Z component) were used. The synthetic in (d) had random noise
(amplitudes equal to 30 percent of average spike amplitudes) added. Synthetics are displayed with traces normalized and using the polarity
convention outlined in Chapter 3. The x-axis unit of absolute offset is feet. See text for discussion.
the primary P-P reflection. Under noise-free conditions (as the modeling results

indicate), it should be possible to observe each of these events in field reflection records.

Adding random noise (having amplitudes equal to 30 percent of average spike

amplitudes) to the synthetic data (Figure 5.14d) demonstrates that because of their

relatively low amplitudes, it would be difficult with the presence of such random noise to

process and image the converted-mode SV-P reflections. The affects of applying varying

levels of random noise to the synthetic data are demonstrated in Figure 5.15. The

primary P-P reflection, because of its high amplitudes could be imaged over a range of

random noise levels, whereas the converted SV-P reflections are only observed at low

random noise levels. The recording of coherent noise modes (e.g. airwave, source-

generated, surface-wave, and traffic-generated) in addition to random noise in acquired

field data, further degrades low amplitude converted-mode (SV-P) reflections from the I-

70 subsurface interfaces, causing them to not be observed in the field data.

5.3.4.3.2 P-SV and SV-SV Synthetics

Shown in Figure 5.16 are synthetic seismograms that were generated using a

seismic source that was equally divided between SV- and P-wave energy, and X-oriented

receivers. A spike seismogram (Figure 5.16a) shows the arrival times and moveouts of

converted-mode (P-SV) and common-mode (SV-SV) reflections from the I-70 model

subsurface interfaces (Figure 5.7). The results from convolving the model reflectivity

sequence with a 250 Hz source wavelet, and with an 80 Hz source wavelet are shown in

Figures 5.16b and 5.16c respectively (traces are each normalized to their maximum

amplitude). At both of these dominant source frequencies, it is seen that the common-
145
146
Figure 5.15. Synthetic seismograms generated using the model in Table 5.1 and Figure 5.7. Source energy was equally divided between P-
and SV-waves (Z and X components respectively), and vertical receivers (Z component) were used. The synthetics in (b) - (f) had random
noise (amplitudes equal to a percentage of average spike amplitudes) added. Synthetics are displayed with traces normalized and using the
polarity convention outlined in Chapter 3. The x-axis unit of absolute offset is feet. See text for discussion.
147
Figure 5.16. Synthetic seismograms generated using the model in Table 5.1 and Figure 5.7. Source energy was equally divided between P-
and SV-waves (Z and X components respectively), and inline receivers (X component) were used. The synthetic in (d) had random noise
(amplitudes equal to 30 percent of average spike amplitudes) added. Synthetics are displayed with traces normalized and using the polarity
convention outlined in Chapter 3. The x-axis unit of absolute offset is feet. See text for discussion.
mode S-wave reflection from the top-of-bedrock has the highest amplitudes of all

reflections. The P-SV reflections from the top two model interfaces are separated in time

(arriving earlier) than common-mode S-wave reflection energy, but are relatively low

amplitude. Under noise-free conditions (as the modeling results indicate), it should be

possible to observe each of these events in field reflection records. However, when

random noise (having amplitudes equal to 30 percent of average spike amplitudes) is

added to the data (Figure 5.16d), only the top-of-bedrock common-mode S-wave

reflection is observed in the synthetic record. As was concluded regarding SV-P

reflections (see above), the recording of coherent noise modes in addition to random

noise in acquired field data, further degrades low amplitude P-SV reflections from the I-

70 subsurface interfaces, causing them to not be observed in the field data.

148
CHAPTER 6

SH-WAVE REFLECTION DATA ANALYSIS CONSIDERATIONS

6.1 Overview

Factors affecting the ability of SH-wave reflection measurements for allowing

near-surface interfaces and subsidence-related discontinuities to be effectively imaged are

described in this chapter. The necessity of considering resolution issues, velocity model

construction methods, and imaging approaches during SH-wave reflection data analysis

to ensure accurate subsurface interpretation is demonstrated. Numerical models and

analyses of data acquired across a section of the I-70 study area (Appendix A) roadway

where a collapse previously occurred into underground coal mine workings are presented.

A data analysis workflow is developed, that given the study area subsurface

conditions and acquired SH-wave data characteristics, provides a methodology for

delineating subsurface areas where subsidence processes have been active. The

processing and imaging flow was developed to allow an accurate subsurface

interpretation to be made, but it was also designed to be efficient so that it could be used

as a basis for the processing of an additional large volume of SH-wave reflection data

(Appendix C) acquired as part of an ongoing research program in the study area.

149
The effectiveness of developed data analysis strategies for meeting the seismic

reflection survey objectives is demonstrated, and it is shown that SH-wave reflection

surveys can be used to assess mine-subsidence problems along roadways. Field data

quality permitted vertical offsets of 3-4 feet and laterally discontinuous zones of about 20

feet in the otherwise continuous bedrock reflection to be resolved in the study area.

Processed SH-wave data indicate that there are two locations near the previous roadway

collapse where a relatively high potential for future mine-related surface failure exists.

6.2 Shear-Wave Reflection Data

The data presented in this chapter were acquired in an effort to determine whether

subsurface subsidence processes had continued subsequent to remediation, at a location

of the I-70 study area (Appendix A) roadway that had previously collapsed into

underground coal mine workings. The data were also acquired in an effort to identify

additional possible areas where a high potential for future failure exists along a 200-foot

(61 m) section of the roadway encompassing the previous surface collapse.

The locations of the SH-wave reflection (Chapter 2) lines that are analyzed in this

chapter are shown relative to the eastbound lanes of I-70, and relative to the location of

the previous collapse and underground mine workings in Figure 6.1. Data were analyzed

from a 1999 survey, during which reflection line Test-1 was positioned parallel to and 60

feet south of the southern edge of the eastbound lanes of I-70. Line GUE-I70-1 was also

acquired during the 1999 survey, and was positioned along the southern edge of the

eastbound lanes. Data from a 2001 survey were also analyzed, during which Line

EBPassYY was positioned along the north edge of the eastbound lanes of I-70. These
150
Figure 6.1. Map view of the eastbound lanes of I-70 (road stations 48300 to 48500)
showing the locations of seismic reflection lines (Test-1, GUE-I70-1, and EBPassYY)
relative to the locations of the underground mine workings. The area of previous
roadway failure is also shown, where a surface collapse feature roughly 10 ft in diameter
was centered in the travel lane at road station 48345. Road stations are given in feet from
the western county line. Due to the small map scale and possible errors that exist with
regards to room and pillar locations, the spatial relationship between these features and
the road stations is regarded as only approximately accurate.

data were acquired using sources that generated preferential shear particle motion

transverse to the line, and geophones containing horizontal elements oriented transverse

to the line. Information regarding acquisition of the data, and discussion regarding data

geometry and pre-processing are presented in Chapter 3.

6.3 Shot Gather Analysis

A consistently strong reflection event with zero-offset times ranging from 105 to

120 ms, and apparent NMO velocities ranging from 600 to 800 ft/s, was observed on shot

151
gathers acquired in the previous roadway collapse region. For line Test-1 and line GUE-

I70-1 filtered shot gathers, the average dominant reflection frequency was 80 Hz. For

line EBPassYY filtered shot gathers, the average dominant frequency was 100 Hz. Depth

estimates using velocities derived from reflections when correlated with drill logs

indicate that this strong reflection is from the overburden and bedrock boundary.

Shot gathers from lines Test-1, GUE-I70-1, and EBPassYY with computer-

generated hyperbolae fit to interpreted reflections are shown in Figure 6.2. For the line

Test-1 shot gather (Figure 6.2a), depth conversion indicates that the reflector is about 39

ft deep. Assuming the velocity and zero-offset time picks used for this calculation were

no more than 0.3 percent and 2 ms in error respectively (reasonable ranges of error in

picking using this approach), this calculated depth would be within plus or minus 2 feet

of the actual reflector depth. A borehole logged near this line (25 feet north of line Test-1

at road station 48350) indicates that bedrock is at a depth of 38 ft.

A reflection correlating to the overburden and bedrock interface was consistently

identified in the majority of shot gathers from the three reflection lines. Relatively lower

amplitude reflections above this interface were also evident, but on a much smaller

percentage of shots, and generally exhibited moveout similar to that of the local top-of-

bedrock reflection. Such events correlated as unconsolidated sandy units within the

overburden that were mapped from drill logs.

6.3.1 Sources of Noise

Love waves (surface waves) require a low velocity near surface layer to exist, and

are a concern during shallow S-wave surveying as they can make reflection imaging

152
Figure 6.2. Examples of shot gathers (uninterpreted and interpreted) with zero-phase
Ormsby bandpass filters (< 12 dB/octave slopes) and AGC gain (100 ms window)
applied: (a) line Test-1 gather (source located at road station 48382.5, eastbound direction
is to the left) with 80-180 Hz filter and gain, (b) line GUE-I70-1 gather (source located at
road station 48363.5, eastbound direction is to the right) with 100-180 Hz filter and gain,
and (c) line EBPassYY gather (source located at road station 48464, eastbound direction
is to the right) with 80-160 Hz filter and gain. Road stations are given in feet from the
western county line. Hyperbolic reflection events interpreted on shot gathers correlate as
the top-of-bedrock. The x-axes scales of absolute offset from the source are in feet.

153
difficult or impossible (Deidda and Ranieri, 2001; Miller et al., 2001). Background

information regarding Love wave propagation and other noise sources is provided in

Chapter 4. Because the S-wave velocity structure across this portion of the study area

was such that near-surface road-construction materials (asphalt and road-fill) had a higher

velocity than the underlying overburden materials at most locations, detrimental Love

wave noise was not observed on the majority of the shot gathers. On a small percentage

of the shot gathers (predominantly line Test-1 gathers), Love wave noise with a relatively

low apparent velocity was observed to interfere with reflection energy from the bedrock

boundary. Such Love wave energy did not require the application of surface wave

attenuation processes (it could be largely suppressed by appropriate post-NMO stretch

muting), and did not degrade the quality of imaged reflections or lead to data

misinterpretations.

Lines Test-1 and EBPassYY were acquired with the source baseplate located on

relatively soft soil, whereas line GUE-I70-1 was acquired with the source baseplate

positioned on the relatively hard roadway pavement. As shown in Figure 6.3, line GUE-

I70-1 correlated shot gathers contained high amplitude periodic noise (at an average

dominant frequency of 65 Hz) that resulted from decoupling of the source baseplate.

Frequencies were introduced through baseplate decoupling within the frequency range of

reflection signal, and source-related noise could therefore not be entirely suppressed

through frequency filtering without degrading reflection quality. As a result, line GUE-

I70-1 data had a lower signal to noise ratio than the line Test-1 and line EBPassYY data

after optimum frequency filtering (after suppressing as much noise as possible without

degrading reflection signal quality).

154
Figure 6.3. Line GUE-I70-1 shot gather (source located at road station 48361.5,
eastbound direction is to the right) with various zero-phase Ormsby bandpass filters (<
12 dB/octave slopes) and AGC gain (100 ms window) applied: (a) unprocessed, (b)
unfiltered with gain, (c) 40-100 Hz filter with gain, (d) 80-140 Hz filter with gain, (e)
120-180 Hz filter with gain, and (f) 160-220 Hz filter with gain. High amplitude periodic
noise evident in shot gathers (predominantly in the range of 50-80 Hz) is related to source
decoupling. The x-axis scale of absolute offset from the source is in feet.

155
Because Interstate traffic was heavy and unable to be stopped during data

acquisition, noise from roadway vehicles was frequently observed on shot gathers. An

example of noise recorded from a passing roadway vehicle is shown in Figure 6.4.

Recorded traffic noise was predominantly low frequency (5-25 Hz, see Figure 6.4a), and

this characteristic of traffic noise has been observed in other shallow S-wave reflection

studies conducted near roadways (Zhang, 1990). Traffic noise was significantly

suppressed through the application of a low-cut frequency filter (Figure 6.4b).

6.3.2 Overburden and Bedrock Interface Reflection Coefficients

The consistently high amplitude of the overburden and bedrock reflection event

observed in the previous roadway collapse region indicates that a high impedance

contrast exists at this interface. For a case where the overburden and bedrock boundary is

a plane that is parallel to the polarization of an incident SH-wave, only scattered SH-

waves would be produced from an incident SH-wave (no mode conversion would occur),

regardless of the incident angle (Chapter 2). Analysis of the shot gather shown in Figure

6.5a, indicates a dominant reflection frequency of about 80 Hz (Figure 6.5b), and

approximate overburden and bedrock S-wave velocities at this location of 670 ft/sec

(apparent NMO velocity) and 2500 ft/sec (unreversed refraction velocity) respectively

(Figure 6.5c). Depth conversion and drill log correlation indicates that the interpreted

reflector is the overburden and bedrock interface. Using the velocity and depth values

derived from the interpreted shot gather in Figure 6.5c, an expected critical distance (the

offset at which the reflection time equals the refraction time) of about 22 ft from the

source is calculated, which agrees with the critical distance observed in the field record.
156
Figure 6.4. Line EBPassYY shot gather (source located at road station 48464, eastbound
direction is to the right) and amplitude spectrum: (a) unprocessed, and (b) with zero-
phase Ormsby bandpass filter (80-160 Hz, < 12 dB/octave slopes) applied. Noise from
roadway traffic (indicated in the circled region) is predominantly low frequency (e.g. 5-
25 Hz), but is also evident in the optimum frequency range for imaging reflection signal.
The x-axis scale of absolute offset from the source is in feet.

157
Figure 6.5. Line Test-1 shot gather (source located at road station 48397.5, eastbound
direction to the left): (a) processed and uninterpreted, (b) amplitude spectrum indicating a
dominant frequency of about 80 Hz, and (c) interpreted. An apparent NMO velocity of
670 ft/sec indicates the hyperbolic reflection is from the top of bedrock at a depth of 39
ft. An apparent refraction velocity (unreversed) of 2500 ft/sec is also interpreted. The x-
axis scale on the shot gathers of absolute offset from the source is in feet.
158
The critical distance (xc) is equivalent to:

xc = 2z*tan(șc) (6.1)

where z is the reflector depth, and șc is the critical angle. When the velocity of the

incident medium is less than that of the refracted / transmitted medium, the refraction

angle is equivalent to 90 degrees at the critical angle (șc), which is defined as:

șc = sin-1(Vs1 / Vs2) (6.2)

where Vs1 and Vs2 are the incident and refracted media shear velocities.

Assuming these data-derived velocity values are constant for the overburden and

bedrock media, and assuming a density ratio of unity, square root energy coefficients (for

equations see Chapter 2) were calculated as a function of incidence angle for a SH-wave

incident on the overburden and bedrock interface, and are plotted in Figure 6.6. As seen

in Figure 6.6, the reflection coefficient is 0.5773 at normal incidence (the square root

energy coefficient curve for the reflected SH-wave is equivalent to the magnitude of the

displacement amplitude or the “reflection coefficient”). When compared with typical

normal incidence reflection coefficient values observed during petroleum exploration

studies, the calculated coefficient value indicates that the overburden and bedrock

boundary at this location can be considered as a high impedance contrast. Reflection

coefficient values for geologic materials are generally classified by Reynolds (1997) as: <

0.1 (weak reflection), 0.1 - 0.2 (moderate reflection), and > 0.2 (strong reflection),

whereas Yilmaz (2001) considers a reflection coefficient > 0.3 as high.

The reflected- and refracted-wave curves change only slightly for near-vertical

values of incidence (Figure 6.6). As the incidence angle increases past 10 degrees

however, the reflected wave curve rapidly approaches zero and the refracted-wave

159
Figure 6.6. Plot showing normalized square root energy coefficients as a function of
incidence angle for a SH-wave incident on the overburden (medium 1) and bedrock
(medium 2) interface. Calculations were made using the velocities determined from
analysis of the shot gather in Figure 6.5c, with the computer program presented in
Appendix B. Note the critical angle (șc) at about 15.5 degrees.

160
curve begins to increase significantly. As the incidence angle increases towards the

critical angle, refracted rays become more horizontal, the surface area between refracted

rays becomes smaller, and the refracted-wave amplitude therefore becomes large. The

refracted-wave curve increases significantly prior to a critical angle due to geometric

reasons (Lay and Wallace, 1995). The amplitude of a wave is proportional to the square

root of the energy per surface area (the refracted wave amplitude is therefore inversely

proportional to the surface area between refracted rays). At the critical angle the

refracted ray is horizontal, and for angles of incidence greater than the critical angle no

real SH-wave energy is transmitted into the bedrock medium normal to the interface.

6.4 Target Resolution Potential

The ability to distinguish two separate features and observe detail using reflection

data is related to resolution. Both vertical and lateral (horizontal) resolution must be

considered in order to assess the potential of data for allowing reflecting horizons to be

detected and resolved, reflectors and diffractors to be differentiated, and discontinuities in

reflecting horizons to be inferred. Lines Test-1, GUE-I70-1, and EBPassYY were

acquired with high CDP (common-depth-point, i.e. the midpoint between a source and

receiver) sampling intervals (0.5 ft, 0.5 ft, and 1 ft respectively) relative to the sampling

intervals used for the majority of near-surface reflection surveys that are conducted.

6.4.1 Reflecting Horizon Resolution Potential

Synthetic shot gathers were generated for comparison with field data and to

investigate the resolution potential of field data relative to the study area geology. The

161
synthetic seismograms were produced using a finite-difference (Alford et al., 1974; Kelly

et al., 1976) modeling code within the ProMAX software package. This code was written

for the purpose of modeling P-wave reflections in acoustic media, but because it does not

consider media elastic parameters or mode conversions, it is also suitable for modeling

SH-wave reflections. Polarity and intromission angle occurrence is different for P-waves

in acoustic media and SH-waves in elastic media, see Chapters 2 and 7, but this does not

affect reflection travel-times, or the objectives of the modeling that was conducted.

Direct arrivals, multiples, and refractions are modeled using the code, but surface

waves are not. A zero-phase Ricker wavelet was used as the source for each model that

was generated, and the source was located at the surface of each model. Reflecting

boundary conditions were specified, and it was computationally feasible to extend the

model bottom to a depth such that bottom reflections were not recorded during the time

range of interest. Grid dispersion was minimized by specifying at least 7 grid points per

wavelength in each region (regardless of velocity) of the models (e.g., at a velocity of 670

ft/s and a frequency of 100 Hz, the grid spacing would be at least 0.96 feet).

The S-wave interval velocity model in Table 6.1 was used to generate synthetic

seismograms, with geometries equivalent to the shot gather in Figure 6.5. Velocity and

depth estimates in Table 6.1 were obtained through analysis of the shot gather shown in

Figure 6.5. Velocity and depth approximations are also included for a 7 ft thick coal

seam, that according to drill log information exists beneath 20.5 ft of bedrock near the

location of the shot gather in Figure 6.5. Reflections correlating to the top or bottom of

the coal seam were not observed in field records and, therefore, the coal’s shear velocity

could not be directly measured. Instead, to model the Upper Freeport Coal the S-wave

162
S-wave interval velocity (ft/sec) Lithology Layer thickness (ft)
670 Overburden 39
2500 Bedrock (shale) 20.5
2395 Coal (bituminous) 7
2500 Bedrock (shale) 150

Table 6.1. Velocity model used to generate synthetic data shown in Figures 6.7 and 6.8.

velocity of 2395 ft/s from the Lower Freeport Coal (Wolfe et al., 1989) was used. The

Lower Freeport coal (middle Pennsylvanian series) is a bituminous coal that is located

stratigraphically beneath the study area Upper Freeport Coal.

Uninterpreted seismograms are shown in Figure 6.7a (300 Hz center frequency

source wavelet) and Figure 6.7c (100 Hz center frequency source wavelet), with

interpretations of these seismograms shown in Figure 6.7b and Figure 6.7d respectively.

When a 300 Hz source wavelet is used, a dominant peak-to-peak reflection frequency of

250 Hz resulted, and reflections from the three model impedance contrasts were

distinguishable at near source to receiver offsets, with a phase reversal evident at these

offsets for the bedrock and coal interface reflection (Figure 6.7b). Reflections from the

coal top and bottom are low amplitude relative to the overburden-bedrock event, but the

reflected wavelets from the top of the bedrock and coal layers (coal layer reflections are

of opposite polarity to one another) are seen to be separated in time because the thickness

of the model coal (7 ft) is greater than half of the wavelength of the seismic wavelet in

the coal (4.8 ft).

163
Figure 6.7. Uninterpreted synthetic seismograms generated using the model in Table 6.1
with a center source frequency of 300 Hz (a), and 100 Hz (c). The three model interfaces
are resolved at near and far offsets with a source frequency of 300 Hz (b), while only the
primary event is easily interpreted with a source frequency of 100 Hz (d). Also shown is
the uninterpreted (e) and interpreted (f) shot gather (from Figure 6.5) used as a basis for
forward modeling. A high reflection coefficient at the overburden-bedrock interface,
near offset noise, a lower signal to noise ratio, interference, and wavelet ringiness result
in only the overburden and bedrock reflection being interpretable from the field data.
The x-axis scale of absolute offset from the source is in feet.
164
The resolution potential of data is strongly dependent on frequency and velocity

(Widess, 1973), however numerous additional factors also influence the resolving

potential of shallow reflection data (Miller et al., 1995). It is apparent from Figure 6.7b

that data with a dominant reflection frequency of 250 Hz should provide the potential to

resolve the top and bottom of a coal seam in this geologic sequence. The bedrock

reflection and refraction interpretations in Figure 6.7b are the same (the same velocities

and zero-offset times) as those that were made for the data in Figure 6.5c.

As a result of the relatively low overburden velocity and the high overburden and

bedrock velocity contrast, the primary reflection event corresponding to this interface in

Figure 6.7b is distorted over a certain range of offsets where deeper reflections from the

coal top and bottom cross this event. Despite interference, the shape and amplitude of the

composite wavelet over this range of offsets is still dominated by the character of the

overburden and bedrock reflection. The primary reflection arrives later in time at

increased offsets than the coal reflections, despite the raypath distances of the coal

reflections being greater than that of the primary reflection. Beyond the offsets where

interference from crossing reflections occurs, the three reflection events are imaged as

separate events, and are also distinguishable from the bedrock refraction.

An interpreted synthetic seismogram generated using the velocity model in Table

6.1 and a dominant source frequency of 100 Hz (resulting in a dominant peak-to-peak

reflection frequency of 80 Hz) is shown in Figure 6.7d. For comparison purposes, the

same interpretations that were given in Figure 6.7b are superimposed on the data in

Figure 6.7d. The high amplitude primary reflection is clearly observed and interpretable

at near source to receiver offsets. Interference of reflection energy does not allow lower

165
amplitude events from the coal top and bottom to be as easily interpreted however, and a

phase reversal from the bedrock and coal interface is not clearly evident.

At near offsets the opposite polarity events from the bedrock and coal tops

interfere (the composite wavelet is dominated by the primary reflection), because the coal

thickness (7 ft) is less than half of the dominant wavelength in the coal (15 ft).

Interference of the primary reflection and lower amplitude crossing events associated

with the coal seam occurs across a larger range of offsets than in Figure 6.7b.

Interference effects have no evident effect on the composite wavelet across this offset

range though, which is still dominated by the character of the primary reflection. It is

therefore unlikely that lateral changes in material properties beneath the overburden-

bedrock interface could be inferred using interference observations and / or amplitude-

based criteria at the dominant field data frequency. At far offsets where reflection energy

from the coal arrives earlier in time than the primary reflection energy, coal top and

bottom reflections cannot be interpreted from the synthetic seismogram.

The shot gather shown in Figure 6.5 is again shown in Figure 6.7e (uninterpreted)

and Figure 6.7f (interpreted). The dominant frequency of the field data and the synthetic

seismogram generated using a 100 Hz source wavelet (Figure 6.7d) are comparable, and

interpretations of the overburden and bedrock interface reflection and refraction events

are the same for the synthetic data and the field data. The observed fit of the same

reflection hyperbola and the same linear refraction event to the field and synthetic data

supports the initial interpretation of these events in the field data and their apparent

velocities. Evidence suggesting the presence of a coal seam beneath the overburden and

bedrock interface does not exist in the field data, likely as a result of the high reflection

166
coefficient at the overburden-bedrock interface, the lower signal to noise ratio of field

data, interference effects, wavelet ringiness (additional wavelet cycles), and the source-

related noise present at near source to receiver offsets.

The specification of different coal S-wave velocities in the model shown in Table

6.1 did not change the main conclusions obtained from forward modeling. Shown in

Figure 6.8 are uninterpreted (left) and interpreted (right) synthetic seismograms generated

using a 100 Hz source wavelet, with varying S-wave velocities (1500 ft/s, 2000 ft/s, and

3000 ft/s) specified for the coal seam. In each of the synthetics shown in Figure 6.8, a

primary reflection from the overburden-bedrock interface dominates the record, and

interference of reflection energy does not allow lower amplitude events from the coal top

or bottom to be easily interpreted.

Based on the modeling in this section, general conclusions are made as follows:

1) The coal seam in the study area subsurface cannot be imaged using

acquired SH-wave reflection data due to: a high reflection coefficient at

the overburden and bedrock interface, inadequate signal-to-noise ratio

of field data, source-related noise, wavelet ringiness, interference and

poor resolution.

2) Lateral changes in material properties beneath the overburden and

bedrock interface cannot be inferred using field data interference

observations and / or amplitude-based criteria. Interference effects from

deeper reflections have no evident effect on the top-of-bedrock

reflection event at the dominant frequency of acquired field data.

167
Figure 6.8. Uninterpreted (left), and with the top of bedrock reflection and refraction
interpreted (right) synthetic seismograms (center source frequency = 100 Hz) generated
using different S-wave velocities for the coal seam specified than those in Table 6.1.
Interference of reflection energy does not allow lower amplitude events from the coal top
or bottom to be easily interpreted at any of these modeled coal S-wave velocities.

168
6.4.2 Discontinuity Resolution Potential

In addition to affecting the potential for imaging reflecting horizons, vertical

resolution also affects the potential for inferring discontinuities along reflecting horizons.

A generally accepted threshold (Sheriff and Geldart, 1982; Yilmaz, 2001) used to

estimate the vertical resolution of reflection data is a quarter of the dominant wavelength.

A similar threshold for easily inferring vertical offset along a reflecting horizon is that the

offset must be at least equivalent to a quarter of the dominant wavelength (Yilmaz, 2001).

A synthetic stacked section that was generated using the overburden and bedrock

velocity and depth values determined from the field data in Figure 6.7f as a basis, is

shown in Figure 6.9. Five faults (associated with mine subsidence in this example)

across an interface separating overburden and bedrock materials were modeled, with the

amount of vertical offset for each fault specified as a fraction of the dominant

wavelength. For the modeled center frequency of 80 Hz, the corresponding dominant

wavelength in the overburden is 8.4 ft.

Vertical offset of the bedrock interface must be at least a quarter of the dominant

wavelength to be easily inferred (Figure 6.9). In field data with a high signal to noise

ratio, inference of offset less than a quarter of the dominant wavelength can be possible

using diffractions. The modeled overburden and bedrock boundary was 39 ft deep at

CDP 100 (the modeled CDP spacing was 0.5 ft), and due to continued downward offset

along faults this boundary existed at a depth of 55.3 ft at the location of CDP 600.

Lateral resolution can be addressed using the concept of Fresnel zones. The first

Fresnel zone (also known as the half wavelength Fresnel zone) is defined as the portion

of a reflecting surface from which energy arrives at a receiver within a half cycle after the

169
Figure 6.9. A synthetic section generated using the overburden and bedrock velocity and
depth parameters determined from the field data in Figure 6.7f as a basis. Five faults
(fault locations are indicated by arrows on the x-axis) were modeled, with the amount of
vertical offset for each fault specified as a fraction of the dominant wavelength (Ȝ). As
seen by the modeled results, vertical offset of the bedrock interface must be at least a
quarter of the dominant wavelength to be easily inferred without relying on diffraction
events (which may not be observed in field data with a lower signal to noise ratio). The
overburden and bedrock boundary is 39 feet deep at CDP 100, and due to continued
downward offset this interface is at a depth of 55.3 ft at CDP 600 (CDP spacing is 0.5 ft).

170
onset of reflection from the Fermat reflecting point (Waters, 1987). Energy reflected

from this zone (containing all points within a certain radius from the Fermat point)

interferes constructively, and thus contributes positively to the reflected amplitude.

Energy reflected from a second Fresnel zone (surrounding the first Fresnel zone) arrives

at a receiver between a half to one cycle later than the onset of reflection, and interferes

destructively. A pattern of alternating zones of constructive and destructive interference

continues outward away from the Fermat point, although, the major contributions to the

reflected signal come from the first Fresnel zone (Sheriff and Geldart, 1982). The

diameter (d) of the Fresnel zone (the adjective “first” is typically dropped) is defined as:

d = 2((V / 2)(t / f)1/2) (6.3)

where V is the average velocity above the reflector, t is the arrival time (two-way travel

time), and f is frequency.

The Fresnel zone diameter can be used to estimate lateral resolution (a smaller

diameter corresponds to higher lateral resolution), and therefore to estimate the potential

of data to allow reflectors and diffractors to be differentiated. Using the field data

parameters in Figure 6.7f, the diameter of the Fresnel zone for the bedrock interface (at a

depth of 39 feet) for a dominant wavelength of 8.4 ft is calculated to be 25.6 ft.

Assuming a higher dominant reflection frequency of 100 Hz, the Fresnel zone diameter

for this interface would be 22.9 ft. For an increasing overburden thickness or overburden

velocity, or a decreasing frequency, the diameter of the Fresnel zone would increase.

For a situation where an otherwise continuous bedrock horizon has collapsed

between stable coal pillars into a mine room, diffractions (with twice the moveout of

reflections from the horizon) would occur at the horizon edges. If the width of the down-

171
dropped bedrock segment were much smaller than width of the first Fresnel zone, it

would be beyond the data lateral resolution. The horizon in this case would appear to be

continuous across the non-reflecting segment on a seismic section, and the non-reflecting

segment would not likely be inferred from field data unless the signal to noise ratio was

high enough to allow diffraction events to be observed. This concept is demonstrated in

Figure 6.10, which shows an unmigrated synthetic stacked section generated at a

dominant frequency of 80 Hz. The overburden and bedrock velocity and depth

parameters determined from the field data in Figure 6.7f were used as a basis for the

synthetic section shown in Figure 6.10. Four graben structures (each resulting from a

segment of the bedrock horizon and overburden being down-dropped by 7 feet into a

mined coal room) were modeled, with the width of each feature specified as a fraction of

the Fresnel zone diameter (d). The top-of-bedrock reflection appears to be continuous

across the modeled subsidence feature when the spatial extent of the feature is much

smaller than the Fresnel zone diameter (Figure 6.10).

Based on the modeling in this section, general conclusions are made as follows:

1) Vertical offset of the bedrock interface (along mine subsidence-related

normal faults) must be at least a quarter of the dominant wavelength to

be easily inferred using field data.

2) Reflections will appear to be continuous across a graben feature

resulting from bedrock subsidence into a mine room when the feature’s

spatial extent is much smaller than the size of the Fresnel zone diameter.

172
Figure 6.10. A synthetic stacked section generated using the overburden and bedrock
velocity and depth parameters determined from the field data in Figure 6.7f as a basis.
Four bedrock graben structures resulting from mine-related subsidence activity (graben
locations are indicated by arrows on the x-axis) were modeled, with the width of each
structure specified as a fraction of the Fresnel zone diameter (d). As seen by the modeled
results, reflections appear to be continuous across a bedrock subsidence feature when the
spatial extent of the feature is much smaller than the size of the Fresnel zone. The top of
bedrock exists at a depth of 39 feet at all CDP locations, except where this boundary has
subsided to a depth of 46 ft in the grabens (CDP spacing is 0.5 ft).
173
6.5 Data Processing

Data processing and imaging operations were performed using ProMAX software

(Landmark Graphics Corporation). The processing flow applied to the each of the lines

was similar, efficient, and effective in allowing the survey objectives to be accomplished.

The processing flow applied to the data in this chapter is shown in Table 6.2. Discussion

regarding data pre-processing and geometry definition was presented in Chapter 3.

6.5.1 Processing Strategy

There were two objectives during the selection of processes and process

parameters for the shear-wave reflection data acquired in the previous collapse region.

One objective was to image the subsurface so that possible areas where subsidence

processes have been active could be accurately delineated. A second objective was to

design the processing flow so that it could be used as a basis for, and applied towards the

analysis of additional SH-wave reflection data acquired in the study area.

Because an event correlating to the overburden and bedrock interface could

consistently be identified in field records, data processing and imaging operations

predominately targeted this high impedance contrast. The resolution potential of the data

(see above) suggested that it should be possible to infer bedrock horizon offsets on the

order of several feet, and disrupted or discontinuous areas along the horizon with lateral

extent less than that of the width of mined-out rooms between coal pillars (Figure 6.1).

Focusing on the character of the bedrock horizon in shot gathers and processed images

seemed to be a reasonable approach for identifying possible subsurface areas where

174
Processing step Description
Data reformat From SEG-Y to ProMAX format
Line Test-1 correlated with synthetic sweep, line GUE-I70-1
Vibroseis correlation correlated with AUX channel –2, line EBPassYY correlated
with pilot sweep
Geometry Defined using field notes and loaded to headers
Data truncation Records truncated to 300 ms
Trace editing Bad / noisy traces killed
Trace equalization 150 ms spatially varying window
CMP sort Sorted from shot gathers to midpoint gathers
Integrated analysis of shot gathers, constant velocity stacks,
Velocity analysis
and semblance plots
NMO correction Applied based on optimum stacking velocities
Stretch mute 30 percent
Zero-phase Ormsby filter: 40-80-180-220 Hz for line Test-1,
Bandpass filter 60-100-180-220 Hz for line GUE-I70-1, and 50-80-160-200
Hz for line EBPassYY
AGC scaling 100 ms window
CMP ensemble / stack Summed NMO-corrected CMP gathers

Table 6.2. Data processing flow for shear-wave reflection data.

175
mine-related subsidence processes have been active, because in order for a mine-related

collapse feature to propagate up to the surface, the bedrock horizon overlying the coal

mine must first be disrupted.

6.5.2 Processing Flow

Trace editing and bandpass filtering were initially conducted to improve the data

quality and to suppress noise outside the optimum signal frequency range. A scalar trace-

to-trace amplitude balancing function was also calculated and applied to the shot gathers

of each line (trace equalization) in order allow a statistically optimum stack to be

produced (Hatton et al., 1986). By calculating this balancing function using CMP gathers

(150 ms spatially varying window), differences in sample amplitudes due to differences

in acquisition equipment performance and source and receiver coupling with location

were minimized (i.e. the negative effect of occasional high amplitude noise would be

reduced, and traces within each gather would contribute equally to stacked sections).

Elevation static corrections did not need to be performed because the data were

acquired on a flat, horizontal surface (less than 1 foot change in elevation per 100 lateral

feet). In many of the field records direct arrival and refraction static shifts were observed

to be inconsistent with reflection static shifts, indicating that for these records reflections

would be incorrectly shifted by applying corrections based on earlier arriving events. For

example, in the EBPassYY shot gather shown in Figure 6.2c, the observed refraction

static delay between offsets of 80 and 100 ft (at 135-140 ms) differs significantly from

that of the later arriving reflection in this offset range. The shooting geometries were

such for lines Test-1 and GUE-I70-1 that the necessary offsets for refraction static

176
calculations were not recorded for most field records (for example, see Figure 6.2b). The

application of surface-consistent residual static corrections (plus or minus 2 ms maximum

shift) did not lead to significant improvements in stacked signal quality or the continuity

of reflections, and did not effect structural interpretations based on the data.

The effectiveness of refracted arrival muting was evaluated, and as expected the

inclusion of top muting into the processing flow significantly increased the required

amount of processing time. As previously mentioned, the line Test-1 and GUE-I70-1

recording geometries were such that for many of the field records, long offsets containing

high amplitude refraction energy from the bedrock interface were not recorded. For

records containing refraction energy, it was found that top muting without degrading the

quality of reflection signal was difficult. For these reasons, and because muting did not

significantly enhance constructed images, refraction mutes were not applied to these data.

A major concern when top muting is not conducted is that non-muted refraction

energy can appear as high amplitude (and often relatively low frequency) coherent events

on processed images (Steeples and Miller, 1998). Care was taken to ensure that structural

interpretations of the subsurface based on constructed images were not affected from top

muting not being conducted. As is demonstrated using synthetic data in Figure 6.11, and

as will be shown in a subsequent section using field data, most refraction energy was able

to be eliminated during stretch muting, due to the fact that refraction arrivals undergo

greater stretching than reflections during NMO correction. As further evidence that

stacked refraction energy did not result in data misinterpretation, the zero-offset arrival

times of reflections on the shot and CMP supergathers presented agree with those

interpreted events on the processed images presented.

177
178
Figure 6.11. Synthetic seismograms (100 Hz center source frequency) that demonstrate the effectiveness of a 30 percent post-
NMO stretch mute in eliminating most refracted energy. The models shown both consist of a 40 ft thick overburden above
bedrock (bedrock velocity = 2500 ft/sec). In example 1 (a) the overburden velocity is 600 ft/sec, and in example 2 (b) the
overburden velocity is 800 ft/sec. The x-axis scale on the gathers of absolute offset from the source is in feet.
6.5.2.1 Velocity Analysis

Velocity analysis can be one of the most time-consuming aspects of processing,

and given the large amount of data acquired in the study area it was necessary to

determine an efficient procedure for obtaining a subsurface velocity model for common-

mode reflection data. It was determined from the analysis of shot gather events

correlating to the overburden and bedrock boundary that the apparent average overburden

velocity varied along lines in the previous collapse region by as much as 20 percent.

Relatively lower amplitude events that were present on certain shot gathers correlated as

unconsolidated sandy units within the overburden, and exhibited moveout similar to that

of the local overburden and bedrock event. A single layer overburden velocity model

with laterally varying velocity above bedrock was therefore determined to be appropriate

for each line for image construction purposes. The procedure for determining velocity

models for the lines acquired in the previous roadway collapse region involved an

integrated analysis of shot gathers, constant velocity stacks, and semblance plots.

Shown in Figure 6.12 are constant velocity stacks of the EBPassYY data with

stacking velocities ranging from 700 to 800 feet per second. Differences in stacked

signal quality and apparent continuity of events are evident at different stacking velocities

(see the circled region on the stacks in Figure6.12). While an event at approximately 110

ms stacks coherently using a velocity of 800 ft/sec at CDP location 48320, an event at

this time at CDP location 48480 (160 feet east of CDP location 48320) stacks coherently

using a velocity of 700 ft/sec.

Shown in Figure 6.13a is the velocity spectrum and velocity pick for a line

EBPassYY supergather (also shown) centered at road station 48415 (corresponding to

179
Figure 6.12. Constant velocity stacks generated for line EBPassYY using a velocity
increment of 25 feet per second (stacking velocities from 700-800 feet per second are
shown). Note the change in event coherency within the circled region at different
stacking velocities. CDP location numbers (CDP_X) correspond to road stations.
180
Figure 6.13. Velocity spectrum and pick for line EBPassYY CDP supergather: (a) before
NMO correction, (b) after NMO correction showing significant stretch at large offsets,
and (c) after stretch mute application. The supergather is centered at road station 48415.
181
CDP_X 48415 in Figure 6.12). The velocity spectrum was derived from the CDP

supergather and represents a measure of coherency, or how well a given hyperbolic

trajectory (dependent on time and velocity) fits the supergather data. Figure 6.13b shows

the Figure 6.13a supergather after the application of an NMO correction based on the

velocity pick (about 700 ft/sec). The hyperbolic reflection event with a zero-offset time

of 110 ms is flattened as a result of this correction, as is a lower amplitude event with a

zero-offset time of about 90 ms, which correlates as a sandy unit above the bedrock

interface. The application of a 30 percent stretch mute (Figure 6.13c) effectively

minimized reflection and refraction energy distorted at large supergather offsets by the

NMO correction. Also shown in Figure 6.13c is a dynamic stack function for the NMO

corrected and stretch muted supergather.

Shown in Figure 6.14 are velocity spectra and picks for 3 additional line

EBPassYY CDP supergathers (shown before and after NMO correction). The centered

locations of these supergathers correspond to road stations 48470 (Figure 6.14a), 48410

(Figure 6.14b), and 48320 (Figure 6.14c). These velocity spectra and picks indicate that

the optimum stacking velocity for the overburden and bedrock event (zero-offset time of

about 110 ms in these supergathers) ranges along this line from about 700 ft/sec at road

station 48470, to about 660 ft/sec at road station 48410, and to about 790 ft/sec at road

station 48320. A relatively lower amplitude event also stacks at these velocities on each

of the supergathers at a zero-offset time of about 90 ms, and the depth of this event

correlates as a sandy unit above the bedrock horizon. The semblance-based velocity

model for each of the reflection lines was built by incorporating only the highest quality

velocity picks, such as the ones shown in Figures 6.13 and 6.14.

182
Figure 6.14. Velocity spectra and picks for line EBPassYY CDP supergathers (before
and after NMO correction and stretch mute application) demonstrating lateral variation in
optimum stacking velocity along this line. The supergathers are centered at: (a) road
station 48470, (b) road station 48410, and (c) road station 48320.
183
6.5.3 CMP Stacking Versus Common Offset Imaging Approaches

Based on aspects of the data and the study area subsurface, it was determined that

the study objectives could be effectively addressed through the combined analysis and

interpretation of individual data gathers and CMP stacked sections. Since the associated

acquisition and processing costs have become affordable for a wide range of near surface

problems, CMP stacking of shallow reflection data has become a common imaging

approach. Because there are potential advantages of the common offset method over

stacking however, the potential effectiveness of this imaging approach for allowing the

study objectives to be met was considered.

CMP stacking involves summing traces that have different source to receiver

offsets, but geometrically correspond to the same subsurface midpoint. The main

advantage of CMP stacking is that the data signal-to-noise ratio can be improved.

Common offset imaging typically involves identifying an optimum offset range within

which reflection energy of interest can consistently be observed with minimal noise and

interference from other events, and then constructing an image using only the traces from

this offset range. There are other potential advantages of the common offset imaging

approach, but the main advantage that is typically experienced is a decrease in acquisition

and processing costs relative CMP stacking. Images of the subsurface produced by CMP

stacking require NMO corrections to be made, whereas common offset images do not.

In areas where severe distortions in reflection curvature exist, or when the

objective is to produce a detailed image of sub-spread length features on a rough horizon

or an image of a horizon with severe relief (Hunter et al., 1984), the common offset

approach may be more applicable than stacking, and may allow more detailed

184
interpretations to be made. A certain amount of frequency distortion and smoothing of

horizon features likely occurred during the NMO correction and stacking of field gathers

that exhibited slight distortions in moveout. However, detailed information regarding

localized horizon structure could still be inferred by paying attention to the character of

reflection events in non-corrected gathers while interpreting stacked sections.

The optimum window for imaging events of interest varied for the acquired data.

For example, the optimum window based upon the shot gather in Figure 6.2b would

include the near normal incidence traces, while this offset range would not likely be

included in the optimum window based upon the gathers shown in Figures 6.2a and 6.2c.

By stacking all offsets of the data (data were of high fold relative reflection data acquired

for the majority of shallow studies), it was possible to improve the signal-to-noise-ratio

and most effectively image the overburden and bedrock interface (Figure 6.15). This was

important for attenuating non-reflection energy and allowing the bedrock horizon and

weaker reflectors within the overburden to be accurately imaged and distinguishable from

non-reflection energy. CMP stacking was determined to provide a better basis for

allowing the study objectives to be met than common offset imaging.

6.6 Results and Interpretations

The strong reflection event correlating to the overburden and bedrock boundary in

shot and CDP gathers is interpretable across stacked sections for each of the three lines.

Lower amplitude and less continuous reflectors above the bedrock horizon are evident on

stacked sections also, and correlate to sandy units mapped within the overburden. The

average dominant frequency of the bedrock horizon on processed sections is 80 Hz for


185
Figure 6.15. Line Test-1 data stacked using different ranges of source to receiver offsets
(ranges specified on stacks are in feet). Stacking all offsets allowed the overburden and
bedrock interface (110 to 120 ms) to be most effectively imaged. Fold (TR_FOLD) plots
are shown on stacks, and CDP location numbers (CDP_X) correspond to road stations.
186
lines Test-1 and GUE-I70-1 (the dominant frequency increases slightly to 90 Hz on the

east end of the line Test-1 section), and 100 Hz for line EBPassYY.

6.6.1 Line Test-1

The processed and stacked line Test-1 (Figure 6.1) data are shown in Figure 6.16.

The reflection event imaged at 110 to 120 ms (Figure 6.16a) correlates to drill log data,

and is the top of bedrock (Figure 6.16b). The horizon is continuous across the line, with

slight apparent dip of the bedrock surface evident between road stations 48345 and

48380. The slight apparent dip in the horizon is also evident after time-to-depth

conversion (Figures 6.16c and 6.16d). No significant vertical displacements or lack of

continuity can be detected along the bedrock horizon in line Test-1 sections, and

therefore no areas of severe disruption related to mine-related subsidence activity (which

would indicate a potential risk for surface collapse) are interpreted across this line.

Inferences regarding the possible effects of the mine on the bedrock horizon

topography and stacking velocities are difficult using line Test-1 data, as no data

regarding the location of room and pillars relevant to the location of line Test-1 were

available (Figure 6.1). However, the coal pillar mapped to the north of the eastern half of

line Test-1 (Figure 6.1) can be projected to intersect line Test-1 at road station 48345. To

the immediate east of this road station is where the apparent dip in the horizon is evident

from stacked data, and where a transition (decrease) in horizon stacking velocity occurs

(Figure 6.16a). These observations suggest that the absence of a coal pillar may have

influenced the bedrock topography along this line, and may have also resulted in a

decrease of the overburden material stiffness.


187
188
Figure 6.16. Line Test-1 stacked section with fold (TR_FOLD) plot (a). The continuous reflection event at 110 to 120 ms (a), is
the top of bedrock (b). The color bar on the bottom x-axis of (a) shows bedrock horizon stacking velocities. Uninterpreted (c)
and interpreted (d) depth sections are also shown. CDP location numbers (CDP_X) correspond to road stations (road stations
are given in feet from the western county line). To the east of CDP 48345 the overburden-bedrock horizon dips down to the east
and stacking velocities decrease, suggesting that the removal of coal in this region (see text for discussion) may have influenced
the bedrock topography and resulted in a decreased overburden stiffness. (continued)
Figure 6.16. (continued)

189
(c) and (d) Line Test-1 stacked depth section.
6.6.2 Line GUE-I70-1

The processed and stacked line GUE-I70-1 data are shown in Figure 6.17. This

line was acquired to the immediate south of the previous roadway collapse feature along

the southern berm of the roadway (Figure 6.1). The reflection event imaged at 110 to 120

ms (Figure 6.17a) correlates to shot gather events and drill log data, and is the top of

bedrock (Figure 6.17b). A number of additional weak reflectors between 50 and 100 ms

are also evident within the bedrock overburden. Due to the frequent presence of

discontinuous overburden units (as indicated by drill log data), it is difficult to infer

whether apparent discontinuity or truncation of these events at certain locations is

depositional / lithological in nature, or due to possible subsidence process-related

fracturing or faulting within the overburden materials.

Apparent dip and undulations of the bedrock horizon are apparent across time

sections, and are evident in depth sections (Figures 6.17c and 6.17d). A bedrock high at

CDP location 48460 agrees with the location of a bedrock high interpreted from drill log

data (Figure 6.18) acquired at approximately the same time as these seismic data. A coal

pillar is present beneath bedrock at this location (Figure 6.1), and the bedrock stacking

velocities are relatively high at this location (Figure 6.17a). These observations suggest

that the coal pillar (by supporting overlying materials) may have allowed the original

bedrock topography and overburden material stiffness to be maintained at this location

subsequent to mining activity. Moving to the west of this location (where Figure 6.1

indicates a room beneath line GUE-I70-1), an apparent dip in the bedrock horizon is

evident from stacked sections, and stacking velocities begin to decrease. It is possible

that this apparent dip in the bedrock surface is related to coal absence, however, it is also

190
191
Figure 6.17. Line GUE-I70-1 stacked section with fold (TR_FOLD) plot (a). The reflection at 110 to 120ms (a), is the top of
bedrock (b). The color bar on the bottom x-axis of (a) shows bedrock stacking velocities. Uninterpreted (c) and interpreted (d)
depth sections are also shown. CDP location numbers (CDP_X) correspond to road stations (road stations are given in feet from
the western county line). A bedrock discontinuity is interpreted at CDP_X 48391, and an area of disrupted bedrock is
interpreted between CDP_X 48380 and 48408 (indicated by x-axis arrows and dashed lines). The interpreted area of disruption
is based upon on wavelet character and analysis of the shot gathers (x-axis flags indicate locations) in Figure 6.19. (continued)
Figure 6.17. (continued)

192
(c) and (d) Line GUE-I70-1 stacked depth section.
possible that the horizon exhibited apparent dip (in this area) prior to mining (data

regarding bedrock elevation prior to mining were not available).

The geologic cross-section shown in Figure 6.18 was interpreted based

predominately upon the wells with core recovery (indicated by solid as opposed to

dashed vertical lines in the figure). Between CDP locations 48330 and 48360 a relative

bedrock low was interpreted from drill log data, and bedrock is also seen to be relatively

low in this region on the seismic sections (from CDP locations 48330-48370). Figure 6.1

indicates the presence of a mine room in this region, and the previous localized collapse

occurred at road station 48345 (about 15 feet north of line GUE-I70-1), however stacking

velocities are somewhat higher this area relative to laterally adjacent areas, and offset or

significant disruption of the bedrock horizon is not evident in seismic sections (Figure

6.17). These observations indicate that either the subsidence processes (at depth) that

were responsible for the previous roadway collapse were predominately active to the

north of Line GUE-I70-1, or that remediation efforts of the previous collapse area have

prevented the continuation of subsidence activity at this location. It is also possible that

evidence for bedrock disruption resulting from ongoing subsidence processes since

remediation was not seismically detectable at this location using these data.

There is no evidence on the seismic sections that indicates that the voids

encountered beneath the bedrock surface during drilling to the west of the previous

collapse (Figure 6.18) have continued to propagate upwards through the bedrock surface

and pose an immediate risk for surface collapse. This interpretation is supported by

cross-hole radar data analyses and drill log data acquired (subsequent to this seismic data

interpretation) during 2002 (Appendix D).

193
194
Figure 6.18. Geologic cross-section constructed from 1999 drill log data acquired along the southern edge of the eastbound
travel lane of I-70 (Modified from BBC&M Inc., 1999).
A discontinuity in the bedrock that resulted in vertical horizon offset is interpreted

on the seismic sections (Figures 6.17b and 6.17d) at CDP location 48391. Based on the

apparent differences in travel time (and depth on the depth sections) across the

discontinuity, the vertical offset is estimated from the seismic section to be between 3 and

4 ft (which is greater than a quarter of the dominant seismic wavelength in the

overburden). The applied stacking velocity did not change abruptly in the vicinity of this

discontinuity (the stacking velocities were relatively low across this region, possibly due

to a fractured and relatively weak overburden), and the horizon appears to be fairly

continuous on either side of the discontinuity. A well drilled during 2002 at road station

48395 based on this interpretation, confirmed that bedrock had in fact been down-

dropped along a normal fault between road stations 48380 and 48395, and indicated

heavy fracturing of the bedrock with no coal encountered in the borehole (Appendix D).

Despite the stacked sections indicating a fairly continuous horizon in the immediate

vicinity of the mapped discontinuity, indications of severe disruption in the bedrock

surface across this area exist in the seismic data.

Figure 6.19a shows uninterpreted shot gathers acquired at the source locations

indicated by flags on the x-axis of Figures 6.17b and 6.17d, and the gathers are

interpreted in Figure 6.19b. For the 48365.5 source location gather, the bedrock event is

represented by a fairly well behaved hyperbola. The apex of this event is shifted to the

right of the source in the 48375.5 gather, indicating an apparent dip of the bedrock

surface (updip direction to the east) at this location (which agrees with the dip apparent at

this location on the seismic sections). Dip-moveout (DMO) corrections were not applied

to the data to compensate for occasionally apparent dips, and as a result less confidence

195
Figure 6.19. Line GUE-I70-1 uninterpreted (a) and interpreted (b) shot gathers with
source locations corresponding to the flag locations on the x-axis of Figures 6.17b and
6.17d. Source locations (SOU_X) and CDP locations (CDP_X) correspond to road
stations. The reflection at 110 to 120 ms (b) is the top of bedrock. The 48375.5 shot
gather indicates an apparent updip direction to the east at this location. The 48389.5 shot
gather supports the interpretation in Figures 6.17b and d of discontinuity and offset in the
bedrock horizon at approximately CDP location 48391. Based on shot gather reflection
character, an area of disrupted bedrock is interpreted between CDP’s 48380 and 48408
(indicated on the x-axis of interpreted gathers by arrows).

196
in a calculated depth of the bedrock horizon estimated using the stacking velocity would

result at such locations. The bedrock horizon event is severely distorted in the 48389.5

source location gather at near offsets, and this supports the interpretation of the

discontinuity at CDP location 48391 on the stacked sections in Figures 6.17b and 6.17d.

Based on disruptions in the character of the bedrock horizon in the shot gathers in

Figure 6.19b, an area of bedrock disturbance resulting from mine-related subsidence

processes is evident between CDP locations 48380 and 48408. Therefore, it is apparent

that a relatively high potential risk for future surface collapse exists along this section of

the roadway. Cross-hole radar data analyses conducted subsequent to this interpretation

indicated that disruption of the bedrock horizon in this region has occurred between the

boreholes and directly beneath the seismic line (Appendix D). The interpreted extent of

the range of subsurface disruption is also supported by the character of near surface

reflectors (within the overburden) in this area, which exhibit apparent dip and offset

across this CDP range on the seismic sections (Figure 6.17). The lateral extent of the

interpreted area of disturbance is indicated by the x-axis arrows in Figures 6.17b and

6.17d, and is also indicated by dashed vertical lines across the bedrock horizon.

Applying NMO corrections using velocities that best flattened the bedrock event to these

shot gathers resulted in a certain amount of smoothing of such sub-spread length features

on the stacked sections. However, by paying attention to event character in shot gathers

(Figure 6.19) during stacked data interpretation, such features were not overlooked.

A void was encountered beneath the bedrock surface at CDP location 48380

during 1999 drilling that was conducted at approximately the same time as the seismic

survey (Figure 6.18). The geologic cross-section interpreted from 1999 drill logs

197
indicates a coherent and continuous bedrock surface to the east of this location (across the

disrupted area interpreted from the seismic data), because a well logged at CDP location

48420 indicated an intact bedrock surface. The mine map (Figure 6.1) indicates that the

southwest end of a coal pillar (with a northeast strike) is present beneath line GUE-I70-1

between road stations 48375 and 48400. Based on the location of the interpreted area of

disruption (between road stations 48380 and 48408), two different mechanisms (pit or

sag subsidence, see Appendix A) were considered as the responsible mechanism for the

observed bedrock horizon disruption. A complete crushing of the pillar between road

stations 48375 and 48400 would have likely resulted in a broad region of bedrock

subsidence, but the seismic data indicate that bedrock is intact and at a relatively high

elevation at road station 48380 (Figure 6.17). It therefore appears that the observed

bedrock disruption resulted from a collapse of the bedrock horizon into the mine room

located to the immediate southeast of the coal pillar. Cross-hole radar data analyses and

2002 drill log data indicate that the mine map in Figure 6.1 has placed the eastern edge of

this coal pillar too far to the east (Appendix D).

6.6.3 Line EBPassYY

The processed and stacked line EBPassYY data are shown in Figure 6.20. This

line was acquired to the north of the previous roadway collapse feature (Figure 6.1). The

reflection event imaged at 105 to 120 ms (Figure 6.20a) correlates to shot gather events

and drill log data, and is the top of bedrock (Figure 6.20b). Bedrock is seen to be

continuous across the length of line EBPassYY from the time sections and the

uninterpreted (Figure 6.20c) and interpreted (Figure 6.20d) depth sections, except for

198
199
Figure 6.20. Line EBPassYY stacked section with fold (TR_FOLD) plot (a). The reflection at 105 to 115 ms (a), is the top of
bedrock (b). The color bar on the bottom x-axis of the uninterpreted section (a) shows bedrock horizon stacking velocities.
Uninterpreted (c) and interpreted (d) depth sections are also shown. CDP location numbers (CDP_X) correspond to road
stations (road stations are given in feet from the western county line). The bedrock horizon is continuous across the section,
except between CDP_X 48329 and 48354 (indicated by x-axis arrows), where discontinuity is interpreted. (continued)
Figure 6.20. (continued)

200
(c) and (d) Line EBPassYY stacked depth section.
between CDP locations 48329 and 48354. In this range (indicated in Figures 6.20b and

6.20d by arrows on the x-axis and normal faults on seismic data), the bedrock horizon is

interpreted to have been down-dropped, and experienced significant disruption resulting

from pit-type mechanism (Appendix A) mine-related subsidence processes. Cross-hole

radar data analyses indicate that disruption of the bedrock horizon in this region has

occurred between the boreholes and directly beneath the seismic line (Appendix D). The

mine map in Figure 6.1 indicates that a mine room is present beneath line EBPassYY to

the immediate east of road station 48330. Cross-hole radar data analyses and drill log

data suggest that the mine map has placed the eastern edge of the coal pillar at 48330 too

far to the east (Appendix D).

The stacking velocities slowly increase towards the west across the disrupted

bedrock region, and the horizon would not stack coherently across this region at any of a

wide range of applied stacking velocities (Figure 6.12). Based on the apparent

differences in depth on the seismic sections, the downward displacement of the horizon is

estimated from the seismic data to be between 3 and 4 ft in the CDP location ranges of

48329 and 48339, and 48348 and 48354. Between CDP locations 48339 and 48348, it

appears that a greater amount of downward displacement has occurred. Disruption

responsible for a seismic anomaly does not necessarily need to be centered directly

beneath the seismic line, due to the Fresnel zone concept (the diameter of the Fresnel

zone at the time of the bedrock horizon for these data is about 23 ft). Low amplitude

scattering was recorded above the down-dropped bedrock across the disrupted region,

and is possibly the result of an impedance contrast that formed from overburden material

subsidence into the collapse feature. Alternatively, this low amplitude energy could have

201
resulted from out of plane scattering related to previous remediation efforts conducted to

the south of this line (i.e. backfilling of the previous roadway collapse feature). The fact

that the previous collapse feature was located approximately 20 feet to the south of this

line however (centered at station 48345 in the eastbound travel lane) suggests that the

latter is less likely.

A geologic cross-section interpreted from drill logs acquired during 1999 along

the north berm of the eastbound passing lane is shown in Figure 6.21. The cross-section

interpreted from these logs indicates that bedrock is continuous across the length of the

seismic line, except for in the vicinity of the disrupted zone interpreted from the seismic

data. Drill log data indicate that a bedrock low exists in the approximate road station

range of 48314 to 48340. Voids were also encountered during drilling below the bedrock

surface in this region. The western edge of the disturbed area in seismic data is about 10

feet to the east of the bedrock low interpreted from the drill log data, indicating that

interpretation based on drill log data placed the western edge too far to the west, or that

the seismic data were not able to exactly delineate the western edge of disturbance. It is

not possible from these data alone, to say whether these features (both with lateral extents

of about 25-26 ft) are the same, as the drill log data and seismic data were not co-

registered to the same survey. Based on the modeled results in Figure 6.10, the apparent

width of a subsidence feature interpreted from seismic data would be expected to be

somewhat less than the actual width of the feature, which may also explain this

interpretation difference.

202
Figure 6.21. Geologic cross-section interpreted from 1999 drill log data acquired along
the north edge of the eastbound passing lane of I-70 in the previous roadway collapse
region (Modified from BBC&M Inc., 1999).

203
A drill log at road station 48340 indicated that bedrock had not subsided at this

location (Figure 6.21), however, the seismic data (Figure 6.20) indicate that mine-related

subsidence processes have been active to the east of road station 48340. Analyses of

cross-hole radar data conducted during 2002, subsequent to the interpretation of these

seismic data, indicate that the area of mine-related bedrock horizon disruption directly

beneath line EBPassYY extends from road stations 48314 to 48354 (Appendix D). Voids

were also encountered beneath the bedrock surface in a well drilled at road station 48360.

The seismic data show an intact bedrock surface at this location, and indicate that these

voids have not yet propagated upwards through the bedrock surface.

204
CHAPTER 7

DEVELOPMENT OF SH-WAVE INTROMISSION ANGLE THEORY

7.1 Overview

With regards to SH plane-wave reflection coefficients, there exists an incidence

angle (intromission angle) for which the reflection coefficient is zero (total transmission)

given certain material property conditions. A SH-wave intromission angle is somewhat

analogous to a P-wave intromission angle (Ziomek, 1995) that can occur under certain

conditions in acoustic media. By knowing the conditions under which a SH-wave

intromission angle can occur, the recognition of such an angle from SH-wave reflection

events in field source or CMP sorted gathers could allow certain inferences regarding

media physical property relationships to be made. This concept could be especially

useful for characterizing the deepest detectable interface, in a situation where it consists

of a relatively high shear velocity layer overlying a layer of lower shear velocity, as a

critically refracted arrival (which would allow the velocity of the lower layer to be easily

determined) would not be recorded. The SH-wave intromission angle concept and its

potential applications in seismology have not previously been discussed in literature to

my knowledge. A brief overview of the concept is presented, along with equations

(derived from the SH plane-wave reflection coefficient) to define this angle.

205
7.2 P-Wave Intromission Angle in Acoustic Media Overview

For P-waves in acoustic media, the plane-wave reflection coefficient (Rayleigh

reflection coefficient) for any angle of incidence is defined by the following equation

(nomenclature modified from Ziomek, 1995):

P1P1 = Z2cosșP1P1 - Z1cosșP1P2 / Zs2cosșP1P1 + Zs1cosșP1P2 (7.1)

Where P1P1 is the P-wave reflection coefficient, șP1P1 is the angle of P-wave incidence

and reflection, șP1P2 is the angle of P-wave refraction, and Z1 and Z2 are the incident

(medium 1) and refracted (medium 2) media compressional impedances.

The P-wave acoustic media reflection coefficient is zero when the incidence angle

is equal to the P-wave intromission angle (șI), defined by Ziomek (1995) as:

șI = sin-1(((ȡ2 / ȡ1)2 - (Vp1 / Vp2)2) / ((ȡ2 / ȡ1)2 - 1))1/2 (7.2)

Or written in terms of impedances and velocities from the Ziomek (1995) equation as:

șI = sin-1(((Z2 / Z1)2 - 1) / ((Z2 /Z1)2 - (Vp2 / Vp1)2))1/2 (7.3)

Where ȡ1 and ȡ2 are the incident and refracted media densities, and Vp1 and Vp2 are the

incident and refracted media compressional velocities.

The P-wave intromission angle is a particular real angle of incidence between

zero and 90 degrees. In order to obtain a real solution to either of the equations

describing this angle, it is seen that the numerator and denominator in the equations must

have the same sign, and the absolute value of the numerator must be equivalent to or less

than the absolute value of the denominator. From Equation 7.2 it is seen that these

conditions will be met if: (ȡ2 / ȡ1) ≥ (Vp1 / Vp2) ≥ 1, or if: 1 ≥ (Vp1 / Vp2) ≥ (ȡ2 / ȡ1).

When: (ȡ2 / ȡ1) = (Vp1 / Vp2) = 1, the reflection coefficient is zero for all incidence angles.

206
Compressional Velocity (Vp) and
Physical Property Scenario
Density (ȡ) Relationships
1 (Vp1 / Vp2) > 1 > (ȡ2 / ȡ1)
2 (ȡ2 / ȡ1) > (Vp1 / Vp2) > 1
3 (ȡ2 / ȡ1) > 1 > (Vp1 / Vp2)
4 1 > (Vp1 / Vp2) > (ȡ2 / ȡ1)

Table 7.1. Incident and refracted media (media 1 and 2 respectively) physical property
scenarios used to model the synthetic seismograms and obtain the solutions presented in
Figures 7.1a - 7.1d.

P-wave intromission angle occurrence in acoustic media and potential usefulness

are demonstrated using synthetic seismograms and plots of reflection coefficient

magnitudes and phase (Figures 7.1a - 7.1d), generated for the four physical property

scenarios in Table 7.1. A P-wave intromission angle has been observed in actual field

data, acquired in a marine environment (at a water and low velocity mud interface) by

Winokur and Bohn (1968). Synthetic seismograms were generated using an acoustic

finite-difference modeling code in the ProMAX software package (for details regarding

this code and modeling parameters see Chapter 6). A zero-phase Ricker wavelet (center

source frequency of 100 Hz) was used for each model. Each model consisted of two

layers (layer one parameters represent water in all cases, and layer two parameters vary

between cases), with the interface between layers at a depth of 60 ft. Relationships

between layer densities and P-wave velocities in models are listed in Table 7.1, and the

actual density (units of g/cm3) and velocity (units of m/s) values used for each model are

207
(a) Synthetic data, reflection coefficients, and phase changes as a function of incident
angle for scenario 1 in Table 7.1.

Figure 7.1. Synthetic seismograms (modeled with a center source frequency of 100 Hz)
and plots showing plane-wave reflection coefficients (magnitude) and phase changes
(versus incidence angle) resulting from an incident P-wave in acoustic media, for the
scenarios in Table 7.1: (a) scenario 1, (b) scenario 2, (c) scenario 3, and (d) scenario 4.
The density (units of g/cm3) and velocity (units of m/s) values used for modeling are
shown on each of the outputs in this figure. Absolute offset values from the source on the
x-axis of synthetic data are in feet. Values of critical distance (xc), critical angle (șc),
intromission distance (xI), and intromission angle (șI) are indicated. Each model
consisted of two layers, with the interface between layers at 60 ft depth. See text for
discussion. Plots were generated using the code listed subsequently. (continued)

208
Figure 7.1. (continued)

(b) Synthetic data, reflection coefficients, and phase changes as a function of incident
angle for scenario 2 in Table 7.1. (continued)

209
Figure 7.1. (continued)

(c) Synthetic data, reflection coefficients, and phase changes as a function of incident
angle for scenario 3 in Table 7.1. (continued)

210
Figure 7.1. (continued)

(d) Synthetic data, reflection coefficients, and phase changes as a function of incident
angle for scenario 4 in Table 7.1.

211
shown on the outputs presented in Figures 7.1a - 7.1d. The plots of reflection coefficient

magnitudes and phase presented in Figures 7.1a - 7.1d were generated using the program

p_acoustic.m (written in MATLAB, version 6), and the short source code for this

program is listed subsequently.

Figure 7.1a contains solutions for a situation where: (Vp1 / Vp2) > 1 > (ȡ2 / ȡ1). In

this case no angle of intromission occurs, and because the incident media (medium 1)

velocity is less than the refracted media velocity (medium 2) no critical refraction occurs

either. The refracted media impedance is less than that of the incident media, and as a

result a polarity reversal occurs from the model interface (the phase is constant with

offset). A low amplitude multiple reflection (at around twice the arrival time of the

primary) is seen on each of synthetic seismograms presented.

For the Figure 7.1b model parameters: (ȡ2 / ȡ1) > (Vp1 / Vp2) > 1. The reflection

event has the same polarity as the direct wave up until the intromission angle (about 53

degrees), at which the reflection coefficient goes to zero, and beyond which a polarity

reversal is evident. In this case there is no refracted arrival, which would allow an easy

velocity determination of the lower model layer. Assuming we did not already know the

model parameters, we could determine the interface depth and the average velocity of

medium 1 above the interface from the arrival time and apparent NMO velocity of the

reflection. By observing the intromission distance (xI) and using the calculated interface

depth we could then determine the angle of intromission. Because the polarity of the

reflection and direct arrival are similar at normal incidence we would know that the

media property relationships must be: (ȡ2 / ȡ1) ≥ (Vp1 / Vp2) ≥ 1, and we could then (by

212
assuming a flat interface, and a density relationship if density measurements were not

available) approximate the velocity of the lower layer.

For the Figure 7.1c model parameters: (ȡ2 / ȡ1) > 1 > (Vp1 / Vp2), and as a result no

intromission angle is observed, but a critical angle occurs at about 42 degrees (the critical

distance, xc, is indicated on the synthetic seismogram). For the Figure 7.1d model

parameters: 1 > (Vp1 / Vp2) > (ȡ2 / ȡ1), and as a result both an intromission angle and a

critical angle are observed (an intromission angle value is always less than a critical angle

value). A polarity reversal is seen to occur across the intromission angle, with the phase

then constant until the critical angle is reached, beyond which the phase is variable

(Figure 7.1d). FDTD modeled acoustic P-wave snapshots using the Figure 7.1d model

parameters are shown in Figure 7.2. The reflection amplitude null and phase changes

indicated in Figure 7.2d are associated with the intromission angle occurrence for these

media physical property conditions.

7.3 Conditions for SH-Wave Intromission Angle

When an interface between two homogenous, isotropic solids is parallel to the

polarization of an incident SH-wave, only scattered SH-waves are generated, regardless

of the incident angle (Chapter 2). For SH-waves, the reflection coefficient for any angle

of incidence is defined by the following equation (nomenclature modified from Shearer,

1999):

SH1SH1 = Zs1cosșSH1SH1 - Zs2cosșSH1SH2 / Zs1cosșSH1SH1 + Zs2cosșSH1SH2 (7.4)

Where SH1SH1 is the SH-wave displacement amplitude reflection coefficient, șSH1SH1 is

the angle of SH-wave incidence and reflection, șSH1SH2 is the angle of SH-wave
213
214
Figure 7.2. FDTD modeled snapshots of acoustic P-wave scattering from a planar interface (located at 60 feet depth), for the
physical properties scenario number 4 in Table 7.1: (a) 35 ms, (b) 45 ms, (c) 55 ms, and (d) 65 ms. The reflection amplitude null
and phase change in (d) are associated with intromission angle occurrence under these modeled media conditions.
refraction, Zs1 is the incident media (medium 1) shear impedance, and Zs2 is the refracted

media (medium 2) shear impedance.

The SH-wave reflection coefficient is zero when the incidence angle is equal to

the SH-wave intromission angle (șI), and this occurs when:

Zs1cosșI = Zs2cosșSH1SH2 (7.5)

To solve for șI:

Zs1 / Zs2 = cosșSH1SH2 / cosșI (7.6)

Using sin2x + cos2x = 1:

Zs1 / Zs2 = (1 - sin2șSH1SH2)1/2 / (1 - sin2șI)1/2 (7.7)

And using Snell’s law to eliminate șSH1SH2:

Zs1 / Zs2 = (1 - (Vs2 / Vs1)2 sin2șI)1/2 / (1 - sin2șI)1/2 (7.8)

Which yields a solution for șI in terms of media shear velocities and impedances:

șI = sin-1(1 - (Zs1 / Zs2)2 / (Vs2 / Vs1)2 - (Zs1 / Zs2)2)1/2 (7.9)

Or a solution for șI in terms of media shear velocities and densities:

șI = sin-1((Vs2 / Vs1)2 - (ȡ1 / ȡ2)2 / (Vs2 / Vs1)4 - (ȡ1 / ȡ2)2)1/2 (7.10)

Where ȡ1 and ȡ2 are the incident and refracted media densities, and Vs1 and Vs2 are the

incident and refracted media shear velocities.

The SH-wave intromission angle is a particular real angle of incidence between

zero and 90 degrees. In order to obtain a real solution to either of the equations

describing this angle above, it is seen that the numerator and denominator in the

equations must have the same sign, and the absolute value of the numerator must be

equivalent to or less than the absolute value of the denominator. From Equation 7.9 it is

seen that these conditions will be met if: (Zs1 / Zs2) ≥ 1 ≥ (Vs2 / Vs1), or if: (Vs2 / Vs1) ≥

215
Physical Property Scenario Shear Velocity (Vs) and Impedance (Zs) Relationships
1 1 > (Vs2 / Vs1) = (Zs1 / Zs2)
2 (Zs1 / Zs2) > 1 > (Vs2 / Vs1)
3 (Vs2 / Vs1) = (Zs1 / Zs2) > 1
4 (Vs2 / Vs1) > 1 > (Zs1 / Zs2)

Table 7.2. Incident and refracted media (media 1 and 2 respectively) physical property
scenarios used to obtain the solutions presented in Figures 7.3 and 7.4.

1 ≥ (Zs1 / Zs2). When: (ȡ1 / ȡ2) = (Vs2 / Vs1) = 1, the reflection coefficient is zero for all

incidence angles.

In order to demonstrate SH-wave intromission angle occurrence and potential

usefulness in elastic media, plots of reflection coefficients and phase (along with

refraction coefficients, square root energy coefficients and energy coefficients) were

generated for the four physical property scenarios listed in Table 7.2, and are shown in

Figures 7.3a - 7.3d. The relationships between the layer densities and shear velocities

used to obtain the solutions presented are listed in Table 7.2, and the actual density and

velocity values used for each case are listed on the plots presented as Figures 7.3a - 7.3d.

The solutions in Figures 7.3a - 7.3d were obtained using the program PSHSV (written in

MATLAB, version 6), and the source code for this program is described and listed in

Appendix B. Synthetic SH-wave reflection seismograms were also generated using the

Sierra (Landmark Graphics Corporation) modeling program, for physical property

scenarios 1 and 2 in Table 7.2, and are shown in Figure 7.4. An impedance contrast was

specified at 29.5 feet depth for the synthetic models, with the shear velocity, impedance,

and density relationships that were used shown on the seismograms in Figure 7.4.
216
217
(a) Solutions for scenario 1 in Table 7.2.

Figure 7.3. Amplitude, square root energy, and energy coefficients, and phase changes from an incident SH-wave, for scenarios
in Table 7.2: (a) scenario 1, (b) scenario 2, (c) scenario 3, and (d) scenario 4. Density (ȡ) units are g/cm3, and shear velocity
(Vs) units are m/s. See text for a discussion of plots. Plots generated using the PSHSV code in Appendix B. (continued)
Figure 7.3. (continued)

218
(b) Solutions for scenario 2 in Table 7.2. (continued)
Figure 7.3. (continued)

219
(c) Solutions for scenario 3 in Table 7.2. (continued)
Figure 7.3. (continued)

220
(d) Solutions for scenario 4 in Table 7.2.
Figure 7.4. Synthetic SH-wave reflection seismograms (modeled with a center source
frequency of 100 Hz), for the scenarios in Table 7.2: (a) scenario 1, and (b) scenario 2.
The shear velocity, impedance, and density relationships used are shown on the plots.
Absolute offset values from the source on the x-axes are in feet. The intromission
distance (xI) in (b) is indicated. Each model consisted of two layers, with the interface
between layers at a depth of 29.5 ft. See text for discussion.
221
Figures 7.3a and 7.4a contain solutions for a situation where: 1 > (Vs2 / Vs1) = (Zs1

/ Zs2). In this case no angle of intromission occurs, and no critical angle occurs because

the incident media (medium 1) velocity is less than that of the refracted media (medium

2). The refracted media impedance is greater than that of the incident media, and no

change in phase occurs from the media interface (phase is constant with offset).

For the Figures 7.3b and 7.4b model parameters: (Zs1 / Zs2) > 1 > (Vs2 / Vs1). The

reflected SH-wave phase is constant until the intromission angle (about 46 degrees), at

which the reflection coefficient goes to zero, and beyond which a 180 degrees phase shift

of the reflected pulse is evident. In this case there is no critically refracted wave, which

should allow an easy velocity determination of the lower layer from field data. Assuming

we has acquired a source (or CMP gather) over media with these physical properties, and

we did not already know the media properties, we could determine the interface depth

and the average velocity of medium 1 above the interface using arrival time and an

apparent NMO velocity of the reflection. By observing the intromission distance (in the

vicinity of which a dimming of the event amplitude may be noticeable, or beyond which

a phase shift may be recognizable) and using the calculated interface depth, we could

then determine the angle of intromission. Because the polarity of the reflection event and

the direct arrival in field data would be similar at normal incidence for these media

properties, we would know that the media property relationships must be: (Zs1 / Zs2) ≥ 1

≥ (Vs2 / Vs1). We could then (by assuming a flat interface, and a density relationship if

density measurements were not available) approximate the velocity of the lower layer. If

a reflection beneath this lower layer was recorded in field data, the velocity of the lower

layer could be approximated using an approach such as the Dix (1955) method.

222
However, in a situation where no reflection from beneath the lower model layer was

recorded, the intromission angle approach for estimating the lower layer velocity could

be especially useful.

For the Figure 7.3c solutions: (Vs2 / Vs1) = (Zs1 / Zs2) > 1, and as a result no

intromission angle is observed, but a critical angle occurs at about 65 degrees. For the

Figure 7.3d solutions: (Vs2 / Vs1) > 1 > (Zs1 / Zs2), and as a result both an intromission

angle and a critical angle are observed. A 180 degrees phase shift for the reflected SH-

pulse is seen to occur at the intromission angle (about 42 degrees), and the phase is then

constant until the critical angle (at about 64 degrees) is reached, beyond which the

reflected pulse phase is variable.

7.4 Source Code Listing: p_acoustic.m

The MATLAB program p_acoustic.m is listed next:

% p_acoustic.m - written and last modified by Erich D. Guy on 3/27/02 %%%%%%%%


% Program computes the magnitude of reflection coefficients (Rayleigh reflection
% coefficients) and phase for compressional plane wave scattering from a planar
% interface in acoustic media.
% User input = P-wave velocity and density values for incident and refracted media.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

clear all; close all;

% User Input
Vp1 = input ('Enter Vp1 (Incident Medium P-Wave Velocity): ');
Vp2 = input ('Enter Vp2 (Refracted Medium P-Wave Velocity): ');
Rho1 = input ('Enter Rho1 (Incident Medium Density): ');
Rho2 = input ('Enter Rho2 (Refracted Medium Density): ');

thetadeg = 0.01:.1:89.99; % Incident angles in degrees


deg = pi/180; rad = 180/pi;
thetarad = thetadeg*deg;
pp = sin(thetarad)/Vp1; % Horizontal slowness
223
Zp1=Vp1*Rho1; Zp2=Vp2*Rho2; % Impedances

% Reflection and refraction angle cosines


costhetaP1P1 = Vp1*(sqrt((1/(Vp1^2))-(pp.^2)));
costhetaP1P2 = Vp2*(sqrt((1/(Vp2^2))-(pp.^2)));

% Critical angle
P1P2thetac = (asin(Vp1/Vp2));

% P intromission angle
thetaI = asin(sqrt((((Rho2/Rho1)^2)-((Vp1/Vp2)^2))/(((Rho2/Rho1)^2)-1)));

% Reflection coefficients
P1P1 = (Rho2*Vp2*costhetaP1P1-
Rho1*Vp1*costhetaP1P2)./(Rho2*Vp2*costhetaP1P1+Rho1*Vp1*costhetaP1P2);

% Plot solutions
set(0,'Units','normal');
ss = get(0,'ScreenSize');
pos = [ss(3)/2-0.8/2,ss(4)/2-0.6/2,.8,.4];
nb = [0,0,.5];

figure('Units','normal','Position',pos,'Name','Incident P-Wave In Medium


1','NumberTitle','off');

subplot(1,2,1);
plot(thetadeg,abs(P1P1),'b-','LineWidth',1.5);
title('Amplitude Coefficients (Magnitude)');
legend('Reflected P');
set(legend,'position',[.001 .003 .113 .067]);
grid on;axis([0 90 -0.05 1.05]);
set(gca,'XTick',0:15:90);
xlabel('Incidence Angle (degrees)');
set(gca,'YTick',0:.25:2);
ylabel('A / A_0');
text(-33,1.1,'Incident P-Wave','fontweight','bold','color',nb);
text(-33,1.05,'In Medium 1','fontweight','bold','color',nb);
N = num2str(Vp1);text(-33,.95,'Vp_1 =','color',nb);text(-21,.95,(N),'color',nb);
O = num2str(Vp2);text(-33,.88,'Vp_2 =','color',nb);text(-21,.88,(O),'color',nb);
P = num2str(Rho1);text(-33,.81,'\rho_1 =','color',nb);text(-21,.81,(P),'color',nb);
Q = num2str(Rho2);text(-33,.74,'\rho_2 =','color',nb);text(-21,.74,(Q),'color',nb);
R = num2str(Rho1*Vp1);text(-33,.67,'Z_1 =','color',nb);text(-21,.67,(R),'color',nb);
S = num2str(Rho2*Vp2);text(-33,.60,'Z_2 =','color',nb);text(-21,.60,(S),'color',nb);
W = num2str(thetaI*rad);text(-33,.38,'\theta_I =','color',nb);
text(-23,.38,(W),'color',nb);

224
X = num2str(real(P1P2thetac)*rad);text(-33,.31,'\theta_c Refracted P =','color',nb);
text(-33,.24,(X),'color',nb);
Y = num2str(Rho2/Rho1);text(-33,.10,'\rho_2/\rho_1 =','color',nb);
text(-19,.10,(Y),'color',nb);
Z = num2str(Vp1/Vp2);text(-33,.03,'Vp_1/Vp_2 =','color',nb);
text(-33,-.04,(Z),'color',nb);

subplot(1,2,2);
plot(thetadeg,angle(P1P1)*rad,'b-','LineWidth',1.5);
title('Phase Changes');
grid on;
axis([0 90 -200 200]);
set(gca,'XTick',0:15:90);
xlabel('Incidence Angle (degrees)');
set(gca,'YTick',-180:90:180);
ylabel('Phase Angle (degrees)');

% End program

225
CHAPTER 8

CONCLUSIONS

This dissertation has improved near-surface characterization potential by testing

and developing non-intrusive methodologies, analyzing key aspects of high-resolution

multicomponent reflection surveying, and designing effective data analysis workflows.

This chapter summarizes the importance and practical application, of developments and

findings that were attained through this dissertation, and provides recommendations on

certain topics for future research.

8.1 Importance of Considering P- and S-Wave Imaging Potential Prior to


Conducting High-Resolution Reflection Surveys
Research concerning the effectiveness of near-surface P- and S-wave reflection

surveys for mapping features in the shallow earth was conducted. Factors influencing the

effectiveness of common-mode P- and S-wave reflection information for allowing near-

surface horizons to be imaged, and discontinuities to be detected were described. Results

of this study demonstrate the necessity of considering the probable usefulness of

recording and analyzing different data components/wave-type reflections for meeting the

objectives of a project, prior to conducting a near-surface seismic reflection survey.

226
Conducting a multicomponent data acquisition test phase and/or modeling based

on geophysical well log information prior to a high-resolution reflection survey will have

associated cost, and for this reason such work is not typically done. Information obtained

through this type of pre-survey work however, will often outweigh the costs of

conducting it, as data components/wave-type reflections can be selected for acquisition

during the actual production-phase survey that will provide the best potential for meeting

project objectives. Dependent upon numerous factors, near-surface characterization may

be able to be most effectively accomplished through the acquisition of one particular data

component/wave-type reflection, or through the acquisition of multiple data

components/wave-type reflections.

The risk of acquiring only one reflection data component without first evaluating

the probability that it will allow project objectives to be successfully met can be great.

For example, in the case of the I-70 mine-subsidence project, seismic reflection survey

objectives would not have been able to be met if only the traditional ZZ (sometimes

called vertical-vertical, or P-P) reflection component had been acquired. Due to

numerous factors (discussed and demonstrated in this dissertation), the acquisition of SH-

wave reflection data was necessary for allowing the I-70 seismic reflection survey

objectives to be effectively met.

8.1.1 Applicability of P- and S-Wave Common-Mode Reflection Information

Due to the field study area subsurface conditions and acquired data

characteristics, the bedrock horizon could be effectively imaged using YY component

(sometimes called crossline-crossline, or SH-SH) reflection information, whereas it could

227
not be effectively imaged or imaged at all using other acquired components. S-wave

reflections from the top-of-bedrock were consistently identified in both XX component

(sometimes called inline-inline, or SV-SV) and YY component data, however, high

amplitude Rayleigh-type surface waves resulted in the optimum reflection window of XX

component data being relatively narrow. Because YY component field records had a

higher reflection signal-to-noise ratio, and a larger optimum reflection window than

corresponding XX component records, accurate estimates of S-wave stacking velocities

could be better obtained, and the targeted bedrock horizon could be more effectively

imaged (and its integrity more accurately inferred) using YY component data.

Although not effective relative to YY component data in the field study area, XX

component data acquisition and analysis may be worthwhile for other near-surface

characterization studies. The near-surface shear-velocity structure was such that high

amplitude Love wave noise did not prevent SH-reflection imaging in the I-70 study area.

In situations where detrimental Love wave noise does saturate the Y-receiver component

however (and prevents effective SH-wave reflection imaging), the XX component may

allow better common-mode S-wave reflection image construction if a large percentage of

incident SV-wave energy is not converted to P-wave energy at a target horizon, and if the

optimum target reflection window is adequate (this will be dependent upon reflection

arrival time, and characteristics of noise modes recorded on X-component receivers).

Although not an objective of this dissertation, and a topic requiring further research, it

may be advantageous to acquire and concurrently analyze multiple S-wave reflection data

components for the purpose of measuring near-surface media anisotropy.

228
S-waves were relatively insensitive to changes in overburden moisture content in

the study area subsurface, whereas common-mode P-wave reflections from the top-of-

saturated-overburden (located above the bedrock) were recorded in ZZ component data,

due to a large P-wave velocity increase across this interface. Airwave and high

amplitude Rayleigh-type surface wave noise had arrival times at near offsets in ZZ

component data that were similar to those of P-wave reflections from this interface. The

frequency content of these noise modes overlapped the optimum P-wave reflection

frequency range, and the necessity of suppressing these noise modes prior to stacking

resulted in a narrow optimum reflection window for this P-wave event. Event arrival

times and characteristics of recorded noise modes made it challenging to process and use

P-wave reflections from the top-of-saturated overburden in the study area.

Forward modeling, shot gather analyses, and consideration of reflection

coefficients indicated that common-mode P-wave reflections from impedance contrasts

beneath the top-of-saturated overburden were not observed in study area field data due to:

1) surface wave and airwave noise, 2) a high-P-wave reflection coefficient at the top-of-

saturated-overburden, 3) low P-wave reflection coefficients at deeper interfaces, and 4)

interference and poor resolution. The percentage of incident P-wave energy reflected (as

P-wave energy) from the top-of-bedrock was small relative to the percentage reflected

from the top-of-saturated overburden, and was also small relative to the percentage of

incident S-wave energy reflected (as S-wave energy) from the top-of-bedrock. Further,

the top-of-bedrock S-wave reflection in YY component data was not severely degraded

by airwave or surface wave noise, and the S-wave reflection window for this event was

wider than the P-wave reflection window for the top-of-saturated overburden event.

229
Because S-waves travel at lower velocity than P-waves, it is possible to increase

resolution potential by using S-wave reflections. In the I-70 study are for example, the

resolution potential of S-waves in dry overburden was calculated as being more than 1.7

times that of P-waves, and in saturated overburden the S-wave resolution potential was

calculated as being more than 4.7 times that of P-waves. Despite the dominant frequency

of recorded S-wave reflections being lower than that of P-wave reflections, recorded

dominant S-wavelengths were smaller than dominant P-wavelengths. The ability to

increase resolution potential (using S-waves) during near surface studies will be

dependent upon media P- and S-wave velocities, P- and S-wave frequencies that can be

generated and recorded, and absorption per unit wavelength of propagating P- and S-

waves in subsurface media.

This study demonstrated that due to differences in P- and S-wave propagation,

media compressional and shear impedance contrasts, and variations in receiver sensitivity

(as a function of orientation), it is necessary to consider the usefulness of recording and

analyzing different data components/wave-type reflections for meeting near-surface

seismic reflection survey objectives. In some situations, one particular component/wave-

type reflection may be necessary, whereas in other cases multiple components/wave-

types may be needed to image all of the targeted horizons. The concurrent analysis of P-

and S-wave reflection information also provides the potential in theory for allowing insitu

elastic media property variations (vertical and lateral) to be mapped in detail. This

objective was unable to be met in the field study area, do to the small number of

reflection events in recorded data, the fact that observed P- and S-wave reflections did

not correlate to similar subsurface interfaces, and the lack of insitu density measurements.

230
8.1.2 Potential for Improving Near-Surface Characterization Using Converted-
Mode Reflection Imaging

The potential for near-surface mode-converted (P-S and S-P) seismic reflection

imaging was tested and evaluated. This topic was addressed experimentally using high-

resolution 9C field reflection data that were acquired over relatively flat-lying geology

(consisting of unsaturated and saturated overburden materials overlying consolidated

units), as well as numerically/theoretically using Zoeppritz equation solutions and elastic-

wave modeling. Primary converted-mode events, having arrival times and moveouts

similar to those predicted from calculations based on acquired common-mode reflection

information, were not observed in acquired multicomponent reflection field data.

Forward modeling results demonstrated that possible (resolution-related) advantages of

near-surface mode-converted reflection imaging could be achieved under noise-free

conditions. Findings of this study also suggested however, that due to the relatively low

amplitudes of mode-converted events and the presence of detrimental random and

coherent noise in field data, it would be difficult from a practical standpoint to use mode-

converted reflection imaging for improving near-surface characterization potential in

studies conducted elsewhere under similar near-surface conditions. Analysis methods

were developed that can be used as a basis for assessing converted-wave imaging

potential in future studies conducted elsewhere.

Factors affecting seismic mode-conversion imaging potential that were different

between previous petroleum multicomponent studies (in which effective P-S converted-

wave imaging has been accomplished) and the engineering-related I-70 multicomponent

field study were recognized. Whereas there is generally an amount of reflected mode-

231
converted energy that is comparable to reflected common-mode energy (at moderate

angles of incidence) for interfaces targeted for petroleum converted-wave imaging, this

was not the case for the I-70 subsurface media conditions. At I-70 subsurface interfaces

of interest, it was found that (due to media impedance relationships) there was little

energy partitioned into converted-modes; scattered mode-converted (P-S and S-P) energy

was not comparable to scattered common-mode (P-P and S-S) energy at any incidence

angles. Further, arrival times of modeled mode-converted reflections from the I-70

interfaces were found to be similar to the observed arrival times of various modes of

high-amplitude coherent noise in the acquired field data. Contrary to this situation, deep-

earth converted-wave events targeted for imaging are often well separated in time and

space from other reflections and detrimental coherent noise modes.

Future field experimentation and theoretical work are necessary to further test and

develop the idea of near-surface converted-wave reflection imaging. Further field

experimentation is necessary in order to investigate the potential for converted-wave

imaging under different environmental/geologic conditions and noise (coherent and

random) levels. Further theoretical work is necessary in order to identify subsurface

media situations that may yield near-surface converted-mode reflections with amplitudes

that are comparable (at moderate angles of incidence and over a substantial range of

incident angles) to those of common-mode reflections.

232
8.2 Factors Affecting SH-Wave Reflection Method Applicability for
Shallow Earth Characterization

Factors affecting the ability of SH-wave reflection measurements for allowing

near-surface horizons to be effectively imaged, and the delineation of subsurface

discontinuities were investigated and described. This was done through modeling and

analyses of SH-reflection data acquired during 1999 and 2001 in the vicinity of a

previous I-70 roadway collapse feature. Results from this study also demonstrated that

the SH-wave reflection technique could be used to image near-surface stratigraphy and

discontinuities in the subsurface that have developed as a result of mine-related

subsidence. A small number of reports have demonstrated that P-wave reflection

methods can provide a means for detecting subsidence activity, however, no examples of

using S-wave reflections for this purpose have been published in refereed literature to

date. By discussing and demonstrating capabilities and limitations of the SH-wave

reflection technique, this study has also increased understanding regarding the

practicality of using S-wave reflection information for near-surface characterization.

This is important considering that the number of published shallow seismic reflection

reports concerning S-waves is very small compared to the number concerning P-waves.

8.2.1 Shallow SH-Reflection Imaging in Complex and High Noise Environments

SH-wave reflection measurements were effective for characterizing the field test

area subsurface, but using these measurements to do so was challenging due to several

factors. The reflection signal-to-noise ratio of acquired SH-wave data suffered from

source-generated noise and high frequency components of traffic noise (which could not

233
be suppressed without degrading reflection quality), and these factors often limited the

range of useful reflection offsets in recorded data. A highly heterogeneous (laterally and

vertically varying) subsurface required that investigations be conducted regarding data

resolution issues, velocity model construction methods, imaging approaches, and

subsurface interpretation criteria prior to data analysis workflow establishment.

This study demonstrated the importance of considering the vertical and lateral

resolution potential of acquired reflection data, as this was a key factor in allowing the

development of an effective processing and interpretation workflow for application to a

large data volume. Through forward modeling it was determined that it was not feasible

to image the complete geologic sequence at a given location in the field study area. Due

to numerous factors, modeling indicated that it was only feasible to image (using SH-

wave reflections) and base interpretations upon the overburden and bedrock interface.

Reflections correlating to this interface could consistently be identified in field records,

and resolution analyses indicated that vertical offsets on the order of several feet and

disrupted or discontinuous areas along this interface with lateral extent less than the

width of mined-out rooms could be inferred from processed reflection information.

Interpretations regarding near-surface media are often based upon shallow seismic

sections that were generated with the assumption of a laterally constant velocity field.

This study demonstrated that large lateral variations in near-surface media velocities are

possible, and must therefore be considered during processing in order to produce accurate

stacked time and depth sections. Over a lateral distance of one hundred feet, variations of

more than 20 percent were observed for average overburden S-wave velocities in the I-70

study area. Due to the low velocity of S-waves (relative to P-waves), the construction of

234
an accurate velocity field for imaging purposes is much more important when using S-

wave reflections than when using P-wave reflections. An integrated velocity analysis

scheme based on shot gathers, constant velocity stacks, and semblance plots was

demonstrated to allow the construction of accurate velocity models, while at the same

time allowing potential drawbacks associated with each of the individual means for

establishing a subsurface velocity field to be avoided.

Due to lower data acquisition and analysis costs than seismic reflection CMP

methods, shallow earth characterization problems are often addressed using common-

offset imaging approaches. In some situations common-offset imaging can offer

additional advantages relative to CMP stacking, and the effectiveness of both approaches

for allowing the I-70 study objectives to be met was therefore evaluated. Common-offset

images that were generated contained a high amount of coherent noise, whereas through

the stacking process CMP sections were produced that had a much higher reflection

signal-to-noise ratio. Although a certain amount of smoothing of the bedrock horizon

was found to occur as a result of NMO corrections prior to CMP stacking, detailed

information regarding localized bedrock structure could still be inferred by paying

attention to the character of reflections in non-corrected field gathers. This study

demonstrated that although more expensive than common-offset imaging approaches,

CMP data acquisition and processing is necessary in some situations for allowing shallow

subsurface characterization to be accomplished using seismic reflection measurements.

235
8.3 Potential Application of SH-Wave Intromission Angle Equations

Equations were derived and presented in this dissertation, which define an

incidence angle for which the SH-wave reflection coefficient is zero (i.e. an angle of

intromission). The potential application of inferring the presence of such an angle from

SH-wave reflection data was also discussed. Such equations and discussion on their

potential application have not been published in the seismology-related literature to date.

Conditions in terms of media density and shear velocity relationships, under

which the SH-wave reflection coefficient will be zero in theory, were described and

demonstrated in this dissertation. Based upon theory, recognition of a SH-wave

intromission angle from reflection events in field source- or CMP-gathers should allow

certain inferences regarding media physical property relationships to be made. This has

previously been demonstrated in the refereed literature under acoustic media conditions

through the recognition of a (somewhat analogous) P-wave angle of intromission in field

data, and was modeled for acoustic media situations in this dissertation.

The SH-wave intromission angle concept could be especially useful for

characterizing the deepest detectable media interface, in a situation where it consists of a

relatively high shear velocity layer overlying a layer of lower shear velocity (an SH-wave

intromission angle can also occur in the opposite case). In this situation a critically

refracted arrival (which would allow the velocity of the lower layer to be easily

determined) would not be recorded. The average velocity and thickness of the overlying

layer (relatively fast in this case) could be determined using the normal incidence arrival

time and measured NMO velocity of a SH-wave reflection. By observing the

intromission distance (in the vicinity of which a dimming of the event amplitude may be
236
noticeable, or beyond which a phase reversal may be recognizable) and using the

calculated interface depth, the angle of intromission could then be determined. By then

considering the direct arrival and reflection (normal incidence) polarities (which would

be similar in this case), assuming a flat interface and a media density relationship (if

density measurements were not available), and applying the SH-wave intromission angle

equations, the shear velocity (and impedance) of the lower layer could be estimated.

Future work and field experimentation are necessary to test the SH-wave

intromission angle theory presented in this dissertation. Such work is necessary in order

to better isolate naturally occurring geologic situations where this theory may be applied,

to determine the practical application of this theory, and to determine the affect that

various factors (e.g. noise modes, other amplitude and phase variations with offset, and

interference) will have on the ability to infer a SH-wave intromission angle from field

reflection data.

8.4 Importance of Developed PSHSV Computer Program

A computer program (PSHSV) was developed that calculates and plots (as a

function of incident angle) amplitude (reflection and refraction/transmission) coefficients,

square root energy ratios, energy coefficients, and phase changes for elastic waves of P-,

SH-, or SV-type incident on a planar interface between elastic media. Source code for

the program is listed and described in this dissertation, and code such as this has not

previously been made publicly available.

The PSHSV program makes rapid analysis of results possible (program

functionality is demonstrated in this dissertation), and it was rigorously tested to ensure


237
solution accuracy. This is important considering the abundance of errors in previous

refereed publications related to seismic amplitude and energy partitioning equations. An

original and intuitive nomenclature that was used for equations in the program source

code is explained, and the equations are well documented in the source code (allowing

the program to be easily modified). Equations describing the amplitude and energy

partitioning of seismic waves at planar geologic interfaces have a wide range of

applications in the earth and physical sciences, and the PSHSV program will therefore be

useful for engineers and scientists that are conducting research, performing exploration

work, and/or teaching.

8.5 Applicability of Cross-Hole Radar for Subsidence Studies: Data


Acquisition and Analysis Considerations

The ability of cross-hole radar methods for providing useful information for mine-

related subsidence studies was demonstrated, by analyzing constant offset profile (COP)

and multiple offset gather (MOG) radar data that were acquired in the I-70 study area.

Data were analyzed, which were acquired using boreholes located near several

subsidence features that had been imaged using SH-wave seismic reflection. This study

demonstrated for the first time that cross-hole radar surveying could provide useful

information for mine-related subsidence studies. Results showed that cross-hole radar

surveys can provide insight into the nature and extent of fracturing and void space within

near-surface media, and reduce uncertainty regarding the locations and extent of mine

rooms, coal pillars, and seismically imaged subsidence features.

238
8.5.1 Borehole Radar Data Acquisition Considerations

Factors influencing the decision on what type of borehole radar data to acquire

and analyze include the surveying objectives, radar system/antenna characteristics,

attenuation characteristics of subsurface media, and separation distances of available

boreholes. Single-hole radar measurements offer the potential for allowing physical

property changes associated with subsidence activity to be detected. Therefore, in studies

where the spacing of available boreholes prevents the recording of high-quality cross-

hole radar data, single-hole measurements can be considered as a valid option for

addressing subsidence concern. In cases where the spacing between wells is generally

small enough to allow the acquisition of high-quality COP and MOG cross-hole data

however, there are benefits of acquiring cross-hole as opposed to single-hole radar data.

It is worthwhile to first acquire and analyze COP data prior to MOG data acquisition.

For most commercial radar systems the antennas are omni-directional, meaning

that potential anomalies in recorded data can be spatially isolated more accurately and

efficiently by acquiring cross-hole measurements. Cross-hole surveys are also better

suited for addressing subsidence concern than single-hole surveys, because media

physical property distributions can be more accurately mapped between boreholes. An

additional factor that must be considered during the design of borehole radar surveys is

that the single-hole receiver response can be strongly influenced by conductive cable-

related affects, and such affects can severely complicate single-hole data analysis.

239
8.5.2 Cross-Hole Radar Data Analysis Considerations

Processing and imaging flows were presented and discussed in this dissertation

for COP and MOG radar data. These flows yielded data-derived plots and images that

conjunctively served as a basis for interpretations regarding mine-subsidence activity in

the subsurface. The data analysis workflows presented in this dissertation, are robust

enough to serve as a basis (with modification of field data-specific process parameters)

for data analyses in other near-surface cross-hole radar studies.

Plots of average absolute amplitude versus depth were shown to be useful for

allowing signal loss related to increases in (fully saturated) media conductivity and

discontinuity scattering to be inferred. Field data amplitude anomalies correlated in

many cases with seismic reflection data discontinuities and fracture zones and voids that

were encountered during drilling. In studies where such correlations/interpretation

quality controls do not exist, consideration of the somewhat qualitative nature of such

amplitude information is important. Recent reports in the literature have demonstrated

that in order to quantitatively relate amplitude values with actual media conductivity and

attenuation values, it will be necessary to consider and correct for the affect that borehole

and surrounding media characteristics have on source and receiver antenna patterns.

Processed velocity tomograms were shown to be useful for allowing lateral and

vertical media variations, and increases in secondary porosity due to subsidence activity

between boreholes to be inferred in the I-70 subsurface. Media depths that were

investigated using cross-hole radar in the I-70 study area were below measured

groundwater levels. For studies that are concerned with detecting fracture zones and

voids in dry/non-fully saturated media, using velocity tomography to detect such features

240
is possible but more difficult. This is because the relative dielectric constant of air is

much closer to relative dielectric constant values of geologic media than that of water,

and changes in secondary porosity in dry media will therefore result in measured velocity

variations of much smaller magnitude than those in fully saturated media cases.

Due to the problem of non-uniqueness, it is important to establish confidence in

processed tomograms prior to basing interpretations on them. In this study, such

confidence was established by considering only high-quality travel time picks for

inversions, initially conducting trial inversions using different starting models, placing

field data analysis-based constraints on inversion solutions, and quality-checking

tomogram velocities against drill logs and COP-data derived average velocities.

241
APPENDIX A

SEISMIC REFLECTION FIELD TEST AREA

A.1 Location

The Interstate 70 (I-70) study area is a 2200 ft section of the highway (between

road stations 46700 and 48900) located between Cambridge and Old Washington in

Guernsey County, Ohio, approximately 4 miles east of Route 77 (Figure A.1). Road

stations are specified in feet from the western Guernsey county line. The site is located

in the unglaciated region of the Appalachian Plateau physiographic province and lies in a

broad valley with steep sloping sides that is drained by the Mud Run Creek. The Mud

Run Creek channel and drainage was significantly altered during the construction of I-70

from 1961-1962 (Hoffman et al., 1995), and surface drainage since construction

additionally flows into poorly graded ditches on the south and north sides of the roadway.

The ground surface elevation in the study area ranges from 821 feet at the east end (road

station 46700) to 830 feet at the west end (road station 48900).

A.2 Geologic Setting

Regionally, the area encompassing the I-70 study site (Figure A.2) is generally

characterized as relatively flat to mildly dipping Paleozoic sedimentary rocks with

242
Figure A.1. Location of study area along I-70 in Guernsey County, Ohio, where the
roadway crosses approximately 2200 feet of the underground Murray Hill No. 2 mine.
The study area is east of Cambridge, Ohio, roughly 4 miles east of Route 77.

occasional gently plunging folds. Unconsolidated overburden materials in the region

were formed by periglacial (glacial margin) erosion and deposition processes. The upper

5-15 feet of material beneath the highway consists of silt and clay fill, and beneath this

fill are silts and clays down to bedrock, with interbedded lenses of sand and gravel

frequently observed. Geologic cross-sections for certain sections of the study area were

constructed using tightly spaced drill log information, and will be presented subsequently

in this dissertation. The total thickness of unconsolidated materials above bedrock

generally ranges from 30-50 feet across the study area, and is predominantly in the range

of 40-50 feet for most of the study area.

243
244
Figure A.2. Photograph and map view of the I-70 study area (stations 46700 to 48900) in Guernsey County, Ohio.
Bedrock beneath the overburden correlates as the Lower Mahoning Sandstone and

Shale (upper Pennsylvanian series) member in the Lower Glenshaw Group, which is

formally classified as a unit consisting of interbedded sandy shales and sandstones

(Crouch et al., 1980). The inspection of drill cores acquired at various locations indicates

that the Lower Mahoning Sandstone and Shale can be predominantly characterized in the

I-70 study area as arenaceous shale. Bedrock above the coal ranges in thickness from 25

feet at the east end, to 10 feet at the west end of the study area, with a regional strike of

N30oE and a dip of less than one degree to the southeast. Stratigraphically, the Lower

Glenshaw Group is located at the bottom of the Conemaugh Formation, which is a thick

(average of 425 feet) repetitive sequence of sandstone, mudstone, sandy shale, and thin

beds of coal, clay, and limestone (Condit, 1912). The Conemaugh Formation lies just

below the Monongahela Formation, and just above the Upper Freeport Coal (No. 7) of

the Allegheny Formation (middle Pennsylvanian series).

The Upper Freeport Coal ranges from 5-7 feet thick at the study site, dips in some

locations as much as 5 or 6 degrees, and is underlain by claystone (also part of the

Allegheny Formation). The inspection of drill cores acquired at various locations

indicates that the Upper Freeport can be characterized in the I-70 study area as

bituminous coal. Condit (1912) indicates that abrupt changes in thickness of the Upper

Freeport coal are common in the Cambridge area, with the thickness generally varying

where the mine top is sandstone, but generally uniform where the mine top is shale. The

coal is reduced in thickness or entirely cut out in many places regionally where overlying

channel sandstones are present. Historically, the Upper Freeport has been the lead low

sulfur coal produced in the Pittsburgh area (Gray and Meyers, 1970), and was mined an

245
average of 6 feet in thickness in the study area. The elevation of the bottom of the coal

ranges from 750-760 feet (60-80 feet below ground surface) in the study area.

Hydrologic well data (static water level measurements) were obtained on a

monthly basis from fully screened monitoring wells that penetrate to the mine level in the

study area. Measurements indicate that water levels in overburden materials have ranged

from within 20-30 feet of the ground surface (above 800 feet elevation) during 1999-

2001, and that the mine had been flooded during this period.

A.3 Subsidence History

Over 6000 abandoned underground mines (the majority of which are coal mines)

exist throughout 35 Ohio counties, and mine-related subsidence has been a problem in the

state dating back to 1923 (Crowell, 1997a). In the study area, I-70 crosses over

underground mine workings that are part of the abandoned Murray Hill No. 2 mine

complex. The location of the I-70 study area relative to the mine, superimposed on a

United States Geological Survey (USGS) topographic map is shown in Figure A.3. A

map showing the I-70 study area relative to the extent of the coal mine room and pillar

workings is presented in Figure A.4. The location and extent of the mine workings

relative to the highway are based upon a map of the Murray Hill No. 2 coal mine (USDI,

1935) that was drafted at a small scale during 1935. Certain areas of the mine workings

were not mapped during mining, and due to the small map scale and possible errors that

exist with regards to room and pillar locations on this map, the mapped spatial

relationship between these features and the I-70 road stations is regarded as approximate.

246
Figure A.3. Location of the I-70 study area relative to the Murray Hill No. 2 coal mine,
superimposed on a USGS topographic map (Modified from BBC&M Inc., 1998).

Figure A.4. Location of the I-70 study area relative to the Murray Hill No. 2 coal mine
room and pillar workings (Modified from ODNR, 1981).

247
During March of 1994 a mine-related collapse pit was recognized by ODOT in

the median of I-70 just east of the 47800 road marker, with two additional subsidence pits

identified off of the berm of the westbound lane just east of the 46800 road marker

(Figure A.5). Studies conducted during April of 1994 concluded that overburden in the

study area was weakened in some locations, and that a high risk for future subsidence at

certain locations along the roadway existed. During August of 1994 the abandoned,

underground Kings coal mine (down dip and south of the Murray Hill No. 2 mine) was

intercepted by a surface mining operation, and pumping of the mine water was necessary

to recover flooded mining equipment and complete surface mining at the location. As a

result, an extensive dewatering (and related roof support loss) of the Murray Hill No. 2

mine complex occurred (the Kings and the Murray Hill No. 2 are connected by main

entries). This dewatering facilitated localized roof failure and soil piping above the mine

workings in the I-70 study area (Hoffman et al., 1995), and subsidence of the weak

overburden resulted in numerous new areas of roadway depression (Figure A.5). ODOT

immediately conducted work (in conjunction with Gannett Fleming Corddry &

Carpenter) to address the accelerated subsidence caused by the mine dewatering. Drilling

conducted at this time encountered voids beneath the roadway and debris in mined areas,

indicating collapse of the mine roof had occurred in several locations in the study area.

Catastrophic failure of the eastbound lanes of I-70 occurred during March of 1995

between the 48300 and 48400 road markers (Figure A.5), resulting in an eastbound travel

lane surface collapse pit roughly 10 feet in diameter that was centered at road station

48345 (Figure A.6). As a consequence of this sudden collapse, a serious accident

involving four vehicles occurred, which resulted in vehicle damage and one person being

248
249
Figure A.5. Study area map: locations of observed subsidence and roadway depression relative to mapped mine workings.
Road stations are in feet from the western county line.
hospitalized. The mine-water pumping conducted at the Kings mine location that

resulted in accelerated subsidence activity in the study area had ceased by this time, and

water levels had returned to pre-dewatering levels (Hoffman et al., 1995).

A.3.1 Remediation Efforts

Immediately subsequent to the recognition of collapse areas along the roadway

during 1994 and 1995, an aggregate material was used to fill depression features, and

they were then patched with asphalt. Following the 1995 collapse the roadway was

closed for 4 months, during which time a mine remediation project was conducted. This

work consisted of drilling and grouting along the study area to secure voids and fill rock

fractures, and the construction of reinforced land bridges (composed of reinforced

concrete) along the westbound lane. Two land bridges were constructed along the

westbound lane at this time: one between the 46850 and 46950 road markers, and one

between the 47400 and 48100 road markers.

After the roadway reopened, additional surface depressions developed during the

spring of 1996, at which time exploratory drilling revealed voids in several locations

where grouting had been previously performed. A second phase of grouting was then

conducted from May through September of 1997. During the two phases of grouting,

approximately 1800 exploratory boreholes were drilled along the roadway in a grid

pattern (Gannett Fleming Corddry & Carpenter, 1995) between the 46882 and 48822

road markers (hereafter referred to as the grouted region of the study area). Relatively

high grout takes were recorded between road stations 46850 and 48700, and grout was

actually pumped into approximately 1500 boreholes in this region (over 18,800 cubic

250
Figure A.6. Photographs of the March 1995 surface collapse pit centered in the
eastbound travel lane of I-70 at road station 48345. The top photograph is from Gannett
Fleming Corddry & Carpenter, and the bottom photograph is from the Ohio Division of
the Federal Highway Administration.

251
yards of grout were pumped into these holes). A map view of the study area showing the

approximate extent of grouting and a contour map indicating grout takes per borehole

relative the I-70 roadway lanes is presented in Figure A.7. Areas of grout takes are seen

to correlate fairly well in this figure with the mapped locations of mine room workings.

Three different types of grout mixtures were used in the study area: 1) barrier, 2)

production, and 3) 80-20 (ODOT, 1995a). A grout curtain was created by tremie

injecting a barrier grout mixture under low pressures into open mine voids that were

encountered in boreholes drilled along the south and north sides of the roadway lanes. A

spacing of 12 ft was used for barrier grout injection boreholes, and the barrier grout

mixture consisted of approximately 10 percent cement, 65 percent flyash, and 25 percent

sand and # 57 gravel. A production grout mixture was tremie injected under low

pressures within the grout curtain when boreholes encountered open mines, and this

mixture consisted of approximately 10 percent cement, 65 percent flyash, and 25 percent

sand. A spacing of 12 ft was used for production grout injection boreholes along the

eastbound and westbound lanes, and a spacing of 24 ft was used for production grout

boreholes drilled along the roadway berms and median. When boreholes encountered

open mine workings a low-pressure injection of a cement grout into the rock overburden

was also conducted. When boreholes encountered coal pillars, an 80-20 grout mixture

(with variable content, but predominantly consisting of 80 percent flyash and 20 percent

cement when used in the grouted region) was tremie injected into the coal pillars, and no

grouting of the rock overburden was conducted.

252
253
Figure A.7. Map of the I-70 study area showing the approximate grouting limits and grout takes (in cubic yards per borehole).
Road stations are in feet from the western county line.
Since the completion of grouting phases in the study area, the roadway has not

collapsed again, however, concern has existed in regards to roadway stability. Research

has continued in the study area to further characterize the subsurface, to determine

whether subsidence processes are ongoing, and to locate potential areas along the

roadway with a risk for future collapse. As a result of subsidence and remediation efforts

(extensive drilling and grouting) that have occurred along the roadway, subsurface

conditions are somewhat more complicated. Cross-sections of the subsurface constructed

from tightly spaced drill logs are presented for certain sections of the study area in this

dissertation (Chapter 6 and Appendix D).

A.4 Coal Mine-Related Subsidence

Subsidence risk associated with coal mining is related to both the method used to

extract coal and the percentage of coal extracted. Additionally, risk is related to the

bearing capacity of the materials above and below the mine, the thickness and weight of

the mine overburden, fractures, hydrogeology, and changes in conditions within the mine

over time (Gray and Meyers, 1970; Speck and Bruhn, 1995). Accelerated subsidence

rates can be caused by soil piping, mine water circulation which can erode pillars and the

mine roof, and by fluctuations of mine water levels which can weaken overlying and

underlying rock, leaving the mine roof unsupported.

A.4.1 Underground Coal Mining Methods

Two mining methods are typically employed for underground coal extraction

(Isphording, 1992): longwall extraction, and the room and pillar method (which has
254
historically been most commonly used in the United States). The longwall mining

method extracts long rectangular coal panels without leaving any pillars to support roof

rock, and although surface subsidence is usually certain, an advantage of the method is

that the related surface subsidence is somewhat controlled and occurs almost immediately

after extraction. The room and pillar method seeks to extract as much coal as possible

(typically 50-70 percent in Ohio mines) while still providing roof support with remaining

coal pillars, wooden or steel supports, and roof bolts (Crowell, 1997b). In some cases a

secondary (sometimes total) extraction of the remaining pillars is conducted in order to

extract the maximum possible amount of coal. Despite the fact that pillars remain in

cases where only an initial (partial) extraction is conducted, a disadvantage of this mining

method is that uncontrolled and unpredictable subsidence often occurs after mining has

ceased (ranging from immediately after to many years later).

A.4.2 Partial Extraction Mining Subsidence Mechanisms

Coal mine-related subsidence in cases where partial extraction (room and pillar)

mining has been conducted can generally be classified as either sag (also referred to as

trough) or pit (also referred to as pothole) subsidence. Sag subsidence is generally

characterized by the settling of a relatively broad area of the surface (referred to as a sag

or trough), while pit subsidence is generally characterized by the formation of a relatively

small and localized collapse feature (referred to as a chimney, crownhole, pit, or pothole).

Whereas sag subsidence features frequently fill with water when they intercept a water

table, pit features generally do not, as they drain into the underlying mine workings

(Crowell, 1997a).

255
Typical examples of subsidence mechanisms into partially extracted room and

pillar coal mine workings are shown in Figure A.8. Sag subsidence (Figure A.8a) is

caused by the failure of coal pillars, either when pillars are crushed under the weight of

overburden, or when the overburden weight causes pillars to punch into an underlying

weak material. Somewhat similar to a situation in which the longwall extraction mining

method is employed, or to a situation in which a high amount of extraction is conducted

when the room and pillar method is employed, differential horizontal strains can occur

(with their magnitudes and directions changing as a function of time) in the marginal

region of a developing sag subsidence feature (Speck and Bruhn, 1995). As seen in

Figure A.8a, a zone of tensile horizontal strain has developed in the area corresponding to

the extent of sag subsidence, whereas a zone of compressive strain (oriented

perpendicular to tensile strain) has developed in the area laterally closer to the center of

the sag feature.

In the case of pit subsidence (Figure A.8b), coal pillars are relatively stable, and

subsidence at the surface results from the upward migration of collapse features that

develop due to the mine roof rock failing between coal pillars. Pit subsidence features

often correlate with the location of mine entry or working intersection locations (where

roof rock spans a large extracted area), with the adjacent pillars and overburden

remaining unaffected (Hoffman et al., 1995). Empirical data (from numerous partial

extraction coal mine subsidence studies) indicate that in situations where there is no

unconsolidated overburden, pit (subsidence) features are typically no larger than 16 feet

across or deep (Waltham, 1989). In situations where unconsolidated overburden does

exist however, a broader expression of subsidence may develop at the surface due to

256
Figure A.8. Typical subsidence mechanisms into room and pillar coal mine workings
(scale is not implied): a) sag subsidence (after Whittaker and Reddish, 1989), and b) pit
subsidence (after Waltham, 1989). Sag subsidence results in the settling of a relatively
broad surface area, and is caused by the failure of coal pillars (when pillars are either
crushed or punched into the underlying material due to overburden weight). Pit
subsidence usually results in the formation of relatively small pit features at the surface,
and is caused by the upward migration of collapse features that develop due to mine roof
failure between coal pillars. See text for a detailed discussion of subsidence mechanisms.

257
the downward migration of overburden material into a collapse feature. In certain cases,

numerous collapse features resulting from adjacent pit subsidence events have coalesced

upon reaching overburden materials, and have resulted in an even broader, non-typical

subsidence feature at the surface (Waltham, 1989).

Empirical criteria (developed through the analysis of data acquired in numerous

partial extraction mine subsidence studies) suggest that pit subsidence often occurs in

situations where the total thickness of overburden (unconsolidated and consolidated) is

less than 165 ft (Crowell, 1997a). For a general pit subsidence case, vertical subsidence

at the surface decreases with increasing overburden thickness and mine depth due to

incomplete compaction of collapse material (Rahn, 1996), and the lateral extent of

subsidence at the surface increases with increasing mine depth (Leveson, 1980). The

amount of vertical subsidence (in situations where piping of overburden materials does

not occur) usually does not exceed the height of the mined area. Detailed methods for

predicting the vertical amount and lateral extent of pit subsidence are presented in Brady

and Brown (1993), Waltham, (1989), and Whittaker and Reddish, (1989).

A.4.3 Interstate 70 Mining and Subsidence Mechanisms

The Murray Hill No. 2 coal mine, which underlies the study area was in operation

from 1912-1935, and was developed using room and pillar methodology (partial

extraction). Four main entries (each consisting of two parallel entries with a remaining

pillar between) were advanced through the mine workings with an orientation of N79oW.

Two of the mine entries (near markers 46900 and 47500) connected large sections of the

mine (Figure A.5), while the other two main entries (near markers 48100 and 48500)

258
were not continuous. Butt entries were driven to the left and right of the main entry

system, with an orientation of N11oE. The rooms from which most of the coal was

extracted were driven at right angles to the butt entries.

Two processes were identified as being responsible for accelerated subsidence

during 1994-1996 in the study area: mine roof failure and soil piping into mine voids

(Hoffman et al., 1995). Localized mine roof failures resulted due to the mine dewatering

caused by the pumping of the nearby Kings mine. Dewatering of the Murray Hill No. 2

mine left roofs unsupported, lowered lateral support pressures on the coal pillars, and of

lesser importance caused overburden materials to be weakened due to air exposure. The

dewatering created a downward flow gradient, which carried soil into the mine workings

through collapsed areas (piping), and caused the upward progression of collapse features

towards the ground surface (Hoffman et al., 1995). Additionally, it was suspected that

water rushing out of the mine workings during dewatering could have removed roof fall

material that was supporting areas where roof failure had previously occurred (ODOT,

1995b). Other factors that are suspected to have possibly contributed to previous

subsidence in the study area were the rotting and weakening with time of timber supports

within the mine workings, and the heavy interstate travel of more than an average of

25,000 vehicles per day.

Although it is possible that pillars have been crushed or have punched into

underlying material within the study area (punching of the Upper Freeport Coal into

weak underclay has previously been observed in the Pittsburgh, Pennsylvania region),

previous studies and data acquired to date indicate that this has not likely occurred in the

study area, and that study area subsidence mechanisms can be predominantly

259
characterized as pit-type (Figure A.8b). When the collapse or punching of a single coal

pillar occurs, increasing loads are placed on adjacent pillars, and this typically will result

in the subsequent failure of additional pillars (Bell, 1999). As seen in Figure A.5

however, the lateral extent of previous surface collapse and roadway depression areas

have not spanned multiple coal pillars, and these areas have been relatively localized in

terms of the previously discussed sag and pit types of subsidence.

Although some uncertainty exists in regards to the exact locations of the mine

workings relative to the I-70 roadway (as previously discussed), the locations of previous

collapse features and roadway depressions correlate fairly well with mine room and

haulage way locations that have been mapped (Figure A.5). These observations serve as

further evidence that past subsidence events in the study area have occurred due to the

mine roof rock collapsing between stable coal pillars. These observations also suggest

that possible ongoing subsidence processes in the study area would be occurring as the

result of similar pit-type subsidence mechanisms, and that the lateral extent of any future

collapses of the surface would most likely be at most on the order of the width of mine

rooms where they are crossed by the roadway.

260
APPENDIX B

PROGRAM FOR APPLICATION OF EQUATIONS DESCRIBING


ELASTIC WAVE SCATTERING FROM PLANAR INTERFACES

B.1 Introduction

A computer program (PSHSV) was written that calculates and plots displacement

amplitude coefficients, normalized square root energy ratios, energy coefficients, and

phase changes for plane P- SH- or SV-waves incident on a planar interface between two

homogeneous and isotropic solids. In subsequent sections of this appendix, functionality

of the program is discussed and demonstrated, and the program source code is listed.

Code such as that which is presented has not previously been made publicly available.

The source code will be available in digital form in the public domain (www.iamg.org) as

part of a paper that has been accepted (during 2002) for publication in the journal

Computers and Geosciences (see vita section of this dissertation for the paper reference).

The complexity of the equations describing seismic amplitude and energy

partitioning has led to numerous papers and books containing erroneous results and

misprints, and many of these mistakes have been recognized and discussed (Gutenburg,

1944; Singh et al., 1970; Hales and Roberts, 1974; Young and Braile, 1976; Denham and

Palmeira, 1984). Additional misprints related to these equations that have not previously

been reported in the literature were recognized during the development of the computer

261
code presented in this appendix. In Table 3.1 of Lay and Wallace (1995), the equation

for variable b is in error. The equation should read: b = ρ 2(1-2β22p2) + 2ρ1β12p2 and not:

b = ρ 2(1-2β22p2) - 2ρ1β12p2 as it is shown. Additionally, in Figure 3.28 of Lay and

Wallace (1995), the y-axis reads: Ai / A0 = (E / Ei)1/2, and this is incorrect for three of the

four curves that are plotted. The quantities plotted in the figure are the magnitudes of the

displacement amplitude coefficients, and only for the reflected P-wave is the plotted

curve equivalent to the square root energy curve that would be plotted. The y-axis should

therefore read: Ai / A0 in the figure. One of the energy coefficient plots in Figure 2 (for

reflected P-wave energy) presented in Baker (1998) also appears to be in error (input

media parameters for the two cases shown in this figure were also reversed).

The abundance of errors in previous refereed publications related to seismic

amplitude and energy-partitioning equations, demonstrates the importance of making

publicly available the program that was developed and presented in this appendix.

B.1.1 Solutions Obtained From Program Application

The developed program was rigorously tested against solutions previously

published (and not subsequently found to be in error) in refereed literature. Displacement

amplitude coefficient solutions were consistent with those for reflected and refracted P-

and SV-waves from incident SV-waves in Costain et al. (1963), reflected and refracted P-

and SV-waves from incident P-waves in Lay and Wallace (1995), reflected and refracted

SH-waves from incident SH-waves in Shearer (1999), and reflected P- and SV-waves

from incident P-waves in Brown (2000). Displacement amplitude coefficient solutions

were also consistent with those for reflected and refracted P- and SV-waves from incident
262
P-waves in Koefoed (1962). The sign of refracted SV-wave coefficients was opposite of

those calculated by Koefoed (1962) however, and this was simply due to opposite

specified conventions for positive refracted SV-wave particle motion direction.

Square root energy coefficient solutions were consistent with those for reflected

and refracted P- SH- and SV-waves from incident P- SH- and SV-waves in Gutenberg

(1944), and reflected and refracted P- SH- and SV-waves from incident SH- and SV-

waves in Crampin (1987). Energy coefficient solutions were consistent with those for

reflected and refracted P- and SV-waves from incident P-waves in Muskat and Meres

(1940), reflected and refracted P- and SV-waves from incident SV-waves in Costain et al.

(1963), reflected and refracted P- and SV-waves from incident P-waves in Tooley et al.

(1965), and reflected and refracted P- and SV-waves from incident P- and SV-waves in

Young and Braile (1976).

B.2 Program Description

The computer program (PSHSV) presented in this appendix was written in

MATLAB (version 6, release 12). Upon execution, the user is prompted to enter media

P- and S-wave velocities and densities. After user-specified parameters are input, exact

solution displacement amplitude coefficients, normalized square root energy ratios,

energy coefficients, and phase changes for plane P- SH- or SV-waves incident on a

planar interface between two homogeneous and isotropic solids are calculated and plotted

as a function of incident angle.

The program was developed using the equations presented in Chapter 2, which

were primarily from, and derived or modified from, those presented by Aki and Richards
263
(1980), Sheriff and Geldart (1982), and Shearer (1999). The signs of calculated

displacement amplitude coefficients depend upon the specified directions of positive

particle displacement in equations. The direction of wave propagation was specified as

the direction of positive polarization for P-waves, and the direction towards the interface

between incident and refracted media was specified as the direction of positive

polarization for SV-waves (SV-wave particle motion is perpendicular to the propagation

direction and in the plane containing the SV-wave ray). The nomenclature used for user-

input parameters, and for variables within the source code, is consistent with that used in

Chapter 2.

B.2.1 Critical Angles and Phase

When the velocity of the incident medium is less than that of the reflected or

refracted medium, the reflected or refracted angle (ș2) is equivalent to 90 degrees for a

certain incident angle (ș1), which is known as the critical angle (șc). In terms of ș1, the

incident medium velocity (V1), and the reflected or refracted medium velocity (V2), ș2 can

be expressed from Snell’s law as:

ș2 = sin-1((sinș1)(V2 / V1)) (B.1)

and a critical angle (șc) is defined as:

șc = sin-1(V1 / V2) (B.2)

For each common-mode component (i.e. P-to-P, SH-to-SH, SV-to-SV) there is

one possible critical angle to consider, which is the incident angle at which total internal

reflection (in medium 1, the incident medium) is produced, and beyond which no

refracted energy is transmitted normal to the interface into medium 2. An evanescent

264
wave (with no vertical energy flux) will continue to propagate along (parallel to) and

decay exponentially away from the interface post-critical angle. For converted wave

components (i.e. P-to-SV, SV-to-P) additional critical angles must be considered. A

possible incident angle may result in the angle of a mode-converted refracted wave being

equivalent to 90 degrees (for example, when the incident SV-wave velocity is less than

the velocity of a refracted P-wave). An SV-wave incident angle may also result in the

angle of a mode-converted reflected P-wave being equivalent to 90 degrees. Considering

all components, possible critical angles for all cases are given in the following set of

equations:

P1P2șc = (sin-1(Vp1 / Vp2))

P1SV2șc = (sin-1(Vp1 / Vs2))

SH1SH2șc = (sin-1(Vs1 / Vs2))

SV1P1șc = (sin-1(Vs1 / Vp1))

SV1SV2șc = (sin-1(Vs1 / Vs2))

SV1P2șc = (sin-1(Vs1 / Vp2)) (B.3)

Where P1P2șc is a critical angle associated with a refracted P-wave from an incident P-

wave, P1SV2șc is a critical angle associated with a refracted SV-wave from an incident P-

wave, SH1SH2șc is a critical angle associated with a refracted SH-wave from an incident

SH-wave, SV1P1șc is a critical angle associated with a reflected P-wave from an incident

SV-wave, SV1SV2șc is a critical angle associated with a refracted SV-wave from an

incident SV-wave, and SV1P2șc is a critical angle associated with a refracted P-wave

from an incident SV-wave.

265
For angles of incidence greater than critical angles, complex values of ș2 (ray

angles of reflected and refracted waves) may be used in order to satisfy Snell’s law,

calculated values of displacement amplitude coefficients (reflection and refraction)

dependent on ș2 become complex, and variable phase changes occur. Phase angle

calculations made using the code consider both real and imaginary parts of complex

coefficients. The phase angle for a given wave type and incidence angle, is calculated as

the angle between +180 degrees and -180 degrees (with 0 degrees specified between

quadrants 1 and 4) whose tangent is the real part of (b / a), written as:

ij = tan-1 (b / a) (B.4)

Where ij is phase angle, and a and b are the real and imaginary parts respectively of a

given displacement amplitude coefficient.

B.3 Program Functionality

Functionality of the program is demonstrated through application to an example

gas reservoir detection problem (a type of problem commonly addressed by considering

amplitudes and reflection coefficients at different incidence angles to infer physical

material properties). Although bright spots (one type of hydrocarbon indicator) in P-

wave sections have been recognized as often being associated with gas presence in shale

and sand sequences, they can also be caused by geologic features other than gas, such as

carbonate or hard streaks, igneous intrusions, lignites or wet sands (Spratt et al., 1993).

Because the bright spot method cannot distinguish between gas and non-gas anomalies,

266
Lithology Vp (m / s) Vs (m / s) ρ (g / cm3)
Shale 3811 2263 2.4
Sand (without gas) 3811 2302 2.25
Sand (with gas) 3453 2302 2.1
Coal 2652 1280 1.27

Table B.1. Lithologic parameters used to obtain the solutions plotted as a function of
incidence angle in Figures B.1 - B.3.

more definitive methods, involving the analysis of amplitude variation with offset (Allen

and Peddy, 1993; Castagna, 1993; Hilterman, 2001), and the comparison of P- and S-

wave amplitudes (Ensley, 1984) have evolved for anomaly verification.

The program was applied to obtain solutions for three different geologic interface

scenarios: 1) an overlying shale and an underlying sand (no gas), 2) an overlying shale

and an underlying sand (with gas), and 3) an overlying shale and an underlying coal. P-

and S-wave velocities and densities that were used as program input parameters are given

in Table B.1, and were taken as representative values from Dey-Sakar and Svatek (1993).

Shown in Figure B.1, are the three figures that were generated upon program

execution using scenario 1 (shale to sand without gas) parameters. Figure B.1a contains

results from an incident P-wave, Figure B.1b contains results from an incident SH-wave,

and Figure B.1c contains results from an incident SV-wave. Each incident wave-type

figure contains six subplots that show how calculated displacement amplitude

coefficients, normalized square root energy ratios, energy coefficients, and phase angles

change as a function of incident angle, for each wave type generated at the interface. The
267
268
(a) Plots for an incident P-wave.

Figure B.1. Plots (generated with the PSHSV program) showing displacement amplitude, square root energy, and energy
coefficients, and phase changes as a function of incidence angle for a: (a) P-wave, (b) SH-wave, and (c) SV-wave incident on a
shale (medium 1) and sand (medium 2) interface (Table B.1). See text for discussion. (continued)
Figure B.1. (continued)

269
(b) Plots for an incident SH-wave. (continued)
Figure B.1. (continued)

270
(c) Plots for an incident SV-wave.
271
(a) Plots for an incident P-wave.

Figure B.2. Plots (generated with the PSHSV program) showing displacement amplitude, square root energy, and energy
coefficients, and phase changes as a function of incidence angle for a: (a) P-wave, (b) SH-wave, and (c) SV-wave incident on a
shale (medium 1) and gas sand (medium 2) interface (Table B.1). See text for discussion. (continued)
Figure B.2. (continued)

272
(b) Plots for an incident SH-wave. (continued)
Figure B.2. (continued)

273
(c) Plots for an incident SV-wave.
274
(a) Plots for an incident P-wave.

Figure B.3. Plots (generated with the PSHSV program) showing displacement amplitude, square root energy, and energy
coefficients, and phase changes as a function of incidence angle for a: (a) P-wave, (b) SH-wave, and (c) SV-wave incident on a
shale (medium 1) and coal (medium 2) interface (Table B.1). See text for discussion. (continued)
Figure B.3. (continued)

275
(b) Plots for an incident SH-wave. (continued)
Figure B.3. (continued)

276
(c) Plots for an incident SV-wave.
imaginary and real parts, and magnitude of displacement amplitude coefficients are

shown. When a coefficient is not complex the magnitude curve represents the absolute

value of the coefficient, and when a coefficient is complex the magnitude curve

represents the complex modulus of the coefficient. Figure B.2 shows figures generated

using scenario 2 (shale to sand with gas) parameters, and Figure B.3 shows figures

generated using scenario 3 (shale to coal) parameters.

Differences in P-wave reflection coefficient behavior as a function of incidence

angle, and in P- and S-wave reflection coefficients, for the three different geologic

scenarios can be observed through a comparison of the solutions presented in Figures

B.1, B.2 and B.3. The main objective of this appendix is to demonstrate program

functionality within the context of a potential application, and therefore only some

general differences between the solutions that were plotted in these figures are discussed.

For scenario 1, the P-wave velocity is unchanged and a small decrease in density

occurs across the shale and sand (without gas) interface. This results in a polarity

reversal and a low P-wave reflection (displacement amplitude) coefficient magnitude

(from an incident P-wave) at normal incidence (Figure B.1a). For scenario 2 (when gas

occupies sand pores) a higher P-wave reflection coefficient magnitude (from an incident

P-wave) is observed at normal incidence (Figure B.2a), because bulk modulus values are

sensitive to, and can vary substantially with changes in pore fluid. P-wave velocity (Vp),

is dependent upon the bulk modulus (k), the shear modulus (µ), and density ( ρ ):

Vp = (k + 4/3µ / ρ )1/2 (B.5)

The introduction of gas (less stiff than water) into a water-saturated sand will cause a

considerable reduction in the bulk modulus, and therefore a reduction in P-wave velocity
277
(Domenico, 1976). Due to a greater contrast in P-wave velocity (and a slightly greater

contrast in density) at the shale and sand interface when gas as opposed to water occupies

sand pores, a larger normal incidence P-wave reflection coefficient magnitude results.

S-wave velocity (Vs) depends upon the shear modulus and density:

Vs = (µ / ρ )1/2 (B.6)

The rigidity of gases and liquids (ideal) is the same (zero), and the introduction of

gas into a water-saturated sand therefore does not change the S-wave velocity appreciably

relative to the change in P-wave velocity. The difference between SH-wave normal

incidence reflection coefficients (from incident SH-waves) for scenarios 1 and 2 (Figures

B.1b and B.2b) is therefore relatively small when compared to the observed difference in

P-wave reflection coefficients between scenarios 1 and 2 (Figures B.1a and B.2a).

Koefoed (1955) recognized that when Poisson’s ratio differs substantially across a

geologic interface, large changes in P-wave reflection coefficients at the interface as a

function of incidence angle can occur, and that these variations can provide insight

regarding lithology. Poisson’s ratio (ı) for a given medium can be defined as:

ı = ((Vp2 / 2Vs2) - 1) / ((Vp2 / Vs2) - 1) (B.7)

Ostrander (1984) first applied the concept of using reflected P-wave amplitude variations

with incidence angle for gas-sand detection purposes, demonstrating that differences in

amplitude partitioning as a function of offset between gas-sand reflections and other

types of reflections may be observed.

In scenario 1, density and Poisson’s ratio decrease only slightly across the shale

and sand (without gas) interface, and there is little variation in the P-wave amplitude

curve with increasing incidence angle (Figure B.1a). This amplitude behavior as a

278
function of incidence angle is different than that observed for scenario 2. In scenario 2 a

considerable decrease in both compressional velocity and Poisson’s ratio occur across the

shale and sand (with gas) interface, which causes a noticeable increase in P-wave

reflection coefficient magnitude with increasing incidence angle (Figure B.2a). These

observations are consistent with a general conclusion that has been made regarding AVO

by Allen and Peddy (1993): when compressional impedance and Poisson’s ratio change

substantially in the same direction across an interface, an increase in P-wave reflection

coefficient magnitude with offset will result.

Important differences in reflected P-wave amplitude behavior with offset are also

evident between scenario 2 and scenario 3 results. In scenario 3, a substantial decrease in

compressional impedance occurs across the shale and coal interface, while Poisson’s ratio

changes in the opposite direction and increases across the interface. This results in the

observed rapid decrease in reflection coefficient magnitude with increasing incidence

angle from normal (Figure B.3a). This type of difference in P-wave amplitude behavior

with offset can often be used to distinguish high amplitude anomalies caused by

lithologic changes from those caused by gas accumulation (Dey-Sarkar and Svatek,

1993). Consideration of the SH-wave reflection coefficients (resulting from incident SH-

waves) at normal incidence can also be useful for distinguishing lithologic- and gas-

related anomalies. The S-wave velocity of sand is not affected by the addition of gas, and

the SH-wave reflection coefficient at normal incidence is therefore small relative to the

normal incidence P-wave reflection coefficient in scenario 2 (Figures B.2a and B.2b).

The lithologic change from shale to coal across the interface in scenario 3 however

279
results in a high shear impedance contrast across the interface, and therefore a higher SH-

reflection coefficient at normal incidence than in scenario 2 (Figures B.2b and B.3b).

B.4 Source Code Listing

The PSHSV computer program (MATLAB code) is listed next in this appendix:

% pshsv.m - written and last modified by Erich D. Guy on 3/25/02 %%%%%%%%%%


% Program computes the real, imaginary, and magnitude of reflection and refraction/
% transmission (amplitude) coefficients along with phase, square root energy
% coefficients, and energy coefficients for elastic plane-wave scattering from a welded
% planar interface (between isotropic solids), for P-, SH-, and SV-waves.
% User input = P-wave and S-wave velocity and (bulk) density values.
% Copyright (c) Erich D. Guy, 2002, all rights reserved. Permission to use program
% for academic or research purposes is authorized, provided the author is cited.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

clear all; close all;

% User Input
Vp1 = input ('Enter Vp1 (Incident Medium P-Wave Velocity): ');
Vp2 = input ('Enter Vp2 (Refracted Medium P-Wave Velocity): ');
Vs1 = input ('Enter Vs1 (Incident Medium S-Wave Velocity): ');
Vs2 = input ('Enter Vs2 (Refracted Medium S-Wave Velocity): ');
Rho1 = input ('Enter Rho1 (Incident Medium Density): ');
Rho2 = input ('Enter Rho2 (Refracted Medium Density): ');

thetadeg = 0.1:.1:89.99; % Incident angles in degrees


deg = pi/180; rad = 180/pi;
thetarad = thetadeg*deg;
xx = size(thetarad);
p = sin(thetarad)/Vp1; % Horizontal slowness (for incident P-wave)
pp = sin(thetarad)/Vs1; % Horizontal slowness (for incident S-wave)
Zp1 = Vp1*Rho1; Zs1 = Vs1*Rho1; Zp2 = Vp2*Rho2; Zs2 = Vs2*Rho2; % Impedances

% Reflection and refraction angle cosines


costhetaP1P1 = Vp1*(sqrt((1/(Vp1^2))-(p.^2)));
costhetaP1SV1 = Vs1*(sqrt((1/(Vs1^2))-(p.^2)));
costhetaP1P2 = Vp2*(sqrt((1/(Vp2^2))-(p.^2)));
costhetaP1SV2 = Vs2*(sqrt((1/(Vs2^2))-(p.^2)));
costhetaSH1SH1 = Vs1*(sqrt((1/(Vs1^2))-(pp.^2)));
costhetaSH1SH2 = Vs2*(sqrt((1/(Vs2^2))-(pp.^2)));
280
costhetaSV1SV1 = Vs1*(sqrt((1/(Vs1^2))-(pp.^2)));
costhetaSV1SV2 = Vs2*(sqrt((1/(Vs2^2))-(pp.^2)));
costhetaSV1P1 = Vp1*(sqrt((1/(Vp1^2))-(pp.^2)));
costhetaSV1P2 = Vp2*(sqrt((1/(Vp2^2))-(pp.^2)));

% Critical angles
P1P2thetac = (asin(Vp1/Vp2));
P1SV2thetac = (asin(Vp1/Vs2));
SH1SH2thetac = (asin(Vs1/Vs2));
SV1P1thetac = (asin(Vs1/Vp1));
SV1SV2thetac = (asin(Vs1/Vs2));
SV1P2thetac = (asin(Vs1/Vp2));

% SH intromission angle
thetaI = (asin(sqrt ((((1))-((Zs1/Zs2)^2))/(((Vs2/Vs1)^2)-((Zs1/Zs2)^2))))*rad);

% Amplitude coefficient variables and cosine-dependent terms


a = Rho2*(1-2*Vs2^2*p.^2)-Rho1*(1-2*Vs1^2*p.^2);
aa = Rho2*(1-2*Vs2^2*pp.^2)-Rho1*(1-2*Vs1^2*pp.^2);
b = Rho2*(1-2*Vs2^2*p.^2)+2*Rho1*Vs1^2*p.^2;
bb = Rho2*(1-2*Vs2^2*pp.^2)+2*Rho1*Vs1^2*pp.^2;
c = Rho1*(1-2*Vs1^2*p.^2)+2*Rho2*Vs2^2*p.^2;
cc = Rho1*(1-2*Vs1^2*pp.^2)+2*Rho2*Vs2^2*pp.^2;
d = 2*(Rho2*Vs2^2-Rho1*Vs1^2);
dd = 2*(Rho2*Vs2^2-Rho1*Vs1^2);
E = b.*(costhetaP1P1/Vp1)+c.*(costhetaP1P2/Vp2);
EE = bb.*(costhetaSV1P1/Vp1)+cc.*(costhetaSV1P2/Vp2);
F = b.*(costhetaP1SV1/Vs1)+c.*(costhetaP1SV2/Vs2);
FF = bb.*(costhetaSV1SV1/Vs1)+cc.*(costhetaSV1SV2/Vs2);
G = a-d*(costhetaP1P1/Vp1).*(costhetaP1SV2/Vs2);
GG = aa-dd*(costhetaSV1P1/Vp1).*(costhetaSV1SV2/Vs2);
H = a-d*(costhetaP1P2/Vp2).*(costhetaP1SV1/Vs1);
HH = aa-dd*(costhetaSV1P2/Vp2).*(costhetaSV1SV1/Vs1);
D = E.*F+G.*H.*p.^2;
DD = EE.*FF+GG.*HH.*pp.^2;

% Amplitude coefficients
P1P1 = ((b.*(costhetaP1P1/Vp1)-c.*(costhetaP1P2/Vp2)).*F
(a+d*(costhetaP1P1/Vp1).*(costhetaP1SV2/Vs2)).*H.*p.^2)./D;

P1SV1 = (2*(costhetaP1P1/Vp1).*(a.*b+c*d.*(costhetaP1P2/Vp2).*
(costhetaP1SV2/Vs2)).*p*Vp1/Vs1)./D;

P1P2 = (2*Rho1*(costhetaP1P1/Vp1).*F*(Vp1/Vp2))./D;

281
P1SV2 = (2*Rho1*(costhetaP1P1/Vp1).*H.*p*(Vp1/Vs2))./D;

SH1SH1 = (Rho1*Vs1*costhetaSH1SH1-Rho2*Vs2*costhetaSH1SH2)./
(Rho1*Vs1*costhetaSH1SH1+Rho2*Vs2*costhetaSH1SH2);

SH1SH2 = (2*Rho1*Vs1*costhetaSH1SH1)./
(Rho1*Vs1*costhetaSH1SH1+Rho2*Vs2*costhetaSH1SH2);

SV1SV1 = -((bb.*(costhetaSV1SV1/Vs1)-cc.*(costhetaSV1SV2/Vs2)).*EE-
(aa+dd.*(costhetaSV1P2/Vp2).*(costhetaSV1SV1/Vs1)).*GG.*pp.^2)./DD;

SV1P1 = -(2*(costhetaSV1SV1/Vs1).*(aa.*bb+cc*dd.*((costhetaSV1P2)/Vp2).*
((costhetaSV1SV2)/Vs2)).*pp*(Vs1/Vp1))./DD;

SV1SV2 = 2*Rho1*(costhetaSV1SV1/Vs1).*EE*(Vs1/Vs2)./DD;

SV1P2 = -2*(Rho1*(costhetaSV1SV1/Vs1).*GG.*pp*(Vs1/Vp2))./DD;

% Square root energy coefficients


ENP1P1 = abs(P1P1);

ENP1SV1 = abs(P1SV1).*(sqrt((Vs1.*costhetaP1SV1)./(Vp1.*costhetaP1P1)));

ENP1P2 = abs(P1P2).*(sqrt((Rho2*Vp2.*costhetaP1P2)./(Rho1*Vp1.*costhetaP1P1)));
for jj = 1:xx(1,2); if thetarad(:,jj)<P1P2thetac; ENP1P2(:,jj) = ENP1P2(:,jj);
elseif thetarad(:,jj)>=P1P2thetac; ENP1P2(:,jj) = 0; end; end;

ENP1SV2 = abs(P1SV2).*(sqrt((Rho2*Vs2.*costhetaP1SV2)./
(Rho1*Vp1.*costhetaP1P1)));
for jj = 1:xx(1,2); if thetarad(:,jj)<P1SV2thetac; ENP1SV2(:,jj) = ENP1SV2(:,jj);
elseif thetarad(:,jj)>=P1SV2thetac; ENP1SV2(:,jj) = 0; end; end;

ENSH1SH1 = abs(SH1SH1);

ENSH1SH2 = abs(SH1SH2).*(sqrt((Rho2*Vs2*costhetaSH1SH2)./
(Rho1*Vs1*costhetaSH1SH1)));
for jj = 1:xx(1,2); if thetarad(:,jj)<SH1SH2thetac; ENSH1SH2(:,jj) = ENSH1SH2(:,jj);
elseif thetarad(:,jj)>=SH1SH2thetac; ENSH1SH2(:,jj) = 0; end; end;

ENSV1SV1 = abs(SV1SV1);

ENSV1P1 = abs(SV1P1).*(sqrt((Vp1.*costhetaSV1P1)./(Vs1.*costhetaSV1SV1)));
for jj = 1:xx(1,2); if thetarad(:,jj)<SV1P1thetac; ENSV1P1(:,jj) = ENSV1P1(:,jj);
elseif thetarad(:,jj)>=SV1P1thetac; ENSV1P1(:,jj) = 0; end; end;

282
ENSV1SV2 = abs(SV1SV2).*(sqrt((Rho2*Vs2.*costhetaSV1SV2)./
(Rho1*Vs1.*costhetaSV1SV1)));
for jj = 1:xx(1,2); if thetarad(:,jj)<SV1SV2thetac; ENSV1SV2(:,jj) = ENSV1SV2(:,jj);
elseif thetarad(:,jj)>=SV1SV2thetac; ENSV1SV2(:,jj) = 0; end; end;

ENSV1P2 = abs(SV1P2).*(sqrt((Rho2*Vp2.*costhetaSV1P2 )./


(Rho1*Vs1.*costhetaSV1SV1)));
for jj = 1:xx(1,2); if thetarad(:,jj)<SV1P2thetac; ENSV1P2(:,jj) = ENSV1P2(:,jj);
elseif thetarad(:,jj)>=SV1P2thetac; ENSV1P2(:,jj) = 0; end; end;

% Energy coefficients
EP1P1 = (ENP1P1).^2; EP1SV1 = (ENP1SV1).^2; EP1P2 = (ENP1P2).^2;
EP1SV2 = (ENP1SV2).^2; ESH1SH1 = (ENSH1SH1).^2; ESH1SH2 = (ENSH1SH2).^2;
ESV1SV1 = (ENSV1SV1).^2; ESV1P1 = (ENSV1P1).^2; ESV1SV2 = (ENSV1SV2).^2;
ESV1P2 = (ENSV1P2).^2;

% Plot solutions
set(0,'Units','normal');
nb = [0,0,.5];
ss = get(0,'ScreenSize');
pos = [ss(3)/2-0.8/2,ss(4)/2-0.6/2,.8,.6];

% Incident SV-wave
figure('Units','normal','Position',pos,'Name','Incident SV-Wave In Medium
1','NumberTitle','off');

subplot(2,3,1);
plot(thetadeg,imag(SV1SV1),'b-',thetadeg,imag(SV1P1),'k--',thetadeg,imag(SV1SV2),'r-
.',thetadeg,imag(SV1P2),'m:','LineWidth',1.5);
title('Amplitude Coefficients (Imaginary)'); ylabel('A / A_0');
legend('Reflected SV','Reflected P','Refracted SV','Refracted P');
set(legend,'position',[.001 .003 .113 .112]);
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,2);
plot(thetadeg,real(SV1SV1),'b-',thetadeg,real(SV1P1),'k--',thetadeg,real(SV1SV2),'r-
.',thetadeg,real(SV1P2),'m:','LineWidth',1.5);
title('Amplitude Coefficients (Real)'); ylabel('A / A_0');
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,3);

283
plot(thetadeg,abs(SV1SV1),'b-',thetadeg,abs(SV1P1),'k--',thetadeg,abs(SV1SV2),'r-
.',thetadeg,abs(SV1P2),'m:','LineWidth',1.5);
title('Amplitude Coefficients (Magnitude)'); ylabel('A / A_0');
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,4);
plot(thetadeg,abs(ENSV1SV1),'b-',thetadeg,abs(ENSV1P1),'k--
',thetadeg,abs(ENSV1SV2),'r-.',thetadeg,abs(ENSV1P2),'m:','LineWidth',1.5);
title('Square Root Energy Coefficients'); xlabel('Incidence Angle (degrees)'); ylabel('(E /
E_0) ^1^/^2');
axis([0 90 -0.05 1.05]); set(gca,'XTick',0:15:90); set(gca,'YTick',0:.25:1); grid on;

subplot(2,3,5);
plot(thetadeg,abs(ESV1SV1),'b-',thetadeg,abs(ESV1P1),'k--',thetadeg,abs(ESV1SV2),'r-
.',thetadeg,abs(ESV1P2),'m:','LineWidth',1.5);
title('Energy Coefficients'); xlabel('Incidence Angle (degrees)'); ylabel('E / E_0');
axis([0 90 -0.05 1.05]); set(gca,'XTick',0:15:90); set(gca,'YTick',0:.25:1); grid on;

subplot(2,3,6);
plot(thetadeg,angle(SV1SV1)*rad,'b-',thetadeg,angle(SV1P1)*rad,'k--
',thetadeg,angle(SV1SV2)*rad,'r-.',thetadeg,angle(SV1P2)*rad,'m:','LineWidth',1.5);
title('Phase Changes'); xlabel('Incidence Angle (degrees)'); ylabel('Phase Angle
(degrees)');
axis([0 90 -200 200]); set(gca,'XTick',0:15:90); set(gca,'YTick',-180:90:180); grid on;
text(-290,812,'Incident SV-Wave In Medium 1','fontweight','bold','color',nb);
L = num2str(Vp1); text(-290,740,'Vp_1 =','color',nb); text(-273.5,740,(L),'color',nb);
M = num2str(Vp2); text(-290,700,'Vp_2 =','color',nb); text(-273.5,700,(M),'color',nb);
N = num2str(Vs1); text(-290,660,'Vs_1 =','color',nb); text(-273.5,660,(N),'color',nb);
O = num2str(Vs2); text(-290,620,'Vs_2 =','color',nb); text(-273.5,620,(O),'color',nb);
P = num2str(Rho1); text(-290,580,'\rho_1 =','color',nb); text(-273.5,580,(P),'color',nb);
Q = num2str(Rho2); text(-290,540,'\rho_2 =','color',nb); text(-273.5,540,(Q),'color',nb);
R = num2str(Zp1); text(-290,460,'Zp_1 =','color',nb); text(-273.5,460,(R),'color',nb);
S = num2str(Zp2); text(-290,420,'Zp_2 =','color',nb); text(-273.5,420,(S),'color',nb);
T = num2str(Zs1); text(-290,380,'Zs_1 =','color',nb); text(-273.5,380,(T),'color',nb);
U = num2str(Zs2); text(-290,340,'Zs_2 =','color',nb); text(-273.5,340,(U),'color',nb);
V = num2str(real(SV1P1thetac)*rad); text(-290,260,'\theta_c Reflected P =','color',nb);
text(-277,230,(V),'color',nb);
W = num2str(real(SV1SV2thetac)*rad);
text(-290,200,'\theta_c Refracted SV =','color',nb); text(-277,170,(W),'color',nb);
X = num2str(real(SV1P2thetac)*rad); text(-290,140,'\theta_c Refracted P =','color',nb);
text(-277,110,(X),'color',nb);
WW = num2str(Vp1/Vs1);
text(-290,50,'Vp_1/Vs1 =','color',nb); text(-285,10,(WW),'color',nb);
XX = num2str(Vp2/Vs2); text(-290,-30,'Vp_2/Vs1 =','color',nb);

284
text(-285,-70,(XX),'color',nb);
Y = num2str(Sigma1); text(-290,-110,'\sigma_1 =','color',nb);
text(-273.5,-110,(Y),'color',nb);
Z = num2str(Sigma2); text(-290,-150,'\sigma_2 =','color',nb); text(-273.5,-
150,(Z),'color',nb);

% Incident SH-wave
figure('Units','normal','Position',pos,'Name','Incident SH-Wave In Medium
1','NumberTitle','off');

subplot(2,3,1);
plot(thetadeg,imag(SH1SH1),'b-',thetadeg,imag(SH1SH2),'r-.','LineWidth',1.5);
title('Amplitude Coefficients (Imaginary)'); ylabel('A / A_0');
legend('Reflected SH','Refracted SH'); set(legend,'position',[.001 .003 .113 .067]);
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,2);
plot(thetadeg,real(SH1SH1),'b-',thetadeg,real(SH1SH2),'r-.','LineWidth',1.5);
title('Amplitude Coefficients (Real)'); ylabel('A / A_0');
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,3);
plot(thetadeg,abs(SH1SH1),'b-',thetadeg,abs(SH1SH2),'r-.','LineWidth',1.5);
title('Amplitude Coefficients (Magnitude)'); ylabel('A / A_0');
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,4);
plot(thetadeg,abs(ENSH1SH1),'b-',thetadeg,abs(ENSH1SH2),'r-.','LineWidth',1.5);
title('Square Root Energy Coefficients'); xlabel('Incidence Angle (degrees)'); ylabel('(E /
E_0) ^1^/^2');
axis([0 90 -0.05 1.05]); set(gca,'XTick',0:15:90); set(gca,'YTick',0:.25:1); grid on;

subplot(2,3,5);
plot(thetadeg,abs(ESH1SH1),'b-',thetadeg,abs(ESH1SH2),'r-.','LineWidth',1.5);
title('Energy Coefficients'); xlabel('Incidence Angle (degrees)'); ylabel('E / E_0');
axis([0 90 -0.05 1.05]); set(gca,'XTick',0:15:90); set(gca,'YTick',0:.25:1); grid on;

subplot(2,3,6);
plot(thetadeg,angle(SH1SH1)*rad,'b-',thetadeg,angle(SH1SH2)*rad,'r-.','LineWidth',1.5);
title('Phase Changes'); xlabel('Incidence Angle (degrees)'); ylabel('Phase Angle
(degrees)');

285
axis([0 90 -200 200]); set(gca,'XTick',0:15:90); set(gca,'YTick',-180:90:180); grid on;
text(-290,812,'Incident SH-Wave In Medium 1','fontweight','bold','color',nb);
N = num2str(Vs1); text(-290,660,'Vs_1 =','color',nb); text(-273.5,660,(N),'color',nb);
O = num2str(Vs2); text(-290,620,'Vs_2 =','color',nb); text(-273.5,620,(O),'color',nb);
P = num2str(Rho1); text(-290,580,'\rho_1 =','color',nb); text(-273.5,580,(P),'color',nb);
Q = num2str(Rho2); text(-290,540,'\rho_2 =','color',nb); text(-273.5,540,(Q),'color',nb);
T = num2str(Zs1); text(-290,380,'Zs_1 =','color',nb); text(-273.5,380,(T),'color',nb);
U = num2str(Zs2); text(-290,340,'Zs_2 =','color',nb); text(-273.5,340,(U),'color',nb);
V = num2str(real(SH1SH2thetac)*rad);
text(-290,260,'\theta_c Refracted SH =','color',nb); text(-277,230,(V),'color',nb);
W = num2str(thetaI); text(-290,190,'\theta_I =','color',nb); text(-277,190,(W),'color',nb);
XX = num2str(Vs2/Vs1); text(-290,-130,'Vs_2/Vs_1 =','color',nb);
text(-263,-130,(XX),'color',nb);
YY = num2str(Zs1/Zs2); text(-290,-170,'Zs_1/Zs_2 =','color',nb);
text(-263,-170,(YY),'color',nb);
ZZ = num2str(Rho1/Rho2); text(-290,-210,'\rho_1/\rho_2 =','color',nb);
text(-263,-210,(ZZ),'color',nb);

% Incident P-wave
figure('Units','normal','Position',pos,'Name','Incident P-Wave In Medium
1','NumberTitle','off')

subplot(2,3,1);
plot(thetadeg,imag(P1P1),'b-',thetadeg,imag(P1SV1),'k--',thetadeg,imag(P1P2),'r-
.',thetadeg,imag(P1SV2),'m:','LineWidth',1.5);
title('Amplitude Coefficients (Imaginary)'); ylabel('A / A_0');
legend('Reflected P','Reflected SV','Refracted P','Refracted SV');
set(legend,'position',[.001 .003 .113 .112]);
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,2);
plot(thetadeg,real(P1P1),'b-',thetadeg,real(P1SV1),'k--',thetadeg,real(P1P2),'r-
.',thetadeg,real(P1SV2),'m:','LineWidth',1.5);
title('Amplitude Coefficients (Real)'); ylabel('A / A_0');
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

subplot(2,3,3);
plot(thetadeg,abs(P1P1),'b-',thetadeg,abs(P1SV1),'k--',thetadeg,abs(P1P2),'r-
.',thetadeg,abs(P1SV2),'m:','LineWidth',1.5);
title('Amplitude Coefficients (Magnitude)'); ylabel('A / A_0');
ylimits = get(gca,'Ylim'); ymin = ylimits(1)-.05; ymax = ylimits(2)+.05;
axis([0 90 ymin ymax]); set(gca,'XTick',0:15:90); grid on;

286
subplot(2,3,4);
plot(thetadeg,abs(ENP1P1),'b-',thetadeg,abs(ENP1SV1),'k--',thetadeg,abs(ENP1P2),'r-
.',thetadeg,abs(ENP1SV2),'m:','LineWidth',1.5);
title('Square Root Energy Coefficients'); xlabel('Incidence Angle (degrees)'); ylabel('(E /
E_0) ^1^/^2');
axis([0 90 -0.05 1.05]); set(gca,'XTick',0:15:90); set(gca,'YTick',0:.25:1); grid on;

subplot(2,3,5);
plot(thetadeg,abs(EP1P1),'b-',thetadeg,abs(EP1SV1),'k--',thetadeg,abs(EP1P2),'r-
.',thetadeg,abs(EP1SV2),'m:','LineWidth',1.5);
title('Energy Coefficients'); xlabel('Incidence Angle (degrees)'); ylabel('E / E_0');
axis([0 90 -0.05 1.05]); set(gca,'XTick',0:15:90); set(gca,'YTick',0:.25:1); grid on;

subplot(2,3,6);
plot(thetadeg,angle(P1P1)*rad,'b-',thetadeg,angle(P1SV1)*rad,'k--
',thetadeg,angle(P1P2)*rad,'r-.',thetadeg,angle(P1SV2)*rad,'m:','LineWidth',1.5);
title('Phase Changes'); xlabel('Incidence Angle (degrees)'); ylabel('Phase Angle
(degrees)');
axis([0 90 -200 200]); set(gca,'XTick',0:15:90); set(gca,'YTick',-180:90:180); grid on;
text(-290,812,'Incident P-Wave In Medium 1','fontweight','bold','color',nb);
L = num2str(Vp1); text(-290,740,'Vp_1 =','color',nb); text(-273.5,740,(L),'color',nb);
M = num2str(Vp2); text(-290,700,'Vp_2 =','color',nb); text(-273.5,700,(M),'color',nb);
N = num2str(Vs1); text(-290,660,'Vs_1 =','color',nb); text(-273.5,660,(N),'color',nb);
O = num2str(Vs2); text(-290,620,'Vs_2 =','color',nb); text(-273.5,620,(O),'color',nb);
P = num2str(Rho1); text(-290,580,'\rho_1 =','color',nb); text(-273.5,580,(P),'color',nb);
Q = num2str(Rho2); text(-290,540,'\rho_2 =','color',nb); text(-273.5,540,(Q),'color',nb);
R = num2str(Zp1); text(-290,460,'Zp_1 =','color',nb); text(-273.5,460,(R),'color',nb);
S = num2str(Zp2); text(-290,420,'Zp_2 =','color',nb); text(-273.5,420,(S),'color',nb);
T = num2str(Zs1); text(-290,380,'Zs_1 =','color',nb); text(-273.5,380,(T),'color',nb);
U = num2str(Zs2); text(-290,340,'Zs_2 =','color',nb); text(-273.5,340,(U),'color',nb);
V = num2str(real(P1P2thetac)*rad); text(-290,260,'\theta_c Refracted P =','color',nb);
text(-277,230,(V),'color',nb);
W = num2str(real(P1SV2thetac)*rad); text(-290,200,'\theta_c Refracted SV =','color',nb);
text(-277,170,(W),'color',nb);
WW = num2str(Vp1/Vs1); text(-290,50,'Vp_1/Vs1 =','color',nb);
text(-285,10,(WW),'color',nb);
XX = num2str(Vp2/Vs2); text(-290,-30,'Vp_2/Vs1 =','color',nb);
text(-285,-70,(XX),'color',nb);
Y = num2str(Sigma1); text(-290,-110,'\sigma_1 =','color',nb);
text(-273.5,-110,(Y),'color',nb);
Z = num2str(Sigma2); text(-290,-150,'\sigma_2 =','color',nb);
text(-273.5,-150,(Z),'color',nb);

% End program

287
APPENDIX C

INTERPRETED SH-WAVE DEPTH SECTIONS


FOR THE ENTIRE FIELD TEST AREA

C.1 Introduction

Stacked sections constructed using YY component (sometimes called crossline-

crossline, or SH-SH) reflection data acquired along the eastbound and westbound lanes of

I-70 (Appendix A, Figure C.1) are presented in this appendix. Discussion regarding the

acquisition and pre-processing of the data is presented in Chapter 3. Data processing and

imaging operations were performed using ProMAX software (Landmark Graphics

Corporation). The analysis flow applied to the data followed the processing, imaging,

and interpretation strategies previously developed for common-mode component

reflection information (Chapters 4 and 6). It was determined that subsurface areas where

subsidence processes have been active could be most accurately delineated through the

processing and interpretation of YY component data (Chapter 4).

C.2 Results and Interpretations

Processed time (uninterpreted) and depth (interpreted) YY component sections for

seismic lines acquired along I-70 between road stations 46694 and 48900 are shown in

Figures C.2 - C.5. (Note, road stations are in feet along the highway from the western
288
Guernsey county line. Highway engineers designate stationing as hundreds of feet plus

the remaining feet, for example, 489+00. In the text and the figures of this appendix

these numbers are combined, for example, 48900). YY component data from seismic

lines acquired along the eastbound travel and passing lanes (referred to as lines EBTravel

and EBPass respectively), and the westbound passing and travel lanes (referred to as lines

WBPass and WBTravel respectively), are shown in Figure C.2, Figure C.3, Figure C.4,

and Figure C.5 respectively.

Disrupted areas and offsets (normal faults) along the bedrock horizon that are

indicated on depth sections, are interpreted to have resulted from mine-related subsidence

activity. The locations and apparent dip directions of the bedrock horizon faults indicated

in Figures C.2, C.3, C.4, and C.5 are summarized in Table C.1, and are plotted relative to

the I-70 eastbound and westbound lanes in Figure C.6. The locations of bedrock horizon

disruptions interpreted from the processed seismic depth sections correlate well with the

locations of mine workings, observed roadway depressions and subsidence features, and

westbound lane land bridges that were approximately mapped relative to the I-70 road

lanes (Figure C.6). Seismic data interpretations are supported by the drill log information

(Chapters 4 and 6, and Appendix D) presented in this dissertation.

289
290
(a) Map of I-70 lanes (road stations 46700-47800) relative to YY component stacked section locations.

Figure C.1. Map view of I-70 eastbound and westbound lanes (EBTravel, EBPass, WBPass, and WBTravel) showing locations
of YY component stacked sections that are presented in this appendix relative to mapped mine workings, observed roadway
depressions and subsidence features, and westbound lane land bridges for road stations: (a) 46700-47800, and (b) 47800-48900.
Road stations are in feet along the highway from the western Guernsey County (Ohio) line. (continued)
Figure C.1. (continued)

291
(b) Map of I-70 lanes (road stations 47800-48900) relative to YY component stacked section locations.
292
(a) EBTravel YY component (stations 46700-47300).

Figure C.2. Eastbound travel lane (EBTravel) stacked time (uninterpreted) and depth (with bedrock horizon interpreted)
sections (YY component data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-48500, and (d) 48500-48900. CDP
locations (CDP_X) correspond to road stations (in feet from the western Guernsey County line). (continued)
Figure C.2. (continued)

293
(b) EBTravel YY component (stations 47300-47900). (continued)
Figure C.2. (continued)

294
(c) EBTravel YY component (stations 47900-48500). (continued)
Figure C.2. (continued)

(d) EBTravel YY component (stations 48500-48900).

295
296
(a) EBPass YY component (stations 46700-47300).

Figure C.3. Eastbound passing lane (EBPass) stacked time (uninterpreted) and depth (with bedrock horizon interpreted) sections
(YY component data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-48500, and (d) 48500-48900. CDP locations
(CDP_X) correspond to road stations (in feet along the highway from the western Guernsey County line). (continued)
Figure C.3. (continued)

297
(b) EBPass YY component (stations 47300-47900). (continued)
Figure C.3. (continued)

298
(c) EBPass YY component (stations 47900-48500). (continued)
Figure C.3. (continued)

(d) EBPass YY component (stations 48500-48900).

299
300
X
(a) WBPass YY component (stations 46700-47300).

Figure C.4. Westbound passing lane (WBPass) stacked time (uninterpreted) and depth (with bedrock horizon interpreted)
sections (YY component data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-48500, and (d) 48500-48900. CDP
locations (CDP_X) correspond to road stations (in feet along the highway from the western Guernsey County line). (continued)
Figure C.4. (continued)

301
X
(b) WBPass YY component (stations 47300-47900). (continued)
Figure C.4. (continued)

302
X
(c) WBPass YY component (stations 47900-48500). (continued)
Figure C.4. (continued)

(d) WBPass YY component (stations 48500-48900).

303
304
X
(a) WBTravel YY component (stations 46700-47300).

Figure C.5. Westbound travel lane (WBTravel) stacked time (uninterpreted) and depth (with bedrock horizon interpreted)
sections (YY component data) for CDP_X: (a) 46700-47300, (b) 47300-47900, (c) 47900-48500, and (d) 48500-48900. CDP
locations (CDP_X) correspond to road stations (in feet along the highway from the western Guernsey County line). (continued)
Figure C.5. (continued)

305
X
(b) WBTravel YY component (stations 47300-47900). (continued)
Figure C.5. (continued)

306
X
(c) WBTravel YY component (stations 47900-48500). (continued)
Figure C.5. (continued)

(d) WBTravel YY component (stations 48500-48900).

307
Eastbound travel lane (EBTravel) Eastbound pass lane (EBPass) Westbound pass lane (WBPass) Westbound travel lane (WBTravel)
Bedrock horizon fault location, Bedrock horizon fault location, Bedrock horizon fault location, Bedrock horizon fault location,
apparent dip direction apparent dip direction apparent dip direction apparent dip direction
46854, east 46951, east 46874, east 46818, east
46864, east 46970, west 46947, west 46830, east
46906, west 47055, east 47478, east 46896, west
46914, west 47071, west 47508, east 47461, east
46936, east 47117, west 47527, west 47554, west
46953, west 47235, west 47550, east 47568, west
47043, east 47252, west 47573, west 47588, west
47133, east 47316, east 47624, east 47665, east
47318, west 47341, west 47690, west 47690, west
47468, east 47399, east 47754, east 47703, east
47481, west 47411, west 47759, east 47719, west
47501, east 47467, east 47804, west 47756, east
47527, west 47790, east 47965, west 47848, west

308
47707, east 47806, west 48045, east 47883, east
47750, west 48216, east 48073, east 48119, west
47777, east 48236, west 48206, west 48264, west
47781, east 48329, east 48312, east 48396, east
47989, west 48354, west 48332, west 48519, east
48391, east 48542, east 48349, west
48596, east 48556, west 48498, west
48615, west 48602, east 48504, west
48643, west

Table C.1. Locations and apparent dip directions of bedrock horizon faults (mine-related) indicated on seismic YY component
depth sections (Figures C.2, C.3, C.4, and C.5). Fault locations (given in feet from the western county line) correspond to road
stations, and are plotted (with apparent dip directions indicated) relative to the roadway in Figure C.6.
309
(a) Map of I-70 lanes (road stations 46700-47800) relative to YY component-derived mine-related faults.

Figure C.6. Map view of I-70 eastbound and westbound lanes showing locations and apparent dip directions of YY component-
derived mine-related faults (Table C.1), relative to mapped mine workings, observed roadway depressions and subsidence
features, and westbound lane land bridges for road stations: (a) 46700-47800, and (b) 47800-48900. Road stations are in feet
from the western Guernsey County (Ohio) line. (continued)
Figure C.6. (continued)

310
(b) Map of I-70 lanes (road stations 47800-48900) relative to YY component-derived mine-related faults.
APPENDIX D

CROSS-HOLE RADAR EFFECTIVENESS FOR INVESTIGATION


OF SEISMICALLY IMAGED DISCONTINUITIES

D.1 Overview

Constant offset profile and multiple offset gather radar measurements were made

in the I-70 study area (Appendix A), using boreholes located near several coal mine-

related subsidence features that were imaged using S-wave seismic reflection (Chapter 6,

Appendix C). These data were acquired and analyzed to test the ability of cross-hole

radar methods for providing useful information for mine-related subsidence studies, and

to further investigate subsurface areas of the study area where S-wave seismic reflection

indicated that the bedrock horizon (above the mine workings) and overburden had been

disrupted from subsidence processes.

Radar data-derived plots of average EM-wave velocity and amplitude along with

processed velocity tomograms are shown to be useful for allowing the lateral and vertical

distribution of media between boreholes to be inferred. Relative low velocity zones

within fully saturated media between boreholes were found to correlate with decreased

direct wave amplitude, indicating increases in secondary porosity related to fracturing

and voids caused by mine-related subsidence activity. In many cases, such velocity and

amplitude trends are shown to correlate with fracture zones and voids that were
311
encountered during well construction, and with subsurface discontinuities that were

interpreted between boreholes from seismic data. Such correlations rule out the

possibility that these radar data anomalies resulted solely from location differences in

media mineralogy, media primary porosity, and/or groundwater specific conductance.

Cross-hole radar data are also shown to be useful for reducing uncertainty

regarding the locations and extent of mine rooms and coal pillars, and seismically imaged

subsidence features. In two cases, radar results indicate that seismically imaged

discontinuities are located between boreholes and directly beneath the seismic line,

whereas in two other cases radar data suggest that seismically imaged discontinuities are

located within close proximity to, but not between boreholes or directly beneath the

seismic line. Although the techniques measure different physical properties, recorded

wavelengths and processed images indicate that the resolution achieved using cross-hole

radar was slightly higher than that achieved using seismic reflection in the study area.

The results of this study demonstrate within the context of the I-70 mine-

subsidence project, that cross-hole radar surveying can provide useful information for

mine-related subsidence studies and insights into the nature and extent of fracturing and

void space within near-surface media between boreholes.

D.2 Introduction: Borehole Radar Methods

In many cases the detectable depth range of surface-located ground penetrating

radar (GPR) is restricted by either high wave attenuation in conductive near-surface

materials, volume scattering, or the presence of strong reflecting interfaces. The

utilization of boreholes permits radar antennas to be located closer to the subsurface


312
geology and potential anomalies of interest at depth. Demonstrated applications of

borehole radar include fracture detection (Olsson et al., 1992; Sato and Miwa, 2000),

cavity and tunnel detection (Haeni et al., 2002; Olhoeft, 1993), lithologic and structural

interpretation (Nickel et al., 1983; Dubois, 1995), and soil water content determination

(Gilson et al., 1996). The majority of documented successful borehole radar applications

for discontinuity detection/imaging have concerned data that were acquired in either

crystalline rock environments or under controlled experimental conditions. Although

borehole radar measurements can provide the potential for allowing physical property

changes associated with subsurface subsidence activity to be detected, to my knowledge

there have been no reports published concerning/demonstrating the applicability of

borehole radar techniques towards mine-related subsidence problems.

D.2.1 Antennas and Measurement Configurations

Borehole radar systems typically utilize dipole antennas that radiate

electromagnetic (EM) fields with the electric field vector components predominantly

oriented parallel to the long axis of the transmitter antenna. The radiation pattern of a

dipole antenna in a dry borehole with a homogeneous surrounding medium is similar to

that of a dipole in a whole space (Sato and Thierbach, 1991), however, changes in the

surrounding geologic medium and fluid filling the borehole can significantly alter the

pattern (Holliger and Bergmann, 2002). Antennas are oriented axially along a borehole

for measurements (only the co-polarized component is measured with commercial radar

systems), and can be considered as directional (can provide azimuthal resolution relative

to the borehole), or omni-directional (cannot directly provide azimuthal resolution).

313
Radar measurements can be made in a single borehole (single-hole) or between

boreholes (cross-hole). Single-hole measurements are typically made while raising or

lowering the antennas with a constant separation (reflection mode). Reflection

measurements can detect energy scattered from electrical property discontinuities above,

below, and away from the borehole. Useful information can also be obtained through

non-traditional variable offset sounding (VOS) conducted in a single borehole, or on the

ground surface using E- and H-plane antenna configurations (Guy and Radzevicius,

2001). VOS data are recorded while varying the spacing between antennas, and provide

a means of investigating propagation characteristics of radar signal.

Cross-hole (transmission) measurements between boreholes can be made by

raising or lowering transmitter and receiver antennas at the same rate to acquire a

constant offset profile (COP). A COP can provide information on bulk electrical

properties between boreholes by measuring variations in amplitude, period, and/or travel

time, and can also indicate discontinuity presence above, below, away from, or between

the boreholes by recording diffracted, reflected, and/or refracted energy. Measurements

can also be made by keeping a transmitter antenna at a fixed position in one hole and

lowering or raising a receiver antenna in another hole to acquire a multiple offset gather

(MOG). Numerous MOG’s made with the transmitter antenna located at different depths

can be merged and processed to create physical property tomograms (2D models for the

media plane between boreholes). Processed tomograms can allow possible borehole

radar anomalies to be spatially defined and isolated more accurately than processed COP

records. The number of traces that need to be acquired in order to generate tomograms

314
however is on the order of a hundred times greater than the number of traces that are

necessary to generate COP data plots.

Regardless of the borehole radar antennas/measurement configuration, accurate

data analysis depends upon the acquisition of high-quality data, the proper identification

of events, and depending upon imaging objectives, the precise measurement of event

amplitudes, periods, and/or travel times. Difficulties in establishing system parameters,

improper time zero determination, system drift, and deviations in positioning of the

antennas are problems that can occur during data acquisition. Other factors that can

affect measurements include the antenna radiation pattern (see above), refractions

generated by high velocity contrasts (Guy et al., 2001), wave guiding in a borehole or

layered sediments (Ellefsen, 1999), and resonance effects (Holliger and Bergmann,

2002). Additionally, previous work has demonstrated that borehole radar measurements

can be affected when energy radiated from the transmitter antenna induces currents on

conductive cables, and that periodic artifacts in data can result when currents traveling on

conductive cables reflect at system impedance mismatches (Guy and Radzevicius, 2001).

D.3 Cross-Hole Radar Data Acquisition: I-70 Study Area

Based on the I-70 study objectives (the detection of subsurface mine-related

subsidence activity, and the location of areas along the roadway having a relatively high

risk for future surface failure), system and antenna characteristics of the employed

borehole radar system, and radar signal-attenuating characteristics of the study area

geologic media, methods of cross-hole COP and MOG data acquisition were selected for

borehole radar surveying in the I-70 study area.


315
Cross-hole (COP and MOG) radar measurements were made along the eastbound

passing and travel lanes of Interstate 70 (I-70) in Guernsey County, Ohio, during July and

August of 2002. These measurements were made in boreholes that were located in the

vicinity of several mine-related subsidence features (bedrock horizon and overburden

discontinuities) that were previously imaged using seismic reflection (Chapter 6,

Appendix C). Four regions of the study area containing subsidence features were

surveyed using cross-hole radar: 1) eastbound travel lane road stations 46940-46968, 2)

eastbound passing lane road stations 48304-48379, 3) eastbound travel lane road stations

48304-48395, and 4) eastbound travel lane road stations 48530-48640 (Table D.1, Figure

D.1). (Note, road stations are in feet along the highway from the western Guernsey

county line. Highway engineers designate stationing as hundreds of feet plus the

remaining feet, for example, 469+40. In the text and the figures of this appendix these

numbers are combined, for example, 46940). The boreholes that were used for radar

measurements were drilled during either 1999 or 2002 (Table D.1).

Radar data were acquired using a commercial borehole radar system, with omni-

directional dipole antennas (measured center frequency of 100 MHz in air) located in

PVC-cased wells. The radar system employed relied on 30.0 m long conductive cables

(approximately 10 mm in diameter with a dielectric coated surface) for signal

transmission between surface-located electronics and the antennas.

Although single-hole measurements have the potential to provide useful

information regarding the subsurface, the radar antennas employed were omni-

directional, and it was felt that potential anomalies in recorded data could therefore be

spatially isolated in a more time-effective manner (with fewer surveys) by acquiring

316
MOG Transmit Receive Borehole Maximum
COP data Tx borehole location Rx borehole location
measurements (Tx) (Rx) separation antennas
filename (easting, ft), drill date (easting, ft), drill date
made? borehole borehole (m) depth (m)
726run1 Yes GC305 EBT 46962, 5/9/02 GC307 EBT 46947, 5/13/02 4.88 22
726run2 Yes GC307 EBT 46947, 5/13/02 GC306 EBT 46940, 5/10/02 1.96 23
726run3 Yes B407G EBT 46968, 4/18/02 B407H EBT 46954, 4/23/02 4.57 19
726run4 Yes GC212 EBT 48315, 10/19/99 GC211 EBT 48304, 10/14/99 3.5 19
726run5 Yes GC213 EBT 48326, 10/22/99 GC212 EBT 48315, 10/19/99 3.12 19
726run6 Yes GC214 EBT 48330, 10/27/99 CG213 EBT 48326, 10/22/99 1.4 19
726run7 Yes GC215 EBT 48340, 11/4/99 GC214 EBT 48330, 10/27/99 3.05 19
726run8 No GC216 EBT 48360, 10/18/99 GC215 EBT 48340, 11/4/99 5.97 20
726run9 No GC217 EBT 48380, 11/12/99 GC216 EBT 48360, 10/18/99 6.15 20
726run10 Yes B412E EBT 48395, 4/19/02 GC217 EBT 48380, 11/12/99 3.99 21
726run13 Yes B413H EBT 48537, 4/22/02 GC301 EBT 48530, 5/6/02 2.16 24
726run14 Yes GC303 EBT 48550, 5/7/02 B413H EBT 48537, 4/22/02 3.81 23
727run1 No B413F EBT 48605, 4/23/02 GC304 EBT 48583, 5/8/02 6.71 20

317
727run2 Yes GC302 EBT 48615, 5/7/02 B413F EBT 48605, 4/23/02 3.99 23
727run3 No B413E EBT 48640, 4/22/02 GC302 EBT 48615, 5/7/02 6.71 20
727run6 Yes GC202 EBP 48314, 10/19/99 GC201 EBP 48304, 10/8/99 3.25 16
727run7 Yes GC203 EBP 48323, 10/21/99 GC202 EBP 48314, 10/19/99 2.79 16
727run8 Yes GC204 EBP 48328, 10/26/99 GC203 EBP 48323, 10/21/99 1.6 16
727run9 Yes GC205 EBP 48340, 11/19/99 GC204 EBP 48328, 10/26/99 3.5 19
727run10 Yes GC206 EBP 48357, 11/10/99 GC205 EBP 48340, 11/19/99 5.13 22
727run11 No GC207 EBP 48379, 11/16/99 GC206 EBP 48357, 11/10/99 6.65 20

Table D.1. Cross-hole constant offset profile (COP) and multiple offset gather (MOG) radar measurements in the I-70 study
area (Figure D.1). Borehole locations correspond to road stations; EBT = eastbound travel lane, EBP = eastbound passing lane.
Road stations are given in feet from the western county line.
318
Figure D.1. Map of the eastbound lanes of I-70 (road stations 46900 to 48650) showing the locations of cross-hole constant
offset profile (COP) and multiple offset gather (MOG) radar measurements between borings (Table D.1), relative to the mapped
locations of underground mine workings. Due to map scale, mine room and pillar locations are regarded as approximate. Road
stations are given in feet from the western county line.
cross-hole measurements. Other factors supporting the decision to acquire cross-hole

rather than single-hole data were: 1) physical properties of geologic media between wells

could be determined (as a function of both lateral distance and depth) more accurately

and in a more time-effective manner using cross-hole measurements, and 2) the single-

hole receiver response is strongly influenced by conductive cable-related affects with the

employed radar system; such affects complicate single-hole data analysis.

D.3.1 Constant Offset Profile Measurements

Borehole radar COP surveys were conducted in the I-70 study area during July of

2002. COP surveys from which data are presented are described in Table D.1, and the

locations of the boreholes that were used are shown relative to the roadway eastbound

lanes and mapped mine workings in Figure D.1. Due to the small scale of the mine

workings map, room and pillar locations relative to the roadway and borehole locations

are regarded as approximate. Separations between boreholes were measured at the time

of data acquisition (measured separations assume no deviation of the wells from vertical

with depth; borehole deviation surveys were not conducted). The maximum

measurement depth for each record was dependent upon the depths of the surveyed

boreholes, and on the maximum depth that could be surveyed by both antennas (this was

limited by a combination of the length of the radar system cables and the allowable

position of the surface-located electronics relative to the boreholes).

For each COP record, both antennas were started at the maximum possible depth,

and were then both stepped upwards at the same increment (0.125 m), with the final

measurement made with both antenna midpoints positioned at 1.0 m below the ground

319
Figure D.2. Cross-hole radar measurements in the I-70 study area for an example case
(19.0 m maximum depth): (left) constant offset profile (COP), and (right) multiple offset
gathers (MOG’s). For MOG’s the maximum Tx and Rx vertical offset was 4.0 m.

320
surface (Figure D.2). In order to accurately position time zero and allow for the

correction of any possible system-related drift that may have occurred during

measurements, both the first and last 3 to 4 traces of each record were acquired in air with

a known (1.0 m) separation between the antennas. Detailed information regarding the

borehole radar COP data recording parameters is presented in Table D.2.

D.3.2 Multiple Offset Gather Measurements

Borehole radar MOG surveys were conducted in the I-70 study area during

August of 2002. MOG surveys from which data are presented are described in Table

D.1, and the locations of the boreholes used are shown in Figure D.1. Survey design for

MOG data acquisition was based on prior COP data analyses, through which practical

effective penetration distances for the I-70 subsurface media were determined. As

expected, useful source to receiver offsets (i.e. antenna separations yielding traces with

direct arrival signal distinguishable above the background noise level) that could be

recorded in the study area subsurface varied depending upon lithology. With the

employed radar system and antennas, the maximum effective source to receiver offset in

consolidated materials (e.g. coal, sandstone, shale) was about 8.0 m, the maximum

effective offset in unconsolidated fine to coarse sands was about 7.0 m, and the maximum

effective offset in unconsolidated clay-rich units was about 5.0 m. The maximum

borehole separation deemed to be suitable for tomographic imaging, considering study

area lithologies, was about 5.0 m. The maximum vertical source and receiver offset was

therefore limited to 4.0 m during MOG surveying, and measurements were not made

between boreholes with a separation significantly greater than 5.0 m.

321
Description Parameters
Data acquisition system Sensors and Software
Center frequency of antennas 100 MHz (in air)
Antenna step increment 0.125 m (8 traces per m)*, 0.25 m (4 traces per m)**
Time window 600 ns*, 350 ns**
Sample points 750*, 437**
Stacks 32
Transmitter (Tx) and receiver
Tx was always located in the eastern-most borehole
(Rx) spatial relationship
Maximum vertical offset
0 m*, 4.0 m**
between (Tx) and (Rx)
Pulsar voltage 1000 V

Table D.2. Acquisition and recording parameters for cross-hole constant offset profile
(COP)* and multiple offset gather (MOG)** radar measurements in the I-70 study area.

Multiple data files were acquired between each set of boreholes surveyed during

MOG data acquisition. For each MOG data file, the transmitter antenna was located and

kept stationary at a given position in one well, while the receiver antenna position was

stepped downwards at an increment of 0.25 m (Figure D.2). The initial location of the

transmitter and receiver antennas for each set of wells that were surveyed was 1.0 m

below the ground surface, and the final location of both antennas corresponded to the

maximum possible measurement depth for the set of boreholes (see the previous section

concerning COP data acquisition). Measurements with the antennas separated by 1.0 m

in air, were frequently and consistently made during MOG data acquisition, in order to

322
allow for time zero and possible system-related drift correction. Detailed information

regarding the borehole radar MOG data recording parameters is presented in Table D.2.

D.4 Cross-Hole Radar Data Processing and Imaging

Processing and imaging flows applied to cross-hole radar COP and MOG data

acquired in the I-70 study area are summarized in Table D.3 and Table D.4 respectively.

Demonstrations of applying the complete flows to COP and MOG field data are

presented in Figure D.3 and Figure D.4 respectively, with accompanying discussion

provided in the following sub-sections.

D.4.1 Constant Offset Profile Analysis

After initially applying a dewow correction to suppress the low frequency

component of data (Figure D.3a), true time zero (the time at which the transmitter

antenna initially radiates a pulse of energy) was established based on traces that were

acquired in air with a known dipole separation. Trace editing, data truncation, and

frequency filtering (to suppress high frequency noise) were then conducted (Table D.3).

The average absolute amplitude of each COP trace (unscaled) was calculated

across a 200 ns window. A broad window was selected for calculations rather than

tightly windowing direct arrivals for two reasons: 1) because radar signal was often

highly attenuated in certain geologic media for the surveyed offsets, and because

refractions interfered with direct arrivals in the near surface, it was difficult (impossible

in some cases) to always accurately window direct arrivals, and 2) the specification of a

broader window would also allow possible trace amplitude anomalies resulting from
323
Processing step Description
Dewow correction Suppression of low-frequency data component
Data reformat From pulseEKKO to SEG-Y to ProMAX format
Time zero correction True time zero established based on traces acquired in air
Trace editing Bad/noisy traces, and traces acquired in air killed
Data truncation Records truncated to 220 ns
Bandpass filter Ormsby filter: 0-30-90-120 MHz
Amplitude analysis Average absolute amplitude calculated across 200 ns window
Trace normalization Each trace scaled relative to maximum amplitude of the trace
Direct arrival breaks picked for calculation of average media
Direct arrival picking
velocities

Table D.3. Data processing and imaging flow for I-70 study area cross-hole constant
offset profile (COP) radar measurements (Figure D.3).

Processing step Description


Dewow correction Suppression of low-frequency data component
Data merge Merged individual MOG’s into single file
Data reformat From pulseEKKO to SEG-Y to ProMAX format
Time zero correction True time zero established based on traces acquired in air
Trace editing Bad/noisy traces killed
Data truncation Records truncated to 220 ns
Bandpass filter Ormsby filter: 0-30-90-120 MHz
Trace normalization Each trace scaled relative to maximum amplitude of the trace
Direct arrival picking Direct arrival breaks picked for travel time inversion
Travel time inversion Inversion conducted using MIGRATOM code
Tomogram generation Plotting of calculated velocity distribution between boreholes

Table D.4. Data processing and imaging flow for I-70 study area cross-hole multiple
offset gather (MOG) radar measurements (Figure D.4).
324
Figure D.3. Demonstration of constant offset profile data processing and imaging flow:
(a) field data after dewow correction, (b) after time-zero correction, trace editing,
truncation, bandpass filtering, amplitude analysis, and trace normalization, (c) after direct
arrival picks, and (d) processed data with amplitude and velocity plots.
325
Figure D.4. Demonstration of multiple offset gather data processing and imaging flow:
(a) field data after dewow correction, merging, time-zero correction, trace editing,
truncation, bandpass filtering, and trace normalization, (b) zoomed in look at MOG’s
after direct arrival picks, (c) plot of calculated velocity distribution between boreholes
obtained through inversion, and (d) velocity distribution plot after image interpolation.

326
post-direct arrival scattering to be more easily recognized. The approach taken allows an

assessment (although somewhat qualitative) of relative media attenuation characteristics

(related to conductivity and scattering loss) for a given set of measurements to be made,

and also provides a means for comparing recorded amplitudes of different traces acquired

in similar media types but at different borehole separations (effective penetration distance

and recorded amplitude are a function of radar system limitations and other factors in

addition to media attenuation characteristics). (Note: for the Sensors and Software

borehole radar system, a sample amplitude value can be multiplied by a factor of 1.526 to

obtain the unit of micro Volt). Traces were normalized after amplitude analysis for direct

arrival picking and data display purposes (Figure D.3b).

In order to allow the calculation of average radar propagation velocities for

geologic media between boreholes, direct arrivals (direct breaks) were manually picked

(Figure D.3c). Such velocity calculations were based on assumptions that the boreholes

did not deviate from vertical with depth, and that velocity was laterally constant

throughout media between boreholes. Picked direct arrival times were interpolated

across depths where high attenuation of radar signal or refraction interference prevented

accurate direct arrival picks from being made. Direct arrival signal was unable to be

enhanced through bandpass or f-k filtering in cases where near-surface refraction

interference occurred. Using quantities obtained from amplitude and direct break

analyses, plots of average absolute amplitude and average radar propagation velocity

(versus depth) were generated and displayed with processed COP records (Figure D.3d).

327
D.4.2 Multiple Offset Gather Analysis

The processing of MOG data was similar to the processing of COP data in that

dewow correction, time zero correction, trace editing, data truncation, bandpass

frequency filtering, and direct arrival picking were all conducted (Table D.4). One

difference was that individual MOG’s acquired between boreholes were merged together

during the MOG data processing flow (Figure D.4a). Another difference was that

inversion was conducted using MOG direct arrival travel time picks (Figure D.4b) for

tomographic imaging purposes. Detailed information regarding inversion theory as it

applies for the purpose of subsurface physical property tomography can be found in

Stewart (1991). The basic idea behind velocity inversion/tomography is that when

multiple travel time measurements along different ray paths (i.e. at different viewing

angles) through a media plane of interest are obtained, the velocity distribution within the

media plane can then be inferred from these measurements, as the spatial relationships of

the sources and receivers are known. Travel time inversion was conducted in order to

generate EM-wave velocity tomograms using a ray tracing computer code called

MIGRATOM (Jackson and Tweeton, 1993).

EM-wave propagation velocity can be approximated for practical purposes in

most low loss, non-magnetic geologic media as the speed of light (c) divided by the

square root of the media relative dielectric permittivity (k). For geologic (unconsolidated

and consolidated) materials, k values typically range from approximately 3 to 25,

depending upon mineralogy and water content. Air, quartz, and water have k values of 1,

4, and 80 respectively, and water content will therefore have a significant affect on media

EM-wave velocity. Water content is related to porosity in fully saturated media, and

328
changes in primary or secondary porosity of fully saturated media will result in EM-wave

velocity changes (porosity changes in dry media will result in velocity changes of smaller

magnitude). It is reasonable to expect that an increase in secondary porosity due to

fracturing and void formation may occur in subsurface media where subsurface

subsidence processes have been active. Therefore, it was felt that by

determining/imaging lateral and vertical EM-wave velocity distribution using

tomography, the potential to accurately locate possible areas of active mine-related

subsidence in the subsurface would exist (in addition to the potential for accurately

mapping lithologic variations).

The inversion process using MIGRATOM, first involved dividing the media

plane between boreholes into a specified grid of cells. Based on the calculated range

(from COP measurements) of dominant EM wavelengths in the I-70 subsurface materials

(0.9 m to 1.5 m; wavelength variation was mainly caused by media electrical property

variation), a cell size of 0.0625 m2 (0.25 m vertically and horizontally) was specified for

grids. Each cell within a plane was then assigned an initial velocity value (based on the

measured average velocity at the cell depth), and the synthetic travel times through each

cell along the portion of ray paths intersecting each cell were calculated. Through a

process of comparing the sum of synthetic travel times along straight ray paths with the

measured travel times, the velocities assigned to cells were iteratively changed/updated

by the program in order to reduce differences between the synthetic model and measured

travel times. Through trial inversions, it was determined that by limiting the maximum

number of solution iterations to 20, reasonable velocity distribution models would be

329
obtained (i.e. models that indicated media velocity variations between boreholes, but did

not suffer from the development of extremely large/unrealistic variations).

Velocity distributions calculated by MIGRATOM were plotted (Figure D.4c), and

a pixel interpolation function was then applied to gently smooth images (Figure D.4d).

Although MOG measurements were made from 1.0 m depth to the maximum possible

depth for all boreholes surveyed, signal attenuation and refraction interference often

prevented tomography attempts in the top several meters of the subsurface (Figure D.4d).

Concern exists when conducting inversion/tomographic imaging due to the issue

of mathematical/solution non-uniqueness. In the case of borehole radar MOG surveying,

the travel time measurements obtained through a media plane do not provide complete

angular coverage, and therefore do not provide enough information to allow a

mathematically unique reconstruction of the velocity distribution within the plane (i.e. an

obtained solution is regarded as non-unique, because more than one velocity distribution

model could be fit to a given set of travel time measurements). In order to improve the

velocity distribution reconstruction process during travel time inversions, only high

confidence travel time picks were considered (see above), and constraints (based on

calculated ranges of average velocities for the I-70 media types) limiting the range of

permissible solutions were specified.

Confidence in EM-wave velocity tomograms that were generated was established

in several ways. Trial inversions were conducted using different types of starting models

(i.e. mean and horizontal), and images obtained using both types of starting models were

found to be quite similar to one another. As Jackson and Tweeton (1993) discuss, a

unique solution is independent of the type of starting model, and if solutions obtained

330
using different starting models are extremely different from one another, then additional

constraints on inversion solutions are necessary to reduce the problem of non-uniqueness.

Velocity tomograms were also compared to average velocity versus depth plots that were

generated from COP records, and to drill log information that were available for the

surveyed boreholes. In all cases, tomogram velocity distributions were found to correlate

well with trends in average velocity versus depth plots, and with lithologic boundaries

indicated from drill log information (this is apparent from visual inspection of the data

that are subsequently presented and discussed).

D.5 Cross-Hole Radar Results and Interpretations

This section presents processed COP data and MOG data-derived EM-wave

velocity tomogram results, along with interpretations in terms of subsurface subsidence

activity and future roadway collapse risk, for locations of the I-70 study area that were

investigated using cross-hole radar methods. Data results and interpretations are

presented in one of the four following sub-sections, depending upon the location of the

boreholes used to acquire radar data: 1) eastbound travel lane road stations 46940-46968,

2) eastbound passing lane road stations 48304-48379, 3) eastbound travel lane road

stations 48304-48395, and 4) eastbound travel lane road stations 48530-48640 (Table D.1

and Figure D.1). Processed S-wave seismic reflection sections showing subsidence

features (bedrock and overburden discontinuities) that were imaged near the locations of

boreholes used to acquire radar data, and well log information obtained during drilling of

the boreholes used for radar measurements are also presented.

331
D.5.1 Eastbound Travel Lane Road Stations 46940 to 46968

Shown in Figure D.5 are average EM-wave amplitude and velocity plots,

processed radar COP data, drill logs, and MOG-data derived EM-wave velocity

tomograms for three sets of boreholes in the I-70 eastbound travel lane road station range

46940 to 46968 (Table D.1, Figure D.1). A mosaic of the velocity tomograms shown in

Figure D.5 is presented in Figure D.6, along with a geologic cross-section constructed

from 2002 drill logs, an interpreted S-wave seismic reflection depth section, and a map

showing the approximate locations of mine workings relative to the locations of the data

in the figure.

From Figures D.5a - D.5c, it is seen that in general there is very good correlation

between calculated average amplitude and velocity value variations versus depth, and

lithologic changes that are indicated by drill logs. In Figure D.5a for example, mineral

content and porosity differences (i.e. associated differences in electrical properties) of

mapped unconsolidated sand and shale units result in expected EM-wave amplitude and

velocity variations with depth. From 13.5 to 18.0 m depth, shale (having no heavy

fracturing detected during drilling) is seen to correlate with relatively high COP-derived

values of average amplitude (~3800 µV) and velocity (~0.063-0.068 m/ns). There is a

large decrease in both amplitude (to ~500 µV) and velocity (to <0.05 m/ns) over the

depth range of 20.0 to 21.5 m, which correlates with heavily fractured shale mapped

during drilling. Although mineralogic or primary porosity differences between the shale,

or groundwater specific conductance differences at different depths may have contributed

to the observed differences in measured EM-wave amplitudes and velocities, data suggest

that an increase in fracture density (i.e. secondary porosity) of the shale located at greater
332
(a) Borehole radar data and drill-log plots for wells GC307 and GC306.

Figure D.5. Average EM-wave velocity and average absolute amplitude plots, radar COP
data, drill logs, and EM-wave velocity tomograms for I-70 eastbound travel lane road
station range 46940 to 46968 (Table D.1, Figure D.1): (a) wells GC307 and GC306, (b)
wells GC305 and GC307, and (c) wells B407G and B407H. A mosaic of the velocity
tomograms and well log information are shown along with interpreted seismic reflection
data and a mine map in Figure D.6. See text for a discussion of plots. (continued)

333
Figure D.5. (continued)

(b) Borehole radar data and drill-log plots for wells GC305 and GC307.

(c) Borehole radar data and drill-log plots for wells B407G and B407H.
334
335
Figure D.6. I-70 eastbound travel lane road station range 46940 to 46968: (a) EM-wave velocity mosaic (Figure D.5), (b) S-wave reflection
data, (c) geologic cross-section from logs, and (d) approximate locations of mine workings. Velocities (a) and drill logs (c) indicate that the
SE edge of the coal pillar beneath the road actually extends farther SE than mapped in (d), and that the seismically imaged subsidence
feature (b) resulted from bedrock collapse into the mine room located immediately south of this pillar. See text for complete discussion.
depth has had an effect on media electrical properties that was detectable using radar. In

this situation, it is apparent that EM-wave velocity has decreased and radar signal

attenuation has increased due to the fracturing (relative dielectric permittivity and

conductivity have increased due to water content increase; radar scattering losses from

electrical property discontinuities have also likely increased).

From Figure D.5a, it is seen that there is good lateral and vertical correlation of

the EM-wave velocity distribution in the tomogram with drill log information, and that

vertical variations in average EM-wave velocity in the tomogram agree with those in the

COP data-derived average velocity versus depth plot. From 18.0 m to 20.0 m depth, the

drill logs indicate that coal is present at the eastern edge of the tomogram, and that grout

is present at the western edge. Relatively high EM-wave velocities are observed near the

eastern edge of the tomogram in this depth range, while relatively low EM-wave

velocities are observed at the western edge of the tomogram in this depth range. In

Figure D.5b, drill logs indicate that a continuous bituminous coal seam exists between the

boreholes from 18.0 to 20.0 m depth. The velocity tomogram indicates that the EM-wave

velocity of the continuous coal (~0.09 - 0.1 m/ns) is higher than that of the overlying

shale unit (~0.08 m/ns). Average amplitude values are also seen to be relatively high at

the depth of the continuous coal. Based on previous laboratory measurements by Cook

(1975), the field observance of a high velocity and low attenuation (relative to shale)

bituminous coal is reasonable.

Radar data presented in this section were acquired using boreholes that were

located to the immediate north of a seismic reflection line that was positioned just south

the eastbound travel lane. The seismic section in the vicinity of the EM-wave velocity

336
tomogram mosaic (Figure D.6a) indicates that the bedrock horizon has subsided into the

mine workings along normal faults, within close proximity to the line between road

stations 46938 and 46957 (Figure D.6b). It is also apparent that a section of bedrock has

subsided to a certain degree, into a mine room to the immediate east of these faults.

Interpolated drill log information to the north of the seismic line (Figure D.6c)

indicate that the local top-of-bedrock is about 13.5 m deep (between the boreholes) from

road stations 46940 to 46968, while the seismic data indicate that bedrock is around 15.0

m deep across this range to the immediate south. Both drill logs and EM-wave velocity

distribution in the tomogram indicate that a continuous coal seam was encountered

beneath bedrock during drilling to the north of the seismic line, between road stations

46947 and 46962. There is a discrepancy between the drill logs and the mapped locations

of the mine room and pillar workings shown in Figure D.6d, as the mine map indicates

that the boreholes should have encountered a mine room and not a coal pillar. It appears

that the southeastern edge of the coal pillar (seen beneath the eastbound lane near

boreholes GC307 and GC305) actually extends slightly farther to the southeast than the

mine map indicates, and terminates just south of the eastbound travel lane.

Radar COP data (Figure D.5a) and velocity tomograms (Figure D.6a) indicate a

termination of the coal pillar west edge at approximately road station 46947, and a

decrease in EM-wave velocities (relative to velocities to the immediate east) at the

bedrock level between road stations 46940 and 46947. A heavily fractured shale unit that

was mapped west of station 46947 (see above) suggests that mine-related subsidence

processes have been active at the mine level in this area. It is possible that the observed

decrease in bedrock EM-wave velocity (above the mine level) to the west of station

337
46947 has resulted from a certain degree of subsidence-related fracturing (which was not

indicated by drill logs). Such fracturing may be related to the formation of the bedrock

subsidence feature that was seismically imaged to the south of the boreholes. Possible

overburden disruption in the road station range surveyed by borehole radar is not

apparent over the depth range of COP data and EM-wave velocity tomograms.

Considering borehole-derived drill log and EM-wave velocity information, it is

apparent that the subsidence feature imaged using seismic reflection (Figure D.6b),

resulted from a collapse of the bedrock horizon into the mine room located immediately

south of the coal pillar edge encountered by the boreholes. As was discussed previously,

an area of disruption responsible for a 2D seismic section anomaly does not necessarily

need to be located directly beneath the seismic line due to the Fresnel zone concept.

Based on the close proximity of the seismically imaged bedrock subsidence-related

disruption to the southern edge of the eastbound travel lane, the eastbound travel lane

road station range of 46938 to 46957 is regarded as having a relatively high potential for

future surface failure.

D.5.2 Eastbound Passing Lane Road Stations 48304 to 48379

Shown in Figure D.7 are average EM-wave amplitude and velocity plots,

processed radar COP data, drill logs, and MOG-data derived EM-wave velocity

tomograms for six sets of boreholes in the I-70 eastbound passing lane road station range

48304 to 48379 (Table D.1, Figure D.1). A mosaic of the velocity tomograms shown in

Figure D.7 is presented in Figure D.8, along with a geologic cross-section constructed

from 1999 drill logs, an interpreted S-wave seismic reflection depth section, and a map

338
(a) Borehole radar data and drill-log plots for wells GC202 and GC201.

Figure D.7. Average EM-wave velocity and average absolute amplitude plots, radar COP
data, drill logs, and EM-wave velocity tomograms for I-70 eastbound passing lane road
station range 48304 to 48379 (Table D.1, Figure D.1): (a) wells GC202 and GC201, (b)
wells GC203 and GC202, (c) wells GC204 and GC203, (d) wells GC205 and GC204, (e)
wells GC206 and GC205, and (f) wells GC207 and GC206 (no MOG data acquired). A
mosaic of the velocity tomograms and well log information are shown along with
interpreted seismic reflection data and a mine map in Figure D.8. See text for a
discussion of plots. (continued)

339
Figure D.7. (continued)

(b) Borehole radar data and drill-log plots for wells GC203 and GC202.

(c) Borehole radar data and drill-log plots for wells GC204 and GC203. (continued)

340
Figure D.7. (continued)

(d) Borehole radar data and drill-log plots for wells GC205 and GC204.

(e) Borehole radar data and drill-log plots for wells GC206 and GC205. (continued)

341
Figure D.7. (continued)

(f) Borehole radar data and drill-log plots for wells GC207 and GC206.

342
343
Figure D.8. I-70 eastbound passing lane road station range 48300 to 48360: (a) EM-wave velocity mosaic (Figure D.7), (b) S-wave
reflection data, (c) geologic cross-section, and (d) approximate locations of mine workings. Velocities (a) indicate the mine-related bedrock
subsidence interpreted from seismic data (b) has occurred between the boreholes and directly beneath the seismic line. Velocities (a) also
extend the west edge of disruption interpreted from seismic data an additional several meters west. See text for complete discussion.
showing the approximate locations of mine workings relative to the locations of the data

in the figure.

Radar data presented in this section were acquired using boreholes located to the

immediate south of a seismic reflection line that was positioned to the north of the

eastbound passing lane. The bedrock horizon was interpreted from seismic reflection

data to have been down-dropped along normal faults, and to have experienced significant

disruption from mine-related subsidence, near the seismic line between road stations

48329 and 48354 (Figure D.8b). A geologic cross-section interpreted from 1999 drill log

data indicated that a bedrock low exists in the approximate road station range of 48314 to

48340, and that voids below the bedrock surface also exist in this region (Figure D.8c).

Although the lateral extent (25-26 ft) of both interpreted features was similar, the western

edge of disrupted bedrock interpreted from seismic data was located to the east of that

interpreted from drill logs.

Consideration of the EM-wave velocity tomogram mosaic across this region

(Figure D.8a) suggests that seismic data were unable to exactly delineate the western

edge of bedrock horizon disturbance between the boreholes. Between wells GC202 and

GC203, a region of low EM-wave velocity (relative to laterally adjacent velocities) exists

(from 14.0 to 16.0 m deep) at the local surveyed bedrock level. The low velocity zone

between these wells exists between road stations 48314 and 48319. This bedrock low

velocity area is interpreted to be the result of increased water content (the high relative

dielectric permittivity of water reduces EM-wave velocity) due to subsidence related

fracturing and voids (a void was encountered during 1999 drilling at 15.5 m depth at

station 48314). Wells in this region were not drilled to the depths necessary to confirm

344
coal presence or absence. However, data observations suggest that the mine map (which

indicates coal beneath wells GC202 and GC203) in Figure D.8d has placed the eastern

edge of the coal pillar in this region too far to the east.

The EM-wave velocity tomogram mosaic (Figure D.8a) indicates a disrupted and

down-dropped bedrock horizon between wells GC203 and GC205 (road stations 48323 to

48340). Relatively low EM-wave velocities at the surrounding bedrock level in this

region suggest an increased secondary porosity (increased water content) due to

fracturing and bedrock subsidence. The inter-layering of grout and shale beneath the

bedrock surface (as mapped by well GC205) has contributed to relatively low velocities

within the region at the bedrock and mine levels, and along with fracturing has resulted in

COP data wavelet cycle distortion (due to direct arrival and diffraction interference) in

the depth range of 16.0 m to 18.0 m (Figure D.7d).

Between wells GC205 and GC206 (road stations 48340 to 48357) EM-wave

velocities at the bedrock level are somewhat higher than those to the immediate west

(Figure D.8a). Although substantial mineralogic or primary porosity differences are

possible, this observation suggests a relative decrease in bedrock subsidence-related

fracture density, which is consistent with the seismic data that indicate an intact (but

down-dropped) section of bedrock between stations road 48348 and 48354. The location

of wells prevented the eastern edge of bedrock horizon disruption (as indicated by

seismic data) from being accurately interpreted from drill log information alone (Figure

D.8c). A well log located 1.8 m east of well GC206 (well GC206 is located at road

station 48357) indicates that a void was encountered during drilling between 15.8 m to

17.5 m deep. Relatively high average EM-wave velocities at the bedrock level between

345
road stations 48357 to 48379 (Figure D.7f), suggest however that the surface of bedrock

between these wells is intact and has not been disrupted from subsidence. This

observation agrees with seismic data, which show an intact bedrock surface between

these stations.

Although it is difficult to infer possible subsidence-related disruption of near-

surface overburden materials from the processed radar and seismic data in the eastbound

passing lane road station range of 48314 to 48354, these data clearly indicate bedrock

horizon disruption (between the wells and beneath the seismic line) in this region.

Because of the close proximity of disrupted bedrock to the roadway, the eastbound

passing lane is regarded as having a relatively high potential for future mine-related

surface failure in this region.

D.5.3 Eastbound Travel Lane Road Stations 48304 to 48395

Shown in Figure D.9 are average EM-wave amplitude and velocity plots,

processed radar COP data, drill logs, and MOG-data derived EM-wave velocity

tomograms for seven sets of boreholes in the I-70 eastbound travel lane road station range

48304 to 48395 (Table D.1, Figure D.1). A mosaic of the velocity tomograms shown in

Figure D.9 is presented in Figure D.10, along with a geologic cross-section constructed

from 1999 and 2002 drill logs, an interpreted S-wave seismic reflection depth section,

and a map showing the approximate locations of mine workings relative to the locations

of the data in the figure.

Radar data presented in this section were acquired using boreholes located to the

immediate north of a seismic reflection line that was positioned to the south of the

346
(a) Borehole radar data and drill-log plots for wells GC212 and GC211.

Figure D.9. Average EM-wave velocity and average absolute amplitude plots, radar COP
data, drill logs, and EM-wave velocity tomograms for I-70 eastbound travel lane road
station range 48304 to 48395 (Table D.1, Figure D.1): (a) wells GC212 and GC211, (b)
wells GC213 and GC212, (c) wells GC214 and GC213, (d) wells GC215 and GC214, (e)
wells GC216 and GC215 (no MOG data acquired), (f) wells GC217 and GC216 (no
MOG data acquired), and (g) wells B412E and GC217. A mosaic of the velocity
tomograms and well log information are shown along with interpreted seismic reflection
data and a mine map in Figure D.10. See text for a discussion of plots. (continued)

347
Figure D.9. (continued)

(b) Borehole radar data and drill-log plots for wells GC213 and GC212.

(c) Borehole radar data and drill-log plots for wells GC214 and GC213. (continued)

348
Figure D.9. (continued)

(d) Borehole radar data and drill-log plots for wells GC215 and GC214.

(e) Borehole radar data and drill-log plots for wells GC216 and GC215. (continued)

349
Figure D.9. (continued)

(f) Borehole radar data and drill-log plots for wells GC217 and GC216.

(g) Borehole radar data and drill-log plots for wells B412E and GC217.

350
351
Figure D.10. I-70 eastbound travel lane road stations 48304 to 48400: (a) EM-wave velocity mosaic (Figure D.9), (b) S-wave reflection
data, (c) geologic cross-section from logs, and (d) approximate locations of mine workings. Velocities (a) support seismic data
interpretation (b) of an intact bedrock surface (stations 48304 to 48340), and bedrock and overburden subsidence-related disruption (stations
48380 to 48408), indicating this disruption occurred between the boreholes and directly beneath the seismic line. See text for discussion.
eastbound travel lane. The bedrock horizon surface was interpreted from seismic

reflection data to be intact between road stations 48304 and 48340 (Figure D.10b), and

this agrees with a geologic cross-section constructed from drill log data (Figure D.10c).

An EM-wave velocity tomogram mosaic across this road station range (Figure D.10a)

also suggests that the bedrock surface is intact in this region, with relatively high EM-

wave velocities observed for most of the bedrock depth range. Voids encountered

beneath the bedrock surface (17.0 to 19.0 m depth) during the drilling of wells GC212,

GC213, and GC214 are apparent from the velocity tomograms, and are observed as

velocity lows (attributed to increased water content). Average amplitude lows

(suggesting increased attenuation and scattering loss) are also seen in the void locations

from the COP-derived plots (Figures D.9a-D.9d). As was concluded from the seismic

data interpretation, it does not appear from the radar COP-data and EM-wave velocity

tomograms that these voids have yet propagated up through the bedrock horizon, and

therefore an immediate risk for surface failure does not exist for this eastbound travel

lane road station range.

A bedrock discontinuity that resulted in vertical horizon offset was interpreted

from seismic data at road station 48391 (Figure D.10b). Analysis of seismic CDP gathers

indicated that the mine-related bedrock surface disruption exists near the seismic line

from the road station range of 48380 to 48408. Near-surface events in seismic data also

exhibit apparent dip and offset across this road station range, suggesting that overburden

materials in this area have also been disturbed by mine-related subsidence activity. A

well (number B412E) drilled during 2002 (subsequent to seismic data interpretation) at

road station 48395 confirmed that bedrock had in fact subsided to the east of road station

352
48380, along a normal fault located between road stations 48380 and 48395 (Figure

D.10c). A heavily fractured shale and siltstone unit was also encountered beneath the

bedrock surface during the 2002 drilling of well B412E, and this unit extended to mine

level depths with no coal encountered. This suggests that where the mine map (Figure

D.10d) indicates a coal pillar beneath well B412E, there should actually be a mine room

indicated (i.e. the mine map has placed the eastern edge of the coal pillar too far to the

east). Voids were previously encountered (from 13.3 to 14.7 m deep) during the 1999

drilling of well GC217 (located at road station 48380).

The EM-wave velocity tomogram (Figure D.10a) between wells GC217 and

B412E (road stations 48380 to 48395) supports the seismic and drill log data

interpretations of a disturbed bedrock horizon. Bedrock EM-wave velocities are seen to

be low relative to those measured in the eastbound travel lane road station range of 48304

to 48340. Overburden velocities are also seen to be low in this region relative to those of

unconsolidated units located to the west, suggesting an increase in fracture density (and

water content). A comparison of average absolute amplitude plots (Figures D.9e and

D.9g) supports this EM-wave velocity-based interpretation. Higher values of average

radar signal amplitude would be expected for measurements made between wells GC217

and B412E (Figure D.9g), which were 3.99 m apart, than for measurements made

between wells GC215 and GC216 (Figure D.9e), which were 5.97 m apart. The

observation that average amplitude values are similar at the depths of overburden for both

sets of measurements, suggests that an increase in radar signal attenuation due to an

increase in overburden fracture density (and water content) between wells GC217 and

B412E has occurred.

353
Due to borehole separation distance, MOG surveying (and thus tomographic

imaging) was not conducted between wells GC215 and GC216, or between wells GC216

and GC217 (road stations 48340 to 48380). Acquired radar COP data and calculated

average velocity and amplitude values between these boreholes are shown in Figures

D.9e and D.9f. High radar signal attenuation within unconsolidated materials prevents an

interpretation concerning overburden integrity based upon these data from being made.

Decreasing average velocity and amplitude values at the depth of the void mapped by

borehole GC217 (Figure D.9f) suggests that the void beneath the bedrock surface may

extend to the west of this borehole. Average velocities at the bedrock level do not

suggest however, that the region of bedrock surface disruption between road stations

48380 and 48408 (see above) extends west of road station 48380, and this interpretation

is consistent with a previous interpretation based only on seismic and drill log data.

Mine-related subsidence disruption of the bedrock horizon and near-surface

overburden materials is interpreted in the eastbound travel lane road station range of

48380 to 48408 (between the boreholes and beneath the seismic line), based upon

processed radar and seismic data. Based upon the close proximity of the disrupted

bedrock and overburden to the roadway, the eastbound travel lane is regarded as having a

relatively high potential for future mine-related surface failure in this region.

D.5.4 Eastbound Travel Lane Road Stations 48530 to 48640

Shown in Figure D.11 are average EM-wave amplitude and velocity plots,

processed radar COP data, drill logs, and MOG-data derived EM-wave velocity

tomograms for five sets of boreholes in the I-70 eastbound travel lane road station range

354
(a) Borehole radar data and drill-log plots for wells B413H and GC301.

Figure D.11. Average EM-wave velocity and average absolute amplitude plots, radar
COP data, drill logs, and EM-wave velocity tomograms for I-70 eastbound travel lane
road station range 48530 to 48640 (Table D.1, Figure D.1): (a) wells B413H and GC301,
(b) wells GC303 and B413H, (c) wells B413F and GC304 (no MOG data acquired), (d)
wells GC302 and B413F, and (e) wells B413E and GC302 (no MOG data acquired). A
mosaic of the velocity tomograms and well log information are shown along with
interpreted seismic reflection data and a mine map in Figure D.12. See text for a
discussion of plots. (continued)

355
FigureD.11. (continued)

(b) Borehole radar data and drill-log plots for wells GC303 and B413H.

(c) Borehole radar data and drill-log plots for wells B413F and GC304. (continued)
356
FigureD.11. (continued)

(d) Borehole radar data and drill-log plots for wells GC302 and B413F.

(e) Borehole radar data and drill-log plots for wells B413E and GC302.

357
358
Figure D.12. I-70 eastbound travel lane road station range 48530 to 48620: (a) EM-wave velocity mosaic (Figure D.11), (b) S-wave
reflection data, (c) geologic cross-sections from logs, and (d) approximate locations of mine workings. Velocities (a) suggest that the coal
pillar beneath the road (stations 48605 to 48615) actually extends farther south than mapped in (d), and that the seismically imaged
subsidence feature (b) resulted from bedrock collapse into the mine room located just south of this pillar. See text for complete discussion.
48530 to 48640 (Table D.1, Figure D.1). A mosaic of the velocity tomograms shown in

Figure D.11 is presented in Figure D.12, along with geologic cross-sections constructed

from 2002 drill logs, an interpreted S-wave seismic reflection depth section, and a map

showing the approximate locations of mine workings relative to the locations of the data

in the figure.

Radar data presented in this section were acquired using boreholes that were

located to the immediate north of a seismic reflection line that was positioned just south

of the eastbound travel lane. The seismic section in the vicinity of the EM-wave velocity

tomogram mosaic (Figure D.12a) indicates that the bedrock horizon surface is intact

between road stations 48530 and 48550 (Figure D.12b). Both the EM-wave velocity

tomograms and drill logs (Figure D.12c) to the north of the seismic line support the

seismic data interpretation, indicating relatively high EM-wave velocities for most of the

bedrock volume in this area. Grout and void space in the depth range of 19.0 m to 22.0 m

(beneath the bedrock surface) correlate with relative lows in both average EM-wave

velocity and amplitude (Figures D.11a and D.11b). These trends are due to a

combination of grout presence and an increase in water content due to fracturing and void

space. The distribution of velocity lows associated with the grout and increased

secondary porosity is mapped well between the wells by the velocity tomograms. A zone

of low velocity is apparent from a tomogram within the bedrock volume over the depth

range of 18.0 to 20.0 m, between stations 48534 and 48537 (Figure D.11a). This

suggests that a zone of fracturing associated with subsidence activity has propagated up

into the bedrock, although it does not appear that the integrity of the entire horizon has

been affected yet, and a relative high risk for future surface failure does not yet exist here.

359
The seismic data indicate that the bedrock horizon has subsided into a mine room

along normal faults between road stations 48596 and 48615 near the seismic line (Figure

D.12b). Drill log (wells B413F and GC302) information (Figure D.12c) and EM-wave

velocity distribution in the tomogram (Figure D.12a) indicate that a continuous coal seam

was encountered beneath bedrock during drilling to the north of the seismic line, between

road stations 48605 and 48615. There is a discrepancy between the drill logs and the

mapped locations of the mine room and pillar workings shown in Figure D.12d, as the

mine map indicates that these boreholes should have encountered a mine room and not a

coal pillar. It appears that the southwestern edge of the coal pillar (seen beneath the

eastbound lane near these wells) actually extends slightly farther to the south than the

mine map indicates, and terminates just south of the eastbound travel lane. Radar COP

data (Figures D.11c and D.11e) indicate relatively high values of average EM-wave

velocity and amplitude over the bedrock depth range between road stations 48583 and

48605, and between road stations 48615 and 48640. These observations suggest that the

bedrock horizon is intact between the boreholes within these regions. Due to borehole

separations, MOG surveying (and thus tomographic imaging) was not conducted between

wells GC304 and B413F, or between wells GC302 and B413E (Figures D.11c and

D.11e).

Considering borehole-derived log and EM-wave velocity information, it is

apparent that the subsidence feature imaged using seismic reflection (Figure D.12b),

resulted from a collapse of the bedrock horizon into the mine room located immediately

south of the coal pillar edge (see above) encountered by boreholes B413F and GC302

(Figure D.12c). As was previously discussed, an area of disruption responsible for a 2D

360
seismic section anomaly does not necessarily need to be located directly beneath the

seismic line due to the Fresnel zone concept. Heavily fractured shale units were

encountered during the drilling of these boreholes, and such fracturing above the mine

level may be related to the formation of the subsidence feature that was seismically

imaged to the south of the boreholes. Possible disruption of overburden materials in the

road station range surveyed by boreholes is not apparent from the radar COP data and

EM-wave velocity tomograms.

Based on the close proximity of the seismically imaged bedrock horizon

disruption to the southern edge of the eastbound travel lane, the eastbound travel lane

road station range of 48596 to 48615 is regarded as having a relatively high potential for

future surface failure.

361
BIBLIOGRAPHY

Aki, K., and Richards, P.G., 1980, Quantitative seismology: theory and methods, W.H.
Freeman and Company, San Francisco, 557 p.

Alford, R.M., Kelly, K.R., and Boore, D.M., 1974, Accuracy of finite-difference
modeling of the acoustic wave equation, Geophysics, vol. 39, pp. 834-842.

Allen, L.J., and Peddy, C.P., 1993, Amplitude variation with offset: Gulf Coast case
studies (ed. F.K. Levin), Geophysical Development Series, vol. 4, Society of Exploration
Geophysicists, Tulsa Oklahoma, 124 p.

Badley, M.E., 1985, Practical seismic interpretation, International Human Resources


Development Corporation, Boston, Massachusetts, 266 p.

Bacharach, R. and Nur, A., 1998, High-resolution shallow-seismic experiments in sand,


part 1: water table, fluid flow, and saturation, Geophysics, vol. 63, pp. 1225-1233.

Baker, G. S., 1998, Applying AVO analysis to GPR data, Geophysical Research Letters,
vol. 25, no. 3, pp. 397-400.

BBC&M Incorporated, 1998, Abandoned underground mines map, Columbus, Ohio.

BBC&M Incorporated, 1999, Drill log data acquired in the previous roadway collapse
region of the I-70 study area, Columbus, Ohio.

Bell, F.G., 1999, Geologic hazards: Their assessment, avoidance, and mitigation, E & FN
Spon, New York, 648 p.

Benjumea, B., Hunter, J.A., Pullan, S.E., Burns, R.A., Good, R.L., 2001, Near-surface
seismic studies to estimate potential earthquake ground motion amplification at a thick
soil site in the Ottawa river valley, Canada, in: Proceedings of the Symposium on the
Application of Geophysics to Engineering and Environmental Problems, 11 p.

Bortfeld, R., 1961, Approximation to the reflection and transmission coefficients of plane
longitudinal and transverse waves: Geophysical Prospecting, vol. 9, pp. 485-502.

362
Bradford, J.H., 2002, Depth characterization of shallow aquifers with seismic reflection,
part 1 - the failure of NMO velocity analysis and quantitative error prediction,
Geophysics, vol. 67, no. 1, pp. 89-97.

Brady, B.H.G., and Brown, E.T., 1993, Rock mechanics for underground mining,
Chapman and Hall, London, 519 p.

Brown, J.R., 2000, Towards a polarity standard for multicomponent seafloor seismic
data, 2000, in: Annual Meeting Society of Exploration Geophysicists Expanded
Abstracts, 4 p.

Burger, H.R., 1992, Exploration geophysics of the shallow subsurface, Prentice-Hall,


Inc., Englewood Cliffs, NJ, 489 p.

Carr, B.J., Hajnal, Z., and Prugger, A., 1998, Shear-wave studies in glacial till,
Geophysics, vol. 63, pp. 1273-1284.

Castagna, J. P., 1993, AVO analysis - tutorial and review, in: Offset-Dependent
Reflectivity - Theory and Practice of AVO Analysis, Investigations in Geophysics Series,
Volume 8, Society of Exploration Geophysicists, Tulsa, Oklahoma, pp. 3-36.

Chung, W.Y., and Corrigan, D., 1985, Gathering mode-converted shear-waves: a model
study, in: 55th Annual International Meeting of the Society of Exploration Geophysicists
Expanded Abstracts, pp. 602-604.

Clark, J.C., Johnson, W.J., and Miller, W.A., 1994, The application of high resolution
shear wave seismic reflection surveying to hydrogeological and geotechnical
investigations, in: Symposium on the Application of Geophysics to Engineering and
Environmental Problems, pp. 231-245.

Condit, D.D., 1912, Conemaugh formation in Ohio, Ohio Geological Survey Bulletin 17,
Ohio Department of Natural Resources, Division of Geological Survey, Columbus, Ohio
363 p.

Cook, J.C., 1975, Radar transparencies of mine and tunnel rocks, Geophysics, vol. 40, pp.
865-885.

Costain, J.K., Cook, K.L., and Algermissen, S.T., 1963, Amplitude, energy, and phase
angles of plane SV waves and their application to earth crustal studies, Bulletin of the
Seismological Society of America, vol. 53, pp. 1039-1074.

Crampin, S., 1987, Crack porosity and alignment from shear-wave VSPs, in: Domenico,
S.N., and Danbom, S.H., Ed.’s, Shear Wave Exploration, Society of Exploration
Geophysicists, Tulsa, Oklahoma, pp. 227-251.

363
Crouch, T.M., Collins, H.R., and Helgesen, J.O., 1980, Abandoned subsurface coal mines
as a source of water for coal conversion in eastern Ohio, Ohio Department of Natural
Resources, Division of Geological Survey, Report of Investigations 118, 25 p.

Crowell, D.L., 1997a, Mine subsidence, GeoFacts No. 12, Ohio Department of Natural
Resources, Division of Geological Survey, Columbus, Ohio, 6 p.

Crowell, D.L., 1997b, Coal mining and reclamation, GeoFacts No. 15, Ohio Department
of Natural Resources, Division of Geological Survey, Columbus, Ohio, 4 p.

Davis, R.O., and Selvadurai, A.P.S., 1996, Elasticity and geomechanics, Cambridge
University Press, 201 p.

Denham, L.R., and Palmeira, R.A.R., 1984, On: reflection and transmission of plane
compressional waves by R.D. Tooley, T.W. Spencer, and H.F. Sagori, (Geophysics, vol.
30, pp. 552-570, April, 1965), Geophysics, vol. 49, p. 2195.

Dey-Sakar, S.K., and Svatek, S.V., 1993, Prestack analysis – an integrated approach for
seismic interpretation in clastic basins, in: Offset-Dependent Reflectivity - Theory and
Practice of AVO Analysis, Investigations in Geophysics Series, Volume 8, Society of
Exploration Geophysicists, Tulsa, Oklahoma, pp. 57-77.

Diedda, G.P., and Ranieri, G., 2001, Some SH-wave seismic reflections from depths of
less than three meters, Geophysical Prospecting, vol. 49, pp. 499-508.

Dix, C.H., 1955, Seismic velocities from surface measurements, Geophysics, vol. 20, pp.
68-86.

Dobecki, T.L., 1993, High resolution in saturated sediments – a case for shear wave
reflection, in: Proceedings of the Symposium on the Application of Geophysics to
Engineering and Environmental Problems, pp. 319-333.

Domenico, S.N., 1976, Effect of brine-gas mixture on velocity in an unconsolidated sand


reservoir, Geophysics, vol. 41, pp. 882-894.

Dubois, J.C., 1995, Borehole radar experiment in limestone: analysis and data processing,
First Break, vol. 13, pp. 57-67.

Eliasata, and Michelena, R.J., 1995, Mapping distribution of fractures in a reservoir with
P-S converted waves, The Leading Edge, pp. 664-676.

Ellefsen, K.J., 1999, Effects of layered sediments on the guided wave in crosswell data,
Geophysics, vol. 64, pp. 1698-1707.

364
Ensley, R.A., 1984, Comparison of P- and S-wave seismic data: a new method for
detecting gas reservoirs, Geophysics, vol. 49, pp. 1420-1431.

Ensley, R.A., 1985, Evaluation of direct hydrocarbon indicators through a comparison of


compressional- and shear-wave seismic data: a case study of the Myrnam gas field,
Alberta, Geophysics, vol. 50, pp. 37-48.

Fisher, W., 1971, Detection of abandoned underground coal mines by geophysical


methods: U.S. Environmental Protection Agency, Water Pollution Control Research
series, Project 14010EHN, 94 p.

Frasier, C., and Winterstein, D., 1990, Analysis of conventional and converted mode
reflections at Putah sink, California using three-component data, Geophysics, vol. 55, no.
6, pp. 646-659.

Freedman, D., 1999, Vector wavefield seismic evolves, AAPG Explorer, October edition,
American Association of Petroleum Geologists, pp. 12-14.

Fyfe, D.J., Dent, B.E., Kelamis, P.B., Al-Mashouq, K.H., and Nietupski, D.A., 1993,
Three-component seismic experiments in Saudi Arabia, Presented at the 55th Annual
Meeting of the European Association of Exploration Geophysicists.

Galloway, D., Jones, D.R., and Ingebritsen, S.E., 1999, Land subsidence in the United
States, United States Geological Survey, Circular 1182, Reston, Virginia, 177 p.

Gannett Fleming Corddry & Carpenter, 1995, Letter Report, Mine subsidence
investigation, Interstate 70, GUE-070-14.10, Gurnsey County, Ohio: Gannett Fleming
Corddry & Carpenter, Columbus, Ohio.

Garotta, R., 2000, Shear waves from acquisition to interpretation: short course notes,
distinguished instructor series no. 3, Society of Exploration Geophysicists, Tulsa,
Oklahoma.

Gilson, E.W., Redman, J.D., Pilon, J., and Annan, A.P., 1996, Near surface applications
of borehole radar, in: Symposium on the Application of Geophysics to Engineering and
Environmental Problems, pp. 545-553.

Goforth, T., and Hayward, C., 1992, Seismic reflection investigations of a bedrock
surface buried under alluvium, Geophysics, vol. 57, pp. 1217-1227.

Granli, J.R., Arntsen, B., Sollid, A., and Hilde, E., 1999, Imaging through gas-filled
sediments using marine shear-wave data, Geophysics, vol. 64, pp. 668-677.

365
Gray, R.E. and Meyers, J.F., 1970, Mine subsidence and support methods in Pittsburgh
area, Journal of the Soil Mechanics and Foundation Division, Proceedings of the
American Society of Civil Engineers, July 1970, SM 4, pp. 1267-1287.

Gutenburg, B., 1944, Energy ratio of reflected and refracted seismic waves, Bulletin of
the Seismological Society of America, vol. 34, pp. 85-102.

Guy, E.D., and Radzevicius, S.J., 2001, Recognition of borehole radar cable-related
effects using variable offset sounding, Subsurface Sensing Technologies and
Applications: An International Journal, vol. 2, no. 2, pp. 127-139.

Guy, E.D., Radzevicius, S.J., and Frank, G., 2001, Non-geologic events in single- and
cross-hole radar data, in: Annual Meeting Society of Exploration Geophysicists
Expanded Abstracts, 4 p.

Haeni, F.P., Halleux, L., Johnson, C.D., and Lane, J.W., Jr., 2002, Detection and
mapping of fractures and cavities using borehole radar, in: Fractured Rock 2002
Proceedings, National Ground Water Association, 4 p.

Hales, A.L., and J.L., Roberts, 1974, The Zoeppritz equations: more errors, Bulletin of
the Seismological Society of America, vol. 64, pp. 285.

Harris, J.B., Miller, R.D., and Xia, J., 2000, Near-surface shear wave reflection surveys
in the Fraser River Delta, B.C., Canada, in: Annual Meeting Society of Exploration
Geophysicists Expanded Abstracts.

Hasbrouck, W.P., 1990, Results of shear-wave measurements, in: Symposium on the


Application of Geophysics to Engineering and Environmental Problems, Environmental
and Engineering Geophysical Society, Golden, Colorado, p. 201.

Hasbrouck, W.P., and Padget, N., 1982, Use of shear wave seismics in evaluation of
strippable coal resources, Utah Geological and Mineral Survey Bulletin, vol. 118, pp.
203-210.

Hatton, L., Worthington, M.H., and Makin, J., 1986, Seismic data processing: theory and
practice, Blackwell Scientific Publications, Palo Alto, California, 177 p.

Hilterman, F.J., 2001, Seismic amplitude interpretation: short course notes, distinguished
instructor series no. 4, Society of Exploration Geophysicists, Tulsa, Oklahoma.

Hoffman, A.G., Clark, D.M., and Bechtel, T.D., 1995, Abandoned deep mine subsidence
investigation and remedial design, Interstate 70, Guernsey County, Ohio, in: Proceedings
of the 46th Highway Geology Symposium, Charleston, West Virginia, 15 p.

366
Holliger, K., and Bergmann, T., 2002, Numerical modeling of borehole georadar data,
Geophysics, vol. 67, no. 4, pp. 1249-1257.

Holzschuh, J., 2002, Low-cost geophysical investigations of a paleochannel aquifer in the


Eastern Goldfields, Western Australia, Geophysics, vol. 67, pp. 690-700.

Hunter, J.A., Pullan, S.E., Burns, R.A., Gagne, R.M., and Good, R.L., 1984, Shallow
seismic reflection mapping of the overburden-bedrock interface with the engineering
seismograph - some simple techniques, Geophysics, vol. 49, pp. 1381-1385.

Jackson, M.J., and Tweeton, D.R., 1993, MIGRATOM - geophysical tomography using
wavefront migration and fuzzy constraints, Report of Investigations 9497, U.S.
Department of the Interior, Bureau of Mines.

Isphording, W.C., 1992, Surface structural damage associated with longwall mining near
Tuscaloosa, Alabama: a case history, Drill Bits, National Drilling Contractors
Association, pp. 23-30.

Kelly, K.R., Ward, R.W., Treitel, S., and Alford, R.M., 1976, Synthetic seismograms: a
finite-difference approach, Geophysics, vol. 41, pp. 2-27.

Kendall, R.R., and Davis, T.L., 1996, The cost of acquiring shear waves, The Leading
Edge, vol. 15, no. 8, pp. 943-944.

Koefoed, O., 1955, On the effect of Poisson’s ratios on the reflection coefficients of plane
waves, Geophysical Prospecting, vol. 3, pp. 381-387.

Koefoed, O., 1962, Reflection and transmission coefficients for plane longitudinal
incident waves, Geophysical Prospecting, vol. 10, pp. 304-351.

Knott, C.G., 1899, Reflexion and refraction of elastic waves, with seismological
applications, Phil. Mag. London, vol. 48, pp. 64-97.

Larson, G.A., 1996, Acquisition, processing, and interpretation of P-P and P-S 3-D
seismic data: M.S. Thesis, University of Calgary, 115 p.

Lay, T., and Wallace, T.C., 1995, Modern global seismology, Academic Press, New
York, 521 p.

Leveson, D., 1980, Geology and the urban environment, Oxford, New York, 369 p.

Macrides, C.G., and Kelamis, P.G., 2000, A 9-C, 2-D land seismic experiment for
lithology estimation of a Permian clastic reservoir, The Leading Edge, vol. 19, no. 10, pp.
1109-1116.

367
Matthews, M.C., Hope, V.S., and Clayton, C.R.I., 1997, The geotechnical value of
ground stiffness determined using seismic methods, in: Modern Geophysics in
Engineering Geology, Geological Society Engineering Geology Special Publication, no.
12, pp. 113-123.

Miller, G.F., and Pursey, H., 1955, On the partition of energy between elastic waves in a
semi-infinite solid, Proceedings of the Royal Society, A233, pp. 55-69.

Miller, R.D., 1992, Normal moveout stretch mute on shallow-reflection data, Geophysics,
vol. 57, pp. 1502-1507.

Miller, R.D., Anderson, N.L., Feldman, H.R., and Franseen, E.K., 1995, Vertical
resolution of a seismic survey in stratigraphic sequences less than 100 m deep in
southeastern Kansas, Geophysics, vol. 60, pp. 423-430.

Miller, R.D., and Steeples, D.W., 1990, A shallow seismic reflection survey in basalts of
the Snake River Plain, Idaho, Geophysics, vol. 55, pp. 761-768.

Miller, R.D., and Xia, J., 1998, Large near-surface velocity gradients on shallow seismic
reflection data, Geophysics, vol. 63, no. 4, pp. 1348-1356.

Miller, R.D., Xia, J., and Park, C.B., 2001, Love waves: a menace to shallow shear wave
reflection surveying, in: Annual Meeting Society of Exploration Geophysicists Expanded
Abstracts, 4 p.

Miller, R.D., and Xia, J., 2002, High-resolution seismic reflection investigation of a
subsidence feature on U.S. Highway 50 near Hutchinson, Kansas, in: Proceedings of the
Symposium on the Application of Geophysics to Engineering and Environmental
Problems, 17 p.

Munk, J., and Sheets, R.A., 1997, Detection of underground voids in Ohio by use of
geophysical methods, U.S. Geological Survey, Water Resources Investigations report 97-
4221, Columbus, Ohio, 28 p.

National Research Council (NRC), 1991, Mitigating losses from land subsidence in the
United States, Washington, D.C., National Academy Press, 58 p.

Muskat, M., and Meres, M.W., 1940, Reflection and transmission coefficients for plane
waves in elastic media, Geophysics, vol. 5, pp. 115-138.

Neidell, N.S., and Taner, M.T., 1971, Semblance and other coherency measures for
multichannel data, Geophysics, vol. 36, pp. 482-497.

Nickel, H., Sender, F., Thierbach, R., and Weichart, H., 1983, Exploring the interior of
salt domes from boreholes, Geophysical Prospecting, vol. 31, pp. 131-148.

368
Nolen-Hoeksema, R.C., 1999, Department of Civil and Environmental Engineering,
University of Michigan, Ann Arbor, Michigan.

Nolen-Hoeksema, R.C., 2001, Department of Civil and Environmental Engineering,


University of Michigan, Ann Arbor, Michigan.

Ohio Department of Natural Resources (ODNR), 1981, Abandoned underground coal


mines map, Guernsey County, Regional Geologic Section, ODNR Division of Geological
Survey, Columbus, Ohio.

ODNR, Division of Geological Survey, 2002, Geophysical Log, Archer Township,


Harrison County, Ohio, permit number 290.

Ohio Department of Transportation (ODOT), 1995a, Interstate 70 grouting report,


Jacksontown, Ohio, November 29, 1995, 3 p.

ODOT, 1995b, Interstate 70 subsidence report, Columbus, Ohio, December 11, 1995, 9 p.

Olhoeft, G.R., 1993, Velocity, attenuation, dispersion, and diffraction hole-to-hole radar
processing, in: Proceedings of the Fourth Tunnel Detection Symposium on Subsurface
Exploration Technology, Golden, Colorado, p. 309-322.

Olsson, O., Falk, L., Forslund, O., Lundmark, L., and Sandberg, E., 1992, Borehole radar
applied to the characterization of hydraulically conductive fracture zones in crystalline
rock, Geophysical Prospecting, vol. 40, pp. 109-142.

Ostrander, W.J., 1984, Plane wave reflection coefficients for gas sands at nonnormal
angles of incidence, Geophysics, vol. 49, pp. 1637-1648.

Pruett, R., 1987, Acquisition, processing, and display conventions for multicomponent
seismic data: Society of Exploration Geophysicists Technical Standards Committee
Report (Subcommittee on 3-C Orientation).

Purnell, G.W., 1992, Imaging beneath a high-velocity layer using converted waves,
Geophysics, vol. 57, no. 11, pp. 1444-1452.

Rahn, P.H., 1996, Engineering geology: an environmental approach, Prentice Hall, New
Jersey, 657 p.

Reynolds, 1997, An introduction to applied and environmental geophysics, John Wiley


and Sons, New York, 796 p.

Richter, C.F., 1958, Elementary seismology, W.H. Freeman and Col, San Francisco,
671 p.

369
Rüger, A., 2002, Reflection coefficients and azimuthal AVO analysis in anisotropic
media, Society of Exploration Geophysicists, 189 p.

Sato, M., and Miwa, T., 2000, Polarimetric borehole radar system for fracture
measurement, Subsurface Sensing Technologies and Applications: An International
Journal, vol. 1, pp. 161-175.

Sato, M., and Thierbach, R., 1991, Analysis of a borehole radar in cross-hole mode, IEEE
Transactions on Geoscience and Remote Sensing, vol. 29, pp. 899-904.

Seriff A.J., and Kim, W.H., The effect of harmonic distortion in the use of vibratory
surface sources, Geophysics, vol. 35, pp. 234-246.

Shearer, P.M., 1999, Introduction to seismology, Cambridge University Press, New York,
260 p.

Sheriff, R.E., 1999, Encyclopedia of Exploration Geophysicists, Society of Exploration


Geophysicists, Tulsa, Oklahoma, 384 p.

Sheriff, R.E., and Geldart, L.P., 1982, Exploration seismology, vol. 1: history, theory,
and data acquisition, Cambridge University Press, New York, 253 p.

Shuey, R.T., 1985, A simplification of the Zoeppritz equations, Geophysics, vol. 50, pp.
609-614.

Shumway, M., 2000, Correlation of SPT N values with cross-hole seismic shear wave
velocity measurements, M.S. Report, Department of Civil and Environmental
Engineering and Geodetic Science, The Ohio State University, 25 p.

Singh, S.J., Ben-Menahem, A., and Shimshoni, M., 1970, Comments on papers by
Costain et al. and McCamy et al. on the solutions of Zoeppritz’ amplitude equations,
Bulletin of the Seismological Society of America, vol. 60, pp. 277-280.

Slotboom, R.T., 1990, Converted wave (P-SV) moveout estimation, in: 60th Annual
International Meeting of the Society of Exploration Geophysicists Expanded Abstracts,
pp. 1104-1106.

Speck, R.C., and Bruhn, R.W., 1995, Non-uniform mine subsidence ground movement
and resulting surface-structure damage, Environmental and Engineering Geoscience, vol.
1, pp. 61-74.

Spratt, R.S., Goins, N.R., and Fitch, T.J., 1993, Pseudo-shear – the analysis of AVO, in:
Offset-Dependent Reflectivity - Theory and Practice of AVO Analysis, Investigations in

370
Geophysics Series, Volume 8, Society of Exploration Geophysicists, Tulsa, Oklahoma,
pp. 37-56.

Steeples, D.W., 1998, Near-surface seismology, short course notes, Society of


Exploration Geophysicists, Tulsa, Oklahoma.

Steeples, D.W., Green, A.G., McEvilly, T.V., Miller, R.D., Doll, W.E., and Rector, J.W.,
1997, A workshop examination of shallow seismic reflection surveying, The Leading
Edge, vol. 16, no. 11, pp. 1641-1647.

Steeples, D.W., Knapp, R.W., and McElwee, C.D., 1986, Seismic reflection
investigations of sinkholes beneath Interstate Highway 70 in Kansas, Geophysics, vol.
51, no. 2, pp. 295-301.

Steeples, D.W., and Miller, R.D., 1998, Avoiding pitfalls in shallow seismic reflection
surveys, Geophysics, vol. 63, pp. 1213-1224.

Stefanic, V., 2000, 4-C projects harvest gulf data, Explorer, August edition, American
Association of Petroleum Geologists, pp. 4-7.

Stewart, R. R., 1991, Exploration seismic tomography: fundamentals, Society of


Exploration Geophysicists, Course Note Series, vol. 3, 200 p.

Stewart, R.R., and Gaiser, J.E., 1999, Converted-wave seismic exploration, short course
notes, Society of Exploration Geophysicists, Tulsa, Oklahoma, 400 p.

Stewart, R.R., and Lawton, D.C., 1999, 3C-3D Seismic polarity definitions, in: Stewart,
R.R., and Gaiser, J.E., Converted-wave seismic exploration, short course notes, Society
of Exploration Geophysicists, Tulsa, Oklahoma, pp. 184-192.

Stewart, R.R., Gaiser, J.E., Brown, R.J., and Lawton, D.C., 2002, Converted-wave
seismic exploration: methods, Geophysics, vol. 67, no. 5, pp. 1348-1363.

Taner, M.T., and Koehler, F., 1969, Velocity spectra - digital computer derivation and
applications of velocity functions, Geophysics, vol. 34, pp. 859-881.

Tatham, R. H., and McCormack, M. D., 1991, Multicomponent seismology in petroleum


exploration: Vol. 6 of Investigations in Geophysics, Society Of Exploration
Geophysicists, Tulsa, Oklahoma, 248 p.

Telford, W.M., Geldart, L.P., and Sherrif, R.E., 1990, Applied geophysics, second
edition, Cambridge University Press, New York, 770 p.

Tooley, R.D., Spencer, T.W., and Sagoci, H.F., 1965, Reflection and transmission of
plane compressional waves, Geophysics, vol. 30, pp. 552-570.

371
United States Department of the Interior (USDI), 1935, Map of the Murray Hill Mine of
the Akron Coal Company, Office of Surface Mining, Pittsburgh, Pennsylvania.

van der Veen, M., and Green, A.G., 1998, Land streamer for shallow seismic data
acquisition: evaluation of gimbal-mounted geophones, Geophysics, v. 63, pp. 1408-1413.

Waltham, A.C., 1989, Ground subsidence, Blackie & Son Limited, London, 202 p.

Waters, K.H., 1987, Reflection seismology: a tool for energy resource exploration, John
Wiley and Sons, Inc., New York, 538 p.

West, M., and Menke, W., 2000, Fluid-induced changes in shear velocity from surface
waves, in: Proceedings of the Symposium on the Application of Geophysics to
Engineering and Environmental Problems, pp. 21-28.

Whittaker, B.N., and Reddish, D.J., 1989, Subsidence: occurrence, prediction and
control, Elsevier, New York, 528 p.

Widess, M.B., 1973, How thin is a thin bed?, Geophysics, vol. 38, no. 6, pp 1176-1180.

Winokur, R.S., and Bohn, J.C, 1968, Sound reflection from a low-velocity bottom, The
Journal of The Acoustical Society of America, vol. 44, no. 4, pp. 1130-1138.

Wolfe, P.J., Richard, B.H., and Holdeman, T.G., 1989, Seismic waveguide characteristics
of an eastern Ohio coal seam, Transactions of the Society of Mining Engineers of AIME,
vol. 284, pp. 1831-1834.

Yilmaz, Ö, 2001, Seismic data analysis: processing, inversion, and interpretation of


seismic data, Society of Exploration Geophysicists, Tulsa, Oklahoma, 2027 p.

Young, G.B., and Braile, L.W., 1976, A computer program for the application of
Zoeppritz’s amplitude equations and Knott’s energy equations, Bulletin of the
Seismological Society of America, vol. 66, pp. 1881-1885.

Zhang, S., 1990, Shallow SH-wave seismic exploration for subway construction in
Shanghai, in: Ward, S.H. Ed., Geotechnical and Environmental Geophysics, vol. 3,
Society of Exploration Geophysicists, Tulsa, Oklahoma, pp. 175-179.

Ziomek, L.J., 1995, Fundamentals of acoustic field theory and space time signal
processing, CRC Press, Ann Arbor, 692 p.

Zoeppritz, K., 1919, Über reflexion and durchgang seismicscher Wellen durch
Unstetigkerlsfläschen: Berlin, Über Eerdbebenwellen VII B, Nachrichten der
Königlichen Gesellschaft der Wissenschaften zu Göttingen, Math-Phys., K1, pp. 57-84.

372

You might also like