Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Acquisition of Active Multichannel Analysis of Surface Waves (MAS

Download as pdf or txt
Download as pdf or txt
You are on page 1of 173

Scholars' Mine

Doctoral Dissertations Student Theses and Dissertations

Spring 2017

Acquisition of active multichannel analysis of surface waves


(MASW) data in karst terrain
Ghassan Salem Alsulaimani

Follow this and additional works at: https://scholarsmine.mst.edu/doctoral_dissertations

Part of the Geological Engineering Commons, and the Geophysics and Seismology Commons
Department: Geosciences and Geological and Petroleum Engineering

Recommended Citation
Alsulaimani, Ghassan Salem, "Acquisition of active multichannel analysis of surface waves (MASW) data
in karst terrain" (2017). Doctoral Dissertations. 2735.
https://scholarsmine.mst.edu/doctoral_dissertations/2735

This thesis is brought to you by Scholars' Mine, a service of the Missouri S&T Library and Learning Resources. This
work is protected by U. S. Copyright Law. Unauthorized use including reproduction for redistribution requires the
permission of the copyright holder. For more information, please contact scholarsmine@mst.edu.
ACQUISITION OF ACTIVE MULTICHANNEL ANALYSIS OF SURFACE WAVES
(MASW) DATA IN KARST TERRAIN

by

GHASSAN SALEM ALSULAIMANI

A DISSERTATION

Presented to the Faculty of the Graduate School of the

MISSOURI UNIVERSITY OF SCIENCE AND TECHNOLOGY

In Partial Fulfilment of the Requirements for the Degree

DOCTOR OF PHILOSOPHY

in

GEOLOGICAL ENGINEERING

2017

Approved by

Neil L. Anderson, Advisor


J. David Rogers
Kelly Liu
Stephen S. Gao
Jeffery D. Cawlfield
© 2017

Ghassan Salem Alsulaimani

All Rights Reserved


iii

ABSTRACT

This study was designed to verify the effects and data reproducibility when the

length of receiver array, receiver spacing, source offset and array orientation parameters

are changed for data acquired using multichannel analysis of surface waves (MASW), at

intended target depth of 30ft (9m), and to compare the results with electrical resistivity

tomography (ERT) data obtained for the same study site.

The MASW data acquired for 34 sites, along four profiles for each site using

variable source offsets of 10ft (3m) and 30ft (9.1m), and variable receiver spacings of

2.5ft (0.76m) and 5.0ft (0.76m), concurrently. Out of the 272 profiles studied, 136

profiles were oriented east-west, and 136 profiles were oriented north-south. The MASW

data was used in conjunction with ERT data to ensure the accuracy of the ERT data.

The comparative analysis indicated the profile configuration measurements have

significant influence on the quality of the data and that the best inversion analysis is

obtained when the dispersion curves are created using the north-south oriented arrays.

The MASW survey study concluded that the most consistent and beneficial karst

terrain dispersion images were those obtained from the predicted optimal acquisition,

using receiver spacing (dx) = 2.5ft, source offset (X1) =10ft and depth of investigation of

about 30ft.
iv

ACKNOWLEDGEMENTS

First of all, I dedicate everything to God Almighty, who made all things possible.

I sincerely extend my gratefulness to my advisor, Dr. Neil L. Anderson, for the

guidance he provided as well as the understanding and patience. Through his mentorship,

I have availed with a well-grounded experience with the consistency of my research

objectives. He further provided me with encouragement not only to broaden my

experimentalist approach, but also to harness my independent thought. The

overwhelming contributions of my committee members, including Dr. J. David Rogers,

Dr. Kelly H. Liu, Dr. Stephen S. Gao and Dr. Jeffery D. Cawlfield, are also greatly

thanked for their generous support.

I would like to thank the Minister of Energy and Industry and Mineral Resources,

Saudi Arabia, Mr. Khalid Al-Falih. I thank the Saudi Geological Survey President, Dr.

Zohair Nawab, for his support as well.

I would like to extend my gratitude and eternal thanks to my father Salem, my

mother Amnah, my sisters, and my brothers for instilling in me the ability to pursue my

ambitions with their love and encouragement. I am also exceedingly grateful to my

father-in-law, Khalid and mother-in-law, Hayat for their support and prayer.

This work is dedicated to my loving wife, Kholood. Without your love and

sacrifice, I would not be achieving this milestone and encouragement throughout the

whole period of my study, and also to my beloved kids, Salem and Juwan.

Finally, special thanks go to the efforts of Mr. Rafat Ghandoura and Dr. Adel

Kerry. Without them, I would not have been able to complete the project successfully.
v

TABLE OF CONTENTS

Page

ABSTRACT ....................................................................................................................... iii

ACKNOWLEDGEMENTS ............................................................................................... iv

LIST OF ILLUSTRATIONS ............................................................................................. ix

LIST OF TABLES ........................................................................................................... xvi

SECTION

1. INTRODUCTION .............................................................................................. 1

1.1. OVERVIEW .........................................................................................1

1.2. RESEARCH OBJECTIVES .................................................................1

1.3. STRUCTURE OF THE DISSERTATION ...........................................2

1.4. LITERATURE REVIEW .....................................................................3

1.5. SIGNIFICANCE OF THE RESEARCH ..............................................5

2. GEOLOGY OF SOUTHWEST MISSOURI ...................................................... 7

2.1. INTRODUCTION ................................................................................7

2.2. OVERVIEW OF GEOLOGICAL AND STRATIGRAPHIC


SUCCESSION OF MISSISSIPPIAN SYSTEM IN SOUTHWEST
MISSOURI ...........................................................................................8

2.2.1. Lower Mississippian of Lower Osagean............................. 11

2.2.2. Lower Mississippian of Uperer Osagean . .......................... 11

2.2.3. Overview of Structure and Faulting Based on the


Work of Coots (2007)......................................................... 13

2.3. FORMATION OF KARST FEATURES ...........................................14

2.3.1. Groundwater Recharge. ...................................................... 16

2.3.2. Sinkhole or Sink.................................................................. 16


vi

2.3.3. Soluble Bedrock. ................................................................. 16

2.3.4. Natural Bridge or Tunnel. ................................................... 17

2.3.5. Losing Stream. .................................................................... 17

2.3.6. Cave. ................................................................................... 17

2.3.7. Spring. ................................................................................. 18

2.4. OVERVIEW OF SINKHOLE FORMATION PROCESS .................18

2.4.1. Collapse Sinkholes. ............................................................. 18

2.4.2. Solution-Sinkholes. ............................................................. 19

2.4.3. Cover-Subsidence Sinkhole. ............................................... 19

2.5. KARST TOPOGRAPHY IN MISSOURI ..........................................21

3. GEOPHYSICAL INVESTIGATIONS ............................................................. 23

3.1. MULTICHANNEL ANALYSIS OF SURFACE WAVE (MASW)..23

3.1.1. Overview. ............................................................................ 23

3.1.2. Basic Concept.. ................................................................... 24

3.1.3. Seismic Waves. ................................................................... 24

3.1.4. Body Waves. ....................................................................... 24

3.1.5. Surface Waves. ................................................................... 25

3.1.6. Rayleigh Waves in Homogeneous Elastic Half-Space ....... 27

3.1.7. Rayleigh Waves in Vertically Heterogeneous Elastic


Half-Space. ......................................................................... 32

3.2. MULTICHANNEL ANALYSIS OF SURFACE WAVES


DATA ACQUISITION...…...……………………………………….35

3.2.1. Geophone Spread Length. ................................................... 38

3.2.2. Source Offset. ..................................................................... 39


vii

3.2.3. Receiver Spacing ................................................................ 42

3.2.4. Orientation .......................................................................... 43

3.2.5. Topographical Conditions................................................... 44

3.2.6. Resolution of MASW Data. ................................................ 45

3.3. MULTICHANNEL ANALYSIS OF SURFACE WAVES


DATA PROCESSING ........................................................................46

3.3.1. Dispersion Analysis. ........................................................... 47

3.4. MULTICHANNEL ANALYSIS OF SURFACE WAVES


DATA INTERPRETATION ..............................................................50

3.5. ELECTRICAL RESISTIVITY TOMOGRAPHY (ERT) ..................51

51
3.5.1. Overview. ............................................................................ 57

3.5.2. Ohm’s Law and Resistivity................................................. 59


52

3.5.3. Relationship between Geology and Resistivity. ................. 57

3.5.4. Electrical Resistivity Array Configuration ......................... 59

3.5.5. Wenner Array...................................................................... 60

3.5.6. Schlumberger Array ............................................................ 62

3.5.7. Dipole-Dipole Array. .......................................................... 62

3.6. ELECTRICAL RESISTIVITY TOMOGRAPHY


DATA ACQUISITION… ...................................................................63

3.7. ELECTRICAL RESISTIVITY TOMOGRAPHY


DATA PROCESSING ........................................................................66

3.8. ELECTRICAL RESISTIVITY TOMOGRAPHY


DATA INTERPRETATION ..............................................................68

4. RESEARCH EXPERIMENTS ......................................................................... 70

4.1. STUDY SITE ......................................................................................70


viii

4.2. MULTICHANNEL ANALYSIS OF SURFACE WAVES


(MASW)……. ....................................................................................72

4.2.1. Data Acquisition. ................................................................ 72

4.2.2. Data Processing and Interpretation.. ................................... 74

4.3. ELECTRICAL RESISTIVITY TOMOGRAPHY (ERT) ..................81

4.3.1. Data Acquisition. .................................................................81

4.3.2. Data Processing and Interpretation. .....................................81


82

5. COMPARATIVE ANALYSES ........................................................................ 89

5.1. QUALITATIVE COMPARISON ......................................................89

5.1.1. Comparison of Overtone Images and Dispersion Curves. .. 89

5.1.2. MASW Traverse No. 1 Oriented West-East ....................... 89

5.1.2. MASW Traverse No. 1 Oriented North to South. .............. 97

5.1.3. MASW Traverse No. 2 Oriented West-East. .................... 106

5.1.4. MASW Traverse No. 2 Oriented North-South. ................ 114

5.2. DISPERSION OF RESOLUTION CURVE .....................................123

5.3. COMPARISON OF 1-D SHEAR-WAVE VELOCITY


PROFILES….. ..................................................................................132

5.4. QUANTITATIVE ANALYSES .......................................................136

6. CONCLUSIONS............................................................................................. 142

7. RECOMMENDATIONS ................................................................................ 143

BIBLIOGRAPHY ........................................................................................................... 144

VITA…………. .............................................................................................................. 156


ix

LIST OF ILLUSTRATIONS
Page

Figure 2.1. Surface Geology of Missouri............................................................................ 7

Figure 2.2. Regional Distribution of the Mississippian System in Missouri. ..................... 8

Figure 2.3. Geological Map of Southwest Missouri. .......................................................... 9

Figure 2.4. Stratigraphic Column for southwestern Mississippian System in


Southwest Missouri. ...................................................................................... 12

Figure 2.5. Faults and Lineament in Southwest Missouri (Coots, 2007). ........................ 13

Figure 2.6. Figure Showing Karst Topography Features.. ................................................ 16

Figure 2.7. Buildings collapse into a sinkhole at Disney World in, Florida.. ................... 20

Figure 2.8. Buildings collapse into a sinkhole at the Summer Bay Resort in
Clermont, Florida. ......................................................................................... 20

Figure 3.1. Elastic deformations and ground particle motions associated with the
passage of body waves. (a) P-wave. (b) S-wave (Bolt, 1982)....................... 25

Figure 3.2. (a) Rayleigh wave motion. (b) Love wave motion. ........................................ 26

Figure 3.3. Distribution of compressional, shear and Rayleigh waves generated


by a point source in a homogeneous half-space, isotropic, elastic half-
space. ............................................................................................................. 28

Figure 3.4. Displacement amplitude of Rayleigh waves versus dimensionless


depth .............................................................................................................. 30

Figure 3.5. Rayleigh wave components with different Rayleigh wavelengths


propagating through a layered medium. Wave components with
different frequencies reflect soil properties at diverse depths. ...................... 35

Figure 3.6. The instrumentation used in the MASW tomography survey. ....................... 38

Figure 3.7. Refraction arrivals showing the shingling degree pattern often
associated with layers. ................................................................................... 40

Figure 3.8. Topographical conditions are found to have an effect on the quality of
the recorded multichannel surface wave data. Receivers should be
placed on relatively flat terrain for optimum results (A & B). Surface
reliefs greater than 0.1d and the recorded data (C & D). .............................. 44
x

Figure 3.9. (A) Acquisition seismic time series data; (B) Dispersion curves
extraction frequency and phase velocity; and (C) 1-D shear-wave
velocity profiles Vs. 1-D depth curve. ........................................................ 48

Figure 3.10. Equipotential and current lines for a pair of current electrodes, A and
B, in a homogeneous half-space. ................................................................... 55

Figure 3.11. The resistivity of rocks, soils, and minerals. ................................................ 59

Figure 3.12. These are some commonly used electrode arrays and their geometric
factors. Note that for the multiple gradient arrays, the total array
length is‘(s + 2) an’, the distance between the center of the potential
Dipole pair is P1-P2 and the center of the current pair C1-C2 is given
by ‘ma.’ 'British Geological Survey (c) NERC 2013. K = Geometric
Factor. ............................................................................................................ 61

Figure 3.13. SuperSting R8/IP resistivity (left). Switch Box (AGI) connecting
passive cables (middle). ERT field setup (right). .......................................... 65

Figure 3.14. An example of a field data set with a few bad data points. The
apparent resistivity data in (a) pseudosection form and in (b) profile
form.. ............................................................................................................. 67

Figure 4.1. Southwest Missouri map where the study was conducted. ............................ 71

Figure 4.2. Map showing the location of the study area Zone A and Zone B. ................. 71

Figure 4.3. Active MASW data acquisition. ..................................................................... 73

Figure 4.4. Schematic diagram of MASW data acquisition setup. ................................... 74

Figure 4.5. The multichannel seismic records with overlay of preliminary data
quality evaluation provided by the processing software (SurfSeis4). ........... 75

Figure 4.6. The multichannel seismic records with overlay of preliminary data
quality evaluation provided by the processing software (SurfSeis4). ........... 75

Figure 4.7 The multichannel seismic records with overlay of preliminary data
quality evaluation provided by the processing software (SurfSeis4). ........... 76

Figure 4.8. (A) The multichannel seismic record collected with the source offset
of 10ft. (3.04m) and receiver spacing of 2.5ft. (0.76m); (B)
Corresponding overtone image with the fundamental and higher
modes identified.. .......................................................................................... 75
78
xi

Figure 4.9 (A) The multichannel seismic record collected with offset of 10ft.
(3.04 m) and receiver spacing of 2.5ft. (0.76 m) with applied muting.
(B) Corresponding overtone image with the fundamental and higher
modes identified.. .......................................................................................... 79
76

Figure 4.10. The results of the borehole BH1. In the same graph, the 1-D shear
wave velocity profile deduced from the MASW survey along the
Traverse No. 5 is shown. The MASW method estimated with
accuracy the shear wave velocity in the layers and assisted in the
layer identification of all the neighboring traverses... ................................... 80

Figure 4.11 ERT field set up... .......................................................................................... 82

Figure 4.12. Typical example of 2-D uninterpreted ERT model. Traverse No. P1
and three overlapping ERT traverses acquired at the study site along
2,440ft............................................................................................................ 83

Figure 4.13 Typical example of 2-D uninterpreted ERT model. Traverse No. P2
and three overlapping ERT traverses acquired at the study site along
2,440ft............................................................................................................ 83

Figure 4.14. 2-D Interpreted versions of resistivity Traverse No. P1. Depth to top
of bedrock is around 16 ft. and corresponds to the 125 ohm-m contour
interval... ........................................................................................................ 84

Figure 4.15 2-D interpreted versions of resistivity Traverse No. P2. Depth to top
of bedrock varies from 10 to 19 ft. and corresponds to the 125 ohm-m
contour interval.............................................................................................. 85

Figure 4.16. Interpreted versions of resistivity Traverse No. P1. The top of rock
correlates reasonably well with the 125 ohm-m contour interval. The
borehole location has been superposed in a red dashed line. The
200ft. mark on the resistivity profile corresponds with MASW profile
No.1; the 800ft. mark corresponds with profile No. 2 location; the1,
400ft. mark corresponds with profile No.3; the1, 800ft. mark
corresponds with profile No. 4; and the 2,200ft. mark corresponds
with profile No.5.. ......................................................................................... 86

Figure 4.17 (A) ERT Traverse No. P1 tied with MASW profile No. 1. at the 200ft.
(61m) mark; (B) Corresponding 1-D shear-wave velocity profile.
MASW depth to top of weathered rock (“acoustic” top of rock) is
identified at 15ft. (4.6m) depth. Red color line on Figure 8.8 (B)
indicates interpreted depth to top of rock... ................................................... 87
xii

Figure 4.18. (A) ERT Traverse No. P1 tied with MASW profile No. 2.at the 800ft.
(244m) mark; (B) Corresponding1-D shear-wave velocity profile.
MASW depth to top of weathered rock (“acoustic” top of rock) is
identified on 7ft. (2.1m) depth. Red color line on Figure 8.9 (B)
indicates interpreted depth to top of rock.... .................................................. 87

Figure 4.19 (A) ERT Traverse No. P1 tied with MASW profile No. 5 and
borehole (BH1), at the 2,200ft. (670m) mark; (B) Corresponding 1-D
shear-wave velocity profile. MASW depth to top of weathered rock
(“acoustic” top of rock) is identified at 8.5ft. (2.6m) depth. Red color
line on Figure 8.10 (B) indicates interpreted depth to top of rock.... ............ 88

Figure 5.1. MASW survey data collected along Traverse No.1 oriented west-east
array. (A) The 24 channel record (shot gather); (B) The dispersion
image (overtone) with superposed dispersion curve (phase velocity
versus frequency) where the fundamental mode is quite clear. (C) The
1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ........................................................................ 91

Figure 5.2. MASW survey data collected along Traverse No.1 oriented west-east
array. (A) The 24 channel record (shot gather); (B) The dispersion
image (overtone) with superposed dispersion curve (phase velocity
versus frequency) where the fundamental mode is quite clear. (C) The
1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ........................................................................ 93

Figure 5.3. MASW survey data collected along Traverse No.1 oriented west-east
array. (A) The 24 channel record (shot gather); (B) The dispersion
image (overtone) with superposed dispersion curve (phase velocity
versus frequency) where the fundamental mode is quite clear. (C) The
1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ........................................................................ 95

Figure 5.4. MASW survey data collected along Traverse No.1 oriented west-east
array. (A) The 24 channel record (shot gather); (B) The dispersion
image (overtone) with superposed dispersion curve (phase velocity
versus frequency) where the fundamental mode is quite clear. (C) The
1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ........................................................................ 97
xiii

Figure 5.5. MASW survey data collected along Traverse No. 1 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency) where the fundamental mode is quite clear.
(C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ........................................................................ 99

Figure 5.6. MASW survey data collected along Traverse No. 1 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency) where the fundamental mode is quite clear.
(C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ...................................................................... 101

Figure 5.7 MASW survey data collected along Traverse No. 1 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency) where the fundamental mode is quite clear.
(C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ...................................................................... 103

Figure 5.8. MASW survey data collected along Traverse No. 1 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency) where the fundamental mode is quite clear.
(C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N);
3) R.M.S. error; 4) Iterations. ...................................................................... 105

Figure 5.9. MASW survey data collected along Traverse No. 2 oriented west-east
array. (A) The 24 channel record (shot gather); (B) The dispersion
image (overtone) with superposed dispersion curve (phase velocity
versus frequency); (C) The 1-D shear wave velocity profile, deduced
from the inversion technique. 1) Risk of higher mode; 2) The signal-
to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. ..................................... 108

Figure 5.10. MASW survey data collected along Traverse No. 2 oriented west-
east array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 110
xiv

Figure 5.11. MASW survey data collected along Traverse No. 2 oriented west-
east array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 112

Figure 5.12. MASW survey data collected along Traverse No. 2 oriented west-
east array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 114

Figure 5.13. MASW survey data collected along Traverse No.2 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 116

Figure 5.14. MASW survey data collected along Traverse No.2 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 118

Figure 5.15. MASW survey data collected along Traverse No.2 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 120

Figure 5.16. MASW survey data collected along Traverse No.2 oriented north to
south array. (A) The 24 channel record (shot gather); (B) The
dispersion image (overtone) with superposed dispersion curve (phase
velocity versus frequency); (C) The 1-D shear wave velocity profile,
deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations. .......................... 122

Figure 5.17. The multichannel seismic record; (B) Corresponding overtone image
in which the fundamental mode dispersion is identified; (C) The
multichannel seismic record collected with doubled receiver spacing;
(D) Corresponding overtone image in which the fundamental mode
dispersion is identified with noticeable discrepancies.. .............................. 130
126
xv

Figure 5.18. (A) The multichannel seismic record; (B) Corresponding overtone
image in which the fundamental mode dispersion is identified; (C)
The multichannel seismic record collected with doubled receiver
spacing; (D) Corresponding overtone image in which the fundamental
mode dispersion is identified with visible distortion..……………………135 128

Figure 5.19. Dispersion images (overtones) deduced from the surface wave
records collected using various MASW array configurations (north to
south orientation).. .......................................................................................130
135

Figure 5.20. Dispersion images (overtones) deduced from the surface wave
records collected using various MASW array configurations (west to
east orientation).. .........................................................................................131
120

Figure 5.21. Comparison of 1-D shear wave velocity profiles, deduced using the
inversion technique from the seismic data sets collected with various
configurations of active MASW oriented west to east array…..…………122 135

Figure 5.22. Comparison of 1-D shear wave velocity profiles, deduced using the
inversion technique from the seismic data sets collected with various
configurations of active MASW oriented north to south array... ................135
130

Figure 5.23. Overtone images obtained from MASW data of (A) good quality, (B)
fair quality, and (C) poor quality... ..............................................................137
130

Figure 5.24. Histogram showing how often MASW data of good, fair, poor, and
severe quality were acquired using a specific array configuration
(west to east direction)... .............................................................................140
135

Figure 5.25. Histogram showing how often MASW data of good, fair, poor, and
severe quality were acquired using a specific array configuration
(north to south)... .........................................................................................141
135
xvi

LIST OF TABLES

Page

Table 2.1. Geologic and Stratigraphic Units in southwest Missouri. ............................... 11

Table 3.1. Shear wave velocity (Vs) of some earth materials. ............................................. 51

Table 3.2. Resistivity of common earth materials. ........................................................... 68

Table 4.1. Survey parameters used during MASW surveys in this study......................... 73

Table 5.1. Results from all the dispersion curves and generated 1-D shear wave
velocity profiles for MASW Traverse No. 1. X1 - source offset; dx -
geophone spacing; D - receiver spread length; 𝜆𝑚𝑎𝑥 maximum
resolvable Rayleigh wave wavelength; 𝜆𝑚𝑖𝑛 minimum resolvable
Rayleigh wave wavelength; Z max penetration depth; S/N - signal-to-
noise ratio; RMS -root-mean-square error... ................................................. 90

Table 5.2. Results from all the dispersion curves and generated 1-D shear wave
velocity profiles for MASW Traverse No. 2. X1 - source offset; dx -
geophone spacing; D - receiver spread length; 𝜆𝑚𝑎𝑥 maximum
resolvable Rayleigh wave wavelength; 𝜆𝑚𝑖𝑛 minimum resolvable
Rayleigh wave wavelength; Z max penetration depth; S/N - signal-to-
noise ratio; RMS -root-mean-square error... ............................................... 106

Table 5.3. Average Vs at ten 1-D shear-wave velocity profile No.1. ............................. 135
133

Table 5.4. Encoded categories for three criterias. ........................................................... 136

Table 5.5. Data quality categories................................................................................... 138

Table 5.6. Quantitative analysis performed for 136 MASW profiles acquired in
west to east direction... ................................................................................ 139

Table 5.7. Quantitative analysis performed for 136 MASW profiles acquired in
north to south direction................................................................................ 139
1. INTRODUCTION

1.1. OVERVIEW

Subsurface imaging is very critical to study subsurface in complex karst terrain.

Application of non-destructive geophysical methods is one of the effective ways to

evaluate the subsurface conditions. A number of geophysical methods are currently used

to imaging subsurface in karst terrain, including Electrical Resistivity Tomography

(ERT), Self-potential (SP), Induced Polarization (IP), Ground Penetrating Radar (GPR),

Seismic Refraction, Seismic Reflection, and Multichannel Analysis of Surface Waves

(MASW).

The active Multichannel Analysis of Surface Waves (MASW) technique is

typically used to generate 1-D shear-wave velocity profiles that extend to depths of

approximately 100ft. (30.5m). The use of optimal acquisition parameters is vital to

assuring the generation of the most accurate 1-D shear-wave velocity profiles. The

identification of optimal acquisition parameters can be a challenging task, especially in

complex karst terrain, because the acoustic properties of the subsurface can vary due to

the complex and irregular nature of the bedrock surface, the presence of solution-widened

joints, etc. In this research, four configurations are tested in order to develop optimum

acquisition parameters to achieve better data resolution and the minimum investigation

depth of 30ft. (9m).

1.2. RESEARCH OBJECTIVES

The primary objective of the study is to develop optimum MASW array

configuration that can be used to image the subsurface in complex karst terrain. There is
2

an expectation that the use of optimum MASW array configuration will allow to image to

a depth of 30ft., while providing a good data quality. The following are means to

accomplish this objective:

 Acquire active MASW data along a 16 ERT traverses and using different array

configurations,

 Evaluate quality of MASW data on the basis of overtone images,

 Process and interpret MASW and ERT data using bore hole data,

 Compare the MASW and ERT data interpretations in terms of estimating top of

bedrock,

 Perform a statistical data analysis in order to develop an optimum MASW array

configuration.

1.3. STRUCTURE OF THE DISSERTATION

Sections one contains an introduction to the dissertation. The main objective and

significance of the research are discussed in this chapter. A critical review and analysis of

the relevant literature is also provided.

Section two discusses geology of southwest Missouri, overview of geological and

stratigraphic succession of Mississippian system, formation of karst features, overview of

sinkhole formation process and karst in topography in Missouri and geophysical

investigation in karst terrain.

Section three provides overviews of two geophysical methods – MASW and ERT.

Basic concepts, data acquisition, data processing and data interpretation are discussed for

each of the method.


3

Section four presents the experiments conducted in this research. The description

of the study site is also provided in this chapter.

Section five contains comparative analysis of the data. Discussions with focus on

the optimum MASW array configuration are presented in this chapter.

Section six provides the conclusions of the research.

Section seven summarizes recommendations for further studies.

The references that have been used during the study are listed at the end of the

dissertation.

1.4. LITERATURE REVIEW

Geophysical exploration methods, such as Electrical Resistivity Tomography

(ERT) and Multichannel Analysis of Surface Waves (MASW), have been used to study a

wide range of targets within the ground from discovering the deep structure of the

subsurface at thousands of meters to near surface structures and their properties at depths

of a few tens of meters (Elkrry et al., 2015). These methods have been employed in

various applications in engineering geology (Miller et al., 1999a, 1999b; Torgashov and

Varnavina, 2016; Park et al., 1998, 1999b, 2000; Xia et al., 1998, 1999, 2000a, 2000b).

The MASW and ERT methods have different applications in geotechnical

investigations. Each technique has its own advantages and drawbacks. To increase the

accuracy of geophysical investigation, the combination of methods can be effectively

used to constrain and verify the interpretations, especially for geophysical data collected

in complex karst terrain environment. Although karst terrain is often manifested at the
4

subsurface, it is difficult to determine that without more detailed subsurface

investigations.

The ERT method is used to estimate lateral and vertical variations in ground

resistivity values and can be used to map geologic variations. It has been used for

multiple geotechnical projects, including the imaging of the spatial distribution of

moisture in pavement sections (Buettner et al., 1999), river terrace sand and gravel

deposit reserve characterization and estimation using 3-D ERT arrays for bedrock surface

detection (Chambers et al., 2013), imaging of the subsurface in karst terrain (Torgashov,

2012), and characterization of soils using ERT and geotechnical investigations (Sudha,

2009).

The MASW method is used to evaluate the shear-wave velocity distribution and

arrangement of overburden soil deposits and the bedrock as well. A considerable amount

of research has been published regarding the use of both geophysical methods in complex

terrain environment (Doolittle et al., 1998; Kidanu et al., 2016; Miller et al., 2005;

Thierry et al., 2005; He et al., 2006; Schrott and Sass, 2008). Since the method was

introduced into the geophysical community in late 90s, it has been extensively used for

various applications and researched for its data acquisition and processing strategies

(Park et al., 1999). As noted by Park (2013), shear-wave velocity is a proper indicator of

material stiffness, but for a specific rock types, such as carbonate rocks, shear-wave

velocity is affected by its porosity and pore structure, as noted by Baechle et al., 2009.

Shear-wave velocity can be used to differentiate various types of unconsolidated soils and

bedrock (Odum et al., 2007), as typically bedrock exhibits higher velocities than soil,

fractured rock, or karst terrain. Recent studies have shown that the instrumental
5

configurations used for field data acquisition and processing can affect the dispersion

results obtained from surface wave seismic surveys. Attempts to obtain accurate or more

efficient dispersion curves were established by selection of particular acquisition

parameters and careful processing. The optimum parameters of MASW data acquisition,

including the optimum distance between the source and receiver spread (source offset)

and geophone array length, were discussed by Park, Miller, & Miura (2002), Park and

Carnevale (2010), and by Anderson (2012). (Xu et al., 2006) demonstrated the possibility

to develop optimum field parameters for particular MASW survey if preliminary

information on the phase velocity range and interface depth is available. A successful

attempt to develop a formula for quantitative evaluation of a layered homogeneous elastic

model estimation using specific offset was performed by (Xu et al., 2006). Attempted to

develop acquisition and processing parameters for MASW using 3-D ERT as control. In

environmental studies, the MASW method was employed to generate 2-D shear-wave

velocity fields calculated from inversion of Rayleigh-wave phase velocities, define top-

of-rock and subsurface geological structures from 6 to 100 ft. (1.8 to 30 m) (Xia et al.,

1998; Miller et al., 1999), and to determine a collapse feature in an extremely noisy

environment (Xia et al., 2004a).

1.5. SIGNIFICANCE OF THE RESEARCH

The proposed study is not only significant for imaging in karst terrain in

southwest Missouri, but it also has a broader impact on the social and economic lives of

people at all areas. Some of the significant aspects of this research include:
6

 Developing optimum MASW data acquisition and processing parameters to

enhance overall data quality and increase investigation depth.

 Verifying the efficiency of using integrated geophysical tools (MASW, ERT, and

bore hole) to map complex subsurface systems in karts terrain.

 Improving reliability of geophysical data interpretation in complex karst terrain.

 Detecting probabilities for catastrophic collapse, and therefore, minimizing

pollution and contamination of aquifers.


7

2. GEOLOGY OF SOUTHWEST MISSOURI

2.1. INTRODUCTION

Much of the Missouri state is underlain by carbonate rock (Fig. 2.1.), and most of

it is exposed (Vandike, 1997). Karst-related features are formed in dolomite and

limestone bedrocks in Missouri. The presence of clay and mud and shale with low

permeability curbs downward movement of surface water and slows the advancement of

solution cavities in the underlying carbonate rock. Hence, fewer sinkholes and caves are

found in Northern Missouri compared to Southern and Central Missouri.

Figure 2.1. Surface Geology of Missouri (MSDIS, 2003).


http://oewri.missouristate.edu/assets/OEWRI/KShade-Thesis.pdf.
8

2.2. OVERVIEW OF GEOLOGICAL AND STRATIGRAPHIC SUCCESSION


OF MISSISSIPPIAN SYSTEM IN SOUTHWEST MISSOURI

The stratigraphic succession of rock units in the Mississippian System is not

uniform throughout Missouri. Because of that, Missouri is divided into six zones:

northwestern, east-central, southeastern, central, southwestern, and northwestern. Figure

2.2 highlights that southwest Missouri is located in the southwestern region to facilitate a

description of Mississippian Age rocks.

Figure 2.2. Regional Distribution of the Mississippian System in Missouri (Vineyard,


1992).

The geologic map and units in southwest Missouri are displayed in Figure 2.3.

Each geologic rock unit has characteristics that need be addressed on a site-by-site basis.

Most of the southwest Missouri is composed of Burlington-Keokuk Limestone (Mbk) a


9

material that is particularly prone to dissolve, thereby increasing the probability of

catastrophic collapses, and through which surface water reaches the ground water supply.

Figure 2.3. Geological Map of Southwest Missouri (ESRI data source: Missouri
Geological Survey GEOSTRAT system, Sept 2015).

The Mississippian System is divided into two series: the Kinderhook and Osage.

The Mississippian formations of the southwest Missouri are presented in Table 2.1. The

total thickness of the Mississippian System in this part is about 640ft (195m), (Vineyard,

1992).

The Mississippian System starts with the Compton Formation, which has a

thickness of less than 30ft (9m). The Compton Formation is overlain by the Northview

Formation with an average thickness up to 80ft (24m). It is represented by brown

siltstone and blue or bluish-green shale.


10

The Mississippian System continues with the deposition of the Pierson, Reeds

Spring, and Elsey Formations. The total thickness of these three formations is about 260ft

(79m). Cherty limestone is a dominant component in all three formations.

The deposition of the Mississippian System is finalized by the Burlington-Keokuk

Formation. Further, the youngest exposed rock unit is in southwest Missouri. The

thickness of this formation is approximately between 150 to 270ft (45 to 83m) and varies

in thickness from place to place due to erosion. The majority of springs and caves in

southwest Missouri are found in Burlington-Keokuk Limestone.

According to conventional geotechnical procedures, the collections of boreholes

were drilled in southwest Missouri. The bedrock in the study location is intensely

dissolved Burlington-Keokuk Limestone and can be characterized by the presence of

pinnacles and cutters (Fellows, 1970), meaning that the depth in the bedrock varies

significantly. Furthermore, even though the subsurface information obtained at a location

is highly accurate, the interpolation between boreholes can sometimes be incorrect due to

significant lateral variability in karst terrains.

In the study site, the shallow subsurface is mainly represented by rocks of the

Osagean Series in the Mississippian System. The Mississippian System is represented

Table 2.1, by the Pierson, Reeds-Spring, Elsey, and Burlington-Keokuk Formations.

These four formations have similar lithologic characteristics, and they are sometimes

difficult to differentiate.
11

Table 2.1. Geologic and Stratigraphic Units in southwest Missouri (Vandike, 1993).

The following sections discuss the late two collective units of Mississippian

deposition system as shown in Figure 2.4.

2.2.1. Lower Mississippian of Lower Osagean. Lower Mississippian is Lower

Osagean (Mlo) Elsey, Reeds Spring, and Pierson Formations. The Elsey Formation is

light-gray, crystalline to micritic limestone with chert fragments and some crinoids. The

Reeds Spring Formation is gray to brown; finely crystalline limestone with chert

fragments. The Pierson Formation is brown to brown-gray and comprised of Magnesian

limestone with chert nodules. The fine-grained matrix contains some fossil fragments

(Muchaidze, 2008).

2.2.2. Lower Mississippian of Upper Osagean. Lower Mississippian is Upper

Osagean (Muo) or (Mbk) Burlington, and Keokuk Formation, in southwest Missouri. The

Burlington-Keokuk limestone is the most important formation in this research because

the bedrock in the study site is from this formation. The Burlington Keokuk Formation is

coarsely crystalline, light gray limestone with some fossiliferous limestone and chert
12

nodules. The limestone is comprised almost entirely of crinoid fragments. The

Burlington-Keokuk Formation weathers to a red to reddish brown residual soil that

contains a variable amount of chert fragments. It is almost entirely comprised of pure

calcite. Thus, this formation is susceptible to weathering through the dissolution process.

Uneven dissolution of this formation has resulted in a highly irregular bedrock-

overburden interface (Fellows, 1970).

Figure 2.4. Stratigraphic Column for southwestern Mississippian System in southwest


Missouri (Fellows, 1970).
13

2.2.3. Overview of Structure and Faulting Based on the Work of Coots

(2007). The geologic structures in southwest Missouri are highly jointed with orientations

that are trending N 20º W, and N 60º E and are approximately orthogonal to each other.

Coots (2007) discussed three main fault types that are common in the area, namely, the

Kinser Bridge fault, the Danforth Graben Fault, and the Pearson Creek Fault (Figure 2.5),

all of which have a northwest – southeast trend with normal deracination. The Kinser

Bridge Fault trends west-northwest with an average deracination of about 50ft (15m); the

Danfort Graben Fault has a vertical deracination of about 70ft (21m), and trends

northwestward; and the Pearson Creek Fault has a strike direction of N 55º W and a

normal deracination of between 10 and 20 ft (3 and 6m).

Figure 2.5. Faults and Lineament in Southwest Missouri (Coots, 2007).


14

2.3. FORMATION OF KARST FEATURES

According to the Missouri Department of Natural Resources dictionary, karst is “a

type of topography that is formed on limestone, gypsum, and other rocks by dissolution,

and is characterized by caves, sinkholes, and underground drainage”. Karst terrain is

characterized by the presence of springs, caves, sinkholes, and a unique hydrogeology

(USGS, 2012). Karst is a term derived from the German form of the Slavic word “Kras”

or “Krs,” meaning a bleak waterless place (Monroe, 1970).

Features inherent in karst terrain are caves, sinkholes, springs, dry valleys, and

loosening or disappearing streams and springs. The complex terrain results from the

chemical weathering of carbonate and other soluble rocks, and the formation of karst

features are controlled in part by pre-existing fractures within the bedrock (Ford &

Williams, 2007). Sinkhole formation is also controlled by overburden thickness,

fluctuation of the water table, soil type, and the presence of recharge or discharge zones

(Denizman, 2003).

Karst is formed when rain falls and carbon dioxide in the atmosphere is dissolved

making the rainwater acidic and consequently dissolving soluble rocks. It is worth noting

that the acidic rainwater passes through dead plant materials/debris in the soil and even

becomes more acidic as it percolates through cracks, consequently dissolving the bedrock

(limestone, dolostone, marble, gypsum and salt). Dissolution continues as the water

moves sidelong along bedding planes and joints and fractures in the rock, forming

conduits in the rock.

The carbonate rocks, limestone, and dolostone are predominantly composed of

calcite mineral (CaCO3) and dolomite mineral (CaMg (CO3)2), and both mainly calcite,
15

are susceptible to dissolution when slightly acidic water acts on them. Meteoric water

absorbs carbon dioxide (CO2) from the atmosphere and thus becomes slightly acidic.

After meteoric water saturates the ground, it passes through soil that may increase its

CO2 concentration. At the point where water enters carbonate rock, it reacts with soluble

minerals. Dissolved matter will be washed away, and as a result, features such as

dissolution-widened fractures form. As erosion continues underground, caves get hollow

enough and the roofs thin and eventually collapse to form a sink.

Karst terrain is characterized by caves, sinkholes, underground streams, and other

features formed by the slow dissolution, rather than the mechanical eroding, of bedrock.

People have discovered the difficulties (e.g., sinkhole collapse, sinkhole flooding, and

easily polluted groundwater) of living on those terrains as the population has grown and

expanded into those areas as shown in Figure 2.6, and therefore either serve as recharge

or discharge areas depending on certain geological and hydrogeological factors.


16

Figure 2.6. Figure Showing Karst Topography Features.


(www.springfieldmo.gov/documentcenter/view/11091).

2.3.1. Groundwater Recharge. Groundwater recharge (also called deep drainage

or deep percolation) is a hydrologic process where water moves downward from surface

water to groundwater. Recharge is the primary method through which water enters an

aquifer. This process usually occurs in the vadose zone below plant roots and is often

expressed as a flux to the water table surface. Recharge occurs both naturally through the

water cycle or through anthropogenic processes, such as rainwater, melting snow, or

other surface water.

2.3.2. Sinkhole or Sink. A sinkhole or sink is a collapsed part of bedrock above a

void. They may be a sheer vertical opening into a cave or a shallow depression of many

acres. They can appear suddenly, creating havoc for Ozarks landowners or highway

managers.

2.3.3. Soluble Bedrock. Soluble bedrock is characterized by underground

drainage systems with sinkholes and caves. It has also been documented for weathering-
17

resistant rocks. Subterranean drainage may limit surface water with few to no rivers or

lakes. However, in regions where the dissolved bedrock is covered (perhaps by debris) or

confined by one or more superimposed insoluble rock strata, distinctive karst surface

developments might be totally missing.

In Missouri, karst forms in limestone (calcium carbonate) and dolomite

(magnesium calcium carbonate).

2.3.4. Natural Bridge or Tunnel. Natural bridges may be the remains of a cave

that collapsed, with only a portion of the ceiling still standing. But while the entrance to

caves can be mysterious or even foreboding, natural bridges are more inviting with the

light at the end of the tunnel beckoning a visitor explore. Natural bridges are usually

shorter than a tunnel and air-filled rather than filled partially with water.

2.3.5. Losing Stream. A losing stream is a stream or river that loses water as it

flows downstream. The water infiltrates the ground, recharging the local groundwater

because the water table is below the bottom of the stream channel. Losing streams are

also common in regions of karst topography where the stream water may be completely

captured by an underground cavern system and become a subterranean river.

2.3.6. Cave. A cave is a hollow place in the subsurface, or an air-filled

underground void. A specifically, a cave is natural underground space large enough for a

human to enter and often extends deep underground. The formation and development of a

cave is known as speleogenesis. Caves are formed by various geologic processes and can

be variable sizes. This may involve a combination of chemical processes, erosion from

water, tectonic forces, microorganisms, pressure, and atmospheric influences.


18

2.3.7. Spring. A spring can be developed where surface water has infiltrated the

Earth's surface (recharge area) due to karst topography, thereby becoming part of the

area’s groundwater. The underground water then penetrates through a network of cracks

and fissure openings, ranging from intergranular spaces to large caves. The water

emerges from below the surface in the form of a karst spring. A confined aquifer forces

the spring to appear on the surface land.

2.4. OVERVIEW OF SINKHOLE FORMATION PROCESS

Sinkholes, also called dolines, form in the same manner that caves are formed

(dissolution of soluble rocks by surface water and groundwater to create a void or an

opening in the soil or rock). Sinkholes are the most diagnostic surface expression of karst

landscapes. They are important near-surface indicators of active karst features such as

dissolution-enlarged fractures, caves, and conduits. Sinkholes can vary from a few feet to

hundreds of acres wide and from less than 1 to more than 100 feet deep. Some are

rounded in shape, whereas others have vertical walls. Although most sinkholes drain

rapidly, some have natural plugs and may hold water for many years. There are three

types of sinkholes: Collapse Sinkholes, solution-sinkholes, and cover-subsidence

sinkholes.

2.4.1. Collapse Sinkholes. Collapse Sinkholes are karst-related features that are

not bound by a defined drainage area. Collapse sinkholes form as a void in the soil layer

migrates toward the surface. A sudden collapse occurs when the roof of the cavity can no

longer support its own weight and caves-in, creating an abrupt and dramatic sinkhole. A

collapse sinkhole will commonly develop in the floor of an existing depression sinkhole.
19

Sinkhole collapses are very common in the Ozarks. On average, the southwest Missouri

responds to over 30 reported collapses every year. A sinkhole of this type can be

dangerous if someone were to fall in and should therefore be repaired or fenced off as

soon as possible.

2.4.2. Solution-Sinkholes. Solution-sinkholes occur in areas where carbonate

rocks like limestones and dolostones are exposed at land surface or are covered by thin

layers of soil and permeable sand. Dissolution is most active at the limestone surface and

along joints, fractures or other openings in the rock that permit water to move easily into

the subsurface. When rain falls, surface water percolates through joints in the carbonate

rock and the dissolved rock is carried away from the surface, forming a small depression.

Further drainage of the carbonate surface accelerates the dissolution process and enlarges

the depression. As more debris is carried into the developing sinkhole, water outflow may

be plugged to form a pond and may hold water for years. Solution sinkholes tend to have

gently sloping sides and seldom pose a hazard by collapsing.

2.4.3. Cover-Subsidence Sinkhole. Cover-Subsidence Sinkhole is one of the

most dramatic forms of land subsidence is the collapse of the ground surface into natural

underground caverns. The material can no longer support its own weight, and a sudden

collapse forms this type of sinkhole. Heavy rainfall, drought, or mechanical loading can

trigger this type of sinkhole. The collapse is generally abrupt and can be catastrophic. A

sinkhole formed in this manner is often fairly circular with steeply sloping sides. The

sinkhole that developed in Exeter, Missouri is pictured in Figure 2.7 and Figure 2.8.
20

Figure 2.7. Buildings collapse into a sinkhole at Disney World in, Florida.
(http://www.csmonitor.com/Science/2013/0813/Sinkhole-Disney-World-visitors-
walking-on-holey-ground).

Figure 2.8. Buildings collapse into a sinkhole at the Summer Bay Resort in Clermont,
Florida. (http://www.amusingplanet.com/2013/09/disastrous-sinkholes).
21

In Missouri, this type of sinkhole is more common than the other types based on

the mechanism of its formation and nature of overburden materials. In areas where

limestone is buried beneath a sufficient thickness of unconsolidated material, few

sinkholes generally occur. If the overburden is dense plastic clay, its low permeability

may impede downward movement of surface water and slows the development of

solution cavities in the underlying limestone.

2.5. KARST TOPOGRAPHY IN MISSOURI

Missouri is one of seven states with karst terrain. Approximately 59% of Missouri

is covered by carbonate rock, most of which is exposed (Vandike, 1997). Major karst

areas in Missouri occur in the Mississippian rocks.

Three of the four largest metropolitan areas in Missouri (St. Louis, Springfield,

and Columbia) are located almost entirely on karst terrain. Most of the karst features in

Missouri are advanced in the southwest Missouri and Salem plateaus. However, there are

also karst features north of the Missouri River. Karst terrain is particularly susceptible to

catastrophic collapse. Furthermore, contaminants can also become concentrated in karst

depressions, so knowledge of the bedrock topography can be useful in remediation work

(Vandike, 1997).

The Ozark Plateau Aquifer is a large system comprised of many smaller aquifers

spread over a large geographic region across Oklahoma, Missouri, Kansas, and Arkansas.

This system consists of Ozark aquifers, St. Francois aquifers, and the Springfield Plateau.

It has been the most significant water source for southwest Missouri, northeastern

Oklahoma, southeast Kansas, and northern Arkansas (Macfarlane et al., 2005).


22

The Ozark area is experiencing a significant population growth in southern

Missouri regions; the population grew increased by 11% in 1990 and 27% in 2000,

(Missouri Department of Conservation, 2005). According to the Missouri Department of

Natural Resources, there are over 5,500 caves, more than 9,500 sinkholes, more than

2,800 springs, and loosing streams recorded in Missouri (Vandike, 1997).


23

3. GEOPHYSICAL INVESTIGATIONS

Several techniques have been used to characterize karst terrain. More detailed

investigations of the shallow subsurface often rely on boreholes control to characterize

subsurface conditions, but due to the spatial variability of karst terrain, and information

from individual boreholes may be insufficient for a complete site evaluation. Geophysical

exploration methods have been used in several fields to study a wide range of targets

within the subsurface, such as discovering the shallow structure of subsurface and

properties at depths up to 100ft. (30m).

Geophysical data can provide suitable continuous coverage between boreholes

and significantly decrease ambiguity (subsurface conditions). Numerous geophysical

techniques including multichannel analysis of surface waves (MASW) and electrical

resistivity tomography (ERT) can be employed to characterize the subsurface in complex

terrain like karst regions (Doolittle et al., 1998; Miller et al., 2005; Thierry et al., 2005;

Schrott & Sass, 2008). A combination of methods is often used to constrain the

interpretation because each technique has inherent advantages and drawbacks.

3.1. MULTICHANNEL ANALYSIS OF SURFACE WAVES (MASW)

3.1.1. Overview. Multichannel Analysis of Surface Waves (MASW) is a non-

destructive seismic survey method based on the measurement of seismic surface waves

generated by seismic source (active MASW) or ambient noise from cultural activities

(passive MASW) for evaluation of the elastic condition (stiffness) of the ground for

geotechnical purposes. Nowardays MASW is a commonly used geopgysical technique

because of its effectiveness for providing shear-wave velocities measurements and


24

simplicity of use (Park, 2013). The MASW method has become one of the main surface

wave methods to determine shear-wave velocitiy variations for geophysical and civil

engineering applications with observed difference of approximately 15% error between

the results obtained by MASW and borehole control measurements (Xia et al., 2002; Xia,

2014). Typical MASW survey can be broken down into three steps: data acquisition, data

processing, data interpretation. Final shear-waves profile can be presented in 1D, 2D and

3D formats (Park and Taylor, 2010).

3.1.2. Basic Concept. MASW method utilizes frequency-dependent properties of

particular type of seismic surface waves (fundamental-mode Rayleigh waves)

horizontally travelling along the earth surface directly from the impact point to the

receiver spread. The method provides the shear wave velocity information in 1-D (single

vertical shear wave velocity profile), 2-D (shear-wave velocity cross-section) or 3-D

(interpolation between densely distributed 1-D profiles) formats.

3.1.3. Seismic Waves. In general, seismic waves are parcels of strain energy that

propagate outwards from a seismic source. There are two main categories of seismic

waves that propagate within the ground and along its surface: body waves and surface

waves (Aki & Richards, 2002, Evrett, 2013).

3.1.4. Body Waves. There are two types of body waves: compression waves (the

longitudinal, primary or dilatational) and S-waves (the transverse, secondary, or shear-

wave).

Body waves can propagate through the internal energy of an elastic solid and may

be one of two types. In compressional/dilatational primary (or P-) waves, the particles of

the medium move in the direction of wave travel, involving alternating expansion and
25

contraction of the medium. The particle motion of compressional waves is parallel to the

motion of the wave propagation, causing the dilatation and compression of elementary

volume particles (Aki & Richards, 1980), as shown in Figure 3.1(a)).

Figure 3.1. Elastic deformations and ground particle motions associated with the passage
of body waves. (a) P-wave. (b) S-wave (Bolt, 1982).

3.1.5. Surface Waves. Surface waves are results of interfering P-waves and/or S-

waves travelling primarely along the free surfaces or along the boundary of dissimilar

materials (Kearey, Brooks, & Hill, 2002). They induce particles motion which is

perpendicular to the direction of wave propagation and has both a vertical and a

horizontal component. There are two types of surface waves that are most interesting for

engineering purposes based on their modes of propagation, dispersion velocities and the

depth range of the associated particle motion: Rayleigh waves and Love waves. Rayleigh

waves travel along a boundary with particle motion in vertical direction along the wave

patch and always exist in the presence of a free surface. Love waves induce particles
26

motion which is horizontal and transverse to the direction of wave propagation and only

generated in conditions where a soft layer overlying a stiffer layer (Parasnis, 1997). Since

their particle motion is always horizontal, Love waves are rarely recorded in seismic

surveys where only vertical source and receivers are used (Park et al., 1997), as shown in

Figure 3.2.

They represent the strongest portion of the signal received during a seismic

survey. For the preliminary location, an average value of Vp/Vs = 1.73 of the ground

crust is used. However, Vp/Vs can be determined with a fair degree of accuracy by the

Wadati-plot method (Wadati, 1933).

Vp = 1.7 Vs (3.1)

The speed of wave propagation is NOT the speed at which particles move in

solids (~ 0.01 m/s).

Figure 3.2. (a) Rayleigh wave motion. (b) Love wave motion
(http://thinkgeogeek.blogspot.com).
27

More specifically, in this technique, phase velocities are calculated for each

component frequency of the field-recorded Rayleigh waves. The resulting dispersion

curve (phase velocity vs. frequency) is inverted using a least squares approach, and a

vertical shear-wave velocity profile is generated (Miller et al., 2000; Nazarian et al.,

1983; Stokoe et al., 1994; Park et al., 1999a, 1999b, 2000; Xia et al., 1999).

Rayleigh waves result from the interaction of P-waves and vertical (SV) waves

with the surface of the ground (Aki & Richards, 1980). The particle motion of Rayleigh

waves moves perpendicular to the surface but travels along the wave path (Kearey,

Brooks & Hill, 2002). The wave motion is retrograde (counterclockwise) closest to the

surface, but becomes prograde (clockwise) at greater depths. More than two-thirds of the

seismic energy that is generated is imparted (Loke, 2000).

Particle motion associated with Love waves is parallel to the free surface and

perpendicular to the direction of the waves, whereas Rayleigh waves move perpendicular

to the surface but travel along the wave path (Kearey, Brooks and Hill, 2002). Love

waves are a form of a polarized shear wave and are observed in a multilayer media when

the shear wave velocity of the top layer is less than that of the lower layer (Parasnis,

1997). They are the fastest surface wave, move along the ground from side-to-side and

are confined to the surface of the crust. Since their particle motion is always horizontal,

(Love) waves are seldom recorded in seismic surveying where only vertical sources and

receivers are used (Park et al., 1997).

3.1.6. Rayleigh Waves in Homogeneous Elastic Half-Space. Rayleigh waves

generated by a vertical point source on a free surface propagate along cylindrical

wavefronts away from the impact point (Richart, Hall, & Woods, 1970). Body waves
28

propagate radially, from the exterior of the source, both into the medium and along the

surface, along a hemispherical wavefront, as shown in Figure 3.3. As the waves transfer

outward and encounter an increasingly larger volume of material, their amplitude

decreases.

Figure 3.3. Distribution of compressional, shear and Rayleigh waves generated by a point
source in a homogeneous half-space, isotropic, elastic half-space (Richart et al., 1970).

This is called geometrical spreading (or geometrical damping) (Aki & Richards,

1980; Richart et al., 1970). In a homogeneous elastic half-space, the amplitude of

Rayleigh waves decreases as r-0.5, where r is the distance from the impact point. For

comparison, the amplitude of body waves decreases as r-2 along the surface and r-1 into

the medium. As two-thirds of the total seismic energy is imparted into Rayleigh waves

and Rayleigh waves decay more slowly with distance than body waves, the wave field

becomes dominated by Rayleigh wave motion at nearest distances from the seismic

source (Richart et al., 1970). Around 67% of the seismic energy is imparted into Rayleigh

waves, while about 26% is imparted into shear waves and 7% into compressional waves

(Woods, 1968).
29

The amplitude of Rayleigh waves decays exponentially with depth (Richart et al.,

1970). The simplest method is attribution of a factored shear wave phase velocity

(usually 0.9 times the Rayleigh wave velocity (Joh 1996, Foti 2000 & Okada 2003) to a

depth equivalent to a fraction of the Rayleigh wavelength, λ. Fractional depth factors

range from λ/4 to λ/2 (Jones1958, Ballard & McLean 1975; Abbis, 1981).

Gazetas (1982) recommended that λ/4 be used where the stiffness increases

significantly with depth and that λ/2 be used for more homogeneous stiffness profiles.

However, a factor of λ/3 is most commonly used (Bullen 1963 & Richart et al., 1970)

because a significant proportion of the particle motion in the ground associated with

Rayleigh wave propagation is approximately this depth. The horizontal and vertical

Rayleigh wave displacement amplitudes as a function of dimensionless depth are shown

in Figure 3.4 for several values of Poisson’s ratio (v) = 0.25 (Richart et al., 1970).

Rayleigh waves have unique properties that allow them to be transformed into

subsurface shear-wave velocity profiles (Surf-Seis, 2006). Rayleigh waves are dispersive

in nature (different frequencies travel with different phase velocities). The highest usable

Rayleigh wave frequency that has been recorded (for geotechnical purposes) involves

particle motion within the shallowest depth range (approx. 1 Rayleigh wavelength;

typically upper few ft) and travels with a velocity that is mostly a function of the average

shear-wave velocity within that depth range.

Intermediate frequencies for Rayleigh waves involve particle motions to

intermediate depths (to approx. 1 Rayleigh wavelength), and they travel with velocities

that are a function of the average shear-wave velocity over those intermediate depth

ranges. Also, the lowest recorded usable frequency involves particle motions to the
30

greatest depth (1 Rayleigh wavelength) and travels with a velocity that is a function of

the shear-wave velocity over that depth range.

Figure 3.4. Displacement amplitude of Rayleigh waves versus dimensionless


depth(Richart et al., 1970).

Equations describing Rayleigh waves propagate along the free surface of the earth

with particle motions that decay exponentially with depth. The lower component

frequencies of Rayleigh waves involve particle motions at greater depths. In a

homogeneous (non-dispersive) medium, Rayleigh wave phase velocities are constant.

Rayleigh wave phase velocities are a function of both the shear-wave and the

compression wave velocities of the subsurface. The interrelationships between Rayleigh


31

wave velocities (VR), shear-wave velocities (β) and compression wave velocities (α) in a

uniform medium are expressed in Equation 3.2 (Anderson, 2010):

VR6 - 8β2VR4 + (24 - 16β2 /α2) β4VR2 + 16(β2/α2 – 1) β6 = 0 (3.2)

where

VR is the Rayleigh wave velocity within the uniform medium,

β is the shear-wave velocity within the uniform medium (also denoted as Vs), and

α is the compressional wave velocity within the uniform medium (Also denoted as

Vp).

Although the Raleigh wave phase velocity is a function of both compressional (α)

and shear-wave (β) velocities, it is much more sensitive to variations in β than in α in

Equation 3.3.

VR< β < α (3.3)

Lower frequencies involve particle motion at greater depths, causing VR to also

vary with frequency.

Equation 3.4 might initially suggest that it would be difficult to extract the shear-

wave velocity because the equation contains two unknowns (shear and compression wave

velocities). Fortunately, this is not the case because Rayleigh wave phase velocities are

influenced much less by changes in compression wave velocity than by changes in shear

wave velocity. In a uniform medium, Rayleigh wave velocity (VR) and shear-wave

velocity (β) are related by Equation 3.4 (Anderson, 2010):

β = VR/C (3.4)
32

The propagation velocity of individual frequency components is referred to as

phase velocity C. The variable C is a constant that changes slightly depending on

Poisson’s ratio of the material through which the seismic waves travel. Even in extreme

variations of Poisson’s ratio, C only ranges from 0.874 to 0.955 (Anderson, 2010). It is

suggested that if a value for C is assumed and the frequencies (with their respective

surface wave velocities) are recorded, then a shear-wave velocity profile can be

developed through analysis, and a velocity image of the subsurface can be generated

(Anderson, 2010).

Rayleigh wave velocities, as noted in Equation 3.2, are a function of both the

shear-wave velocity and the compressional wave velocity of the subsurface. A plot of the

frequency versus the phase velocity, known as a dispersion curve, visualizes these

relations. The shape of the dispersion curve is referred to as the dispersion characteristic

of the Rayleigh wave (Everett, 2013).

3.1.7. Rayleigh Waves in Vertically Heterogeneous Elastic Half-Space. Shear-

wave and compressional wave velocities vary with depth. Hence, different component

frequencies of Rayleigh waves (involving particle motion over different depth ranges)

exhibit different phase velocities (Bullen, 1963). The phase velocity of each component’s

frequency is a function of the variable body wave velocities over the vertical depth range

associated with that particular Rayleigh wavelength. More specifically, in the subsurface,

the Rayleigh wave phase velocity equation takes the following form:

VR (fj, CRj, β, α, ρ, h) = 0 (j = 1, 2, …., m) (3.5)

where

fj is the frequency in Hz,


33

VRj is the Rayleigh-wave phase velocity at frequency fj,

β = (β1, β2,….., βn) T is the shear-wave velocity vector,

βi is the shear-wave velocity of the i th layer,

α = (α1, α2, ….., αn) T is the compressional P-wave velocity vector,

αi is the P-wave velocity of the i th layer,

ρ = (ρ1, ρ2,…., ρn) T is the density vector,

ρi is the density of the i th layer,

h = (h1, h2,…., hn-1) T is the thickness vector,

hi is the thickness of the i th layer and,

n is the number of layers within the earth model,

The velocity of Rayleigh waves is comparable to the velocity of shear waves. In a

rock formation with a Poisson ratio of approximately 0.25, the velocity of the Rayleigh

wave is approximately 92 % of the velocity of the shear wave. In materials with rates

from 0.4 to 0.5, the percentage increases to 94 to 95.5 %, respectively (Steeples, 1998).

As a general rule, the velocities of Rayleigh waves are assumed to be approximately 90 to

92 percent of the respective shear-wave (Ivanov, Park, & Xia, 2009; Parasnis, 1997). The

shear-wave velocity can be estimated within a ten percent margin of error using these

assumptions (United States Corps of Engineers, 1995).

The shear-wave velocity of a material is very important when predicting the

impact that an earthquake’s seismic waves will have on a material as the waves pass

through it (Wood, 2009). Knowing the shear-wave velocity of material, one can

determine the shear modulus using the relationship between the shear-wave and soil or

rock as related by Equation 3.3:


34

μ=ρVs2 (3.6)

where

μ = shear modulus,

p = mass density, and

Vs = shear-wave velocity.

Rayleigh wave velocities, as noted in Equation 3.5, are a function of both the

shear-wave velocity and the compressional wave velocity of the subsurface. A plot of

frequency versus phase velocity, known as a dispersion curve, visualizes these relations.

The shape of the dispersion curve is referred to as the dispersion characteristic of the

Rayleigh wave (Evrett, 2013). In MASW, phase velocities are calculated for each

component frequency of the field-recorded Rayleigh waves. Then the resulting dispersion

curve (phase velocity vs. frequency) is inverted using a least squares approach, and a

vertical shear-wave velocity profile is generated (Miller et al., 2000; Nazarian et al.,

1983; Stokoe et al., 1994; Park et al., 1999a, 1999b, 2000; Xia et al., 1999). Typical

dispersion curve has multimodal character (multiple phase velocities existing for a certain

frequency). The mode with the lowest phase velocity (at each frequency) is defined as

fundamental mode and exists at all frequencies. Other higher modes, called the first

mode, second mode, etc. They have higher phase velocities and are only exist above a

cut-off frequency that depends on the mode (Evrett, 2013).


35

3.2. MULTICHANNEL ANALYSIS OF SURFACE WAVES DATA


ACQUISITION

Rayleigh wave velocities are, in general, shown to increase with depth (e.g. waves

with longer Rayleigh wavelengths and lower frequencies) and propagate faster than those

with shorter wavelengths. The relation between frequency (f) and Rayleigh wavelength is

called a phase velocity (c(f)). These unique characteristics result in a different Rayleigh

wavelength (λ(f) given as (Kramer, 1996) , as shown in Figure 3.5.

Figure 3.5. Rayleigh wave components with different Rayleigh wavelengths propagating
through a layered medium. Wave components with different frequencies reflect soil
properties at diverse depths.

The material properties of the topmost layer have an impact on the phase velocity

of wave component (1), whereas the phase velocities of wave components (2) and (3)

also depend on the properties of the deeper layers.

λ (f) = c(f)/f (3.7)


36

The variable C is a constant that changes slightly depending on Poisson’s ratio of

the material through which the seismic waves travel. Even in extreme variations of

Poisson’s ratio, C only ranges from 0.874 to 0.955 (Anderson, 2010). It is suggested that

if a value for C is assumed and the frequencies (with their respective surface wave

velocities) are recorded, then a shear-wave velocity profile can be developed through

analysis, and a velocity image of the subsurface can be generated (Anderson, 2010).

These unique characteristics result in a different Rayleigh wavelength λ.

Depending on the nature of the seismic source, MASW method can be

categorized as active or passive. The active MASW adopts the conventional seismic

refraction mode of surveying by using an active seismic source, such as a sledgehammer,

accelerated weight drop or vibroseis deployed in traverse along a linear array of receivers

(Park et al., 1999). The passive MASW method utilizes surface waves generated by

natural sources or cultural activities, highway traffic and construction equipment (Park et

al., 2007). Depending on the receiver configuration there are two ways to conduct a

passive MASW survey - the passive remote MASW (Park et al., 2004; Park et al., 2005)

employs a two-dimensional receiver array and the passive roadside MASW (Park and

Miller, 2008) uses linear one-dimensional receiver array. Data acquisition is more

tolerant in parameter selection than other seismic methods because of the easily achieved

high level of signal-to-noise ratio in seismic field records.

For the acquisition of active MASW data, low frequency (for example, 4.5 Hz),

vertically polarized geophones are lined on the test site surface at the appropriate equal

intervals. The number of geophones used is typically 24 or more with a constant inter-

geophone spacing that is optimized for site-specific geologic conditions (Donohue,


37

Forristal, & Donohue, 2013; Lin, Chang, & Chang, 2004). Each geophone is connected to

a separate recording channel (Park et al., 1997) and the whole array is connected to

engineering seismograph and field laptop equipped with software necessary for data

recording. A surface wave is generated with an impact load at one end of the lineup and

the geophones record the resulting wave motion as a function of time. The seismic source

may be a 20-lb sledgehammer, a mechanical impact device, a shotgun or explosives,

depending upon the depth of the investigation and the site-specific conditions. The arrival

of surface wave is detected along a set of geophones and recorded on a seismograph, with

the output of each geophone being displayed as a single trace.

Proper setup of data acquisition parameters is essential for the success of the

surface wave survey. The important field parameters are total length of the receiver

spread, source offset (the distance between the source and the nearest geophone), and

receiver spacing.

As the geophones only record vertical motion, it is important that they are placed

vertically in the subsurface. Each geophone is connected through a separate recording

channel to a data acquisition card and a computer equipped with the necessary data

acquisition software (Figure 3.6.).


38

Figure 3.6. The instrumentation used in the MASW tomography survey (Park et al.,
2002).

As illustrated in Figure 3.6, geophones are profiled up on the surface of the test

site. A wave is generated, and the wave propagation is recorded. The maximum depth of

the investigation (Zmax) varies with the site, the natural frequency (fe) of the geophones

and the type of seismic source that is used. The maximum investigation depth is

determined by the longest Rayleigh wavelength that is obtained during data acquisition

(λmax). The following is a commonly adopted empirical criterion (Park & Carnevale,

2010):

Zmax ≈ 0.5λmax (3.8)

3.2.1. Geophone Spread Length. The overall geophone spread length (D)

defines the longest Rayleigh waves wavelength that can be analyzed. It is also related to

the maximum investigation depth (Zmax), since the length of the receiver array is related

to the longest Rayleigh wavelength that is obtained during data acquisition. A common
39

criterion is that the longest Rayleigh wavelength (λmax) that can be analyzed is

approximately equal to the length of the receiver array (Park & Carnevale, 2010):

λmax ≈ D. (3.9)

Attempts to analyze longer Rayleigh wavelengths than those indicated by

Equation 3.9 can lead to less accurate results. Recent studies have shown that although

there is fluctuating inaccuracy, it will remain within 5% for the interval D ≤ λmax ≤2D

(Park & Carnevale, 2010), because of the uncertainties (noise) always included in the

measurement. A very long receiver spread should be avoided. Surface waves generated

by the most commonly used seismic sources (e.g. reasonably heavy sledgehammers) will

have attenuated noise level at the end of an excessively long receiver spread, making the

signal from the furthermost receivers too noisy to be usable (Park et al., 1999; Xia et al.,

2009). The maximum wavelength that can be accurately analyzed is approximately equal

to the half of the length of the overall geophones spread.

3.2.2. Source Offset. The appropriate choice of source offset eliminates the near-

field effect (the risk that geophones will pick up surface waves that were not fully

developed. Long source offset could generate energy for long-wavelength surface waves.

It can also result in lack of short-wavelength wave components due to the attenuation.

For long offsets, where the distance traveled along the refractor exceeds several Rayleigh

wavelengths, the head wave amplitude decreases by a factor close to the inverse square of

the offset. At smaller offsets, which is likely to be the case for many LVLs (low-velocity

layer or weathering surveys), the amplitude decrease is slightly less. In addition to the

variation of amplitude with the offset, significant attenuation occurs if the refractor is thin

(Press et al., 1954). This effect is normally associated with shingling, a phenomenon
40

characterized by a shift of energy to successively later cycles as the offset increases,

producing an echelon pattern, as shown in Figure 3.7.

Figure 3.7. Refraction arrivals showing the shingling degree pattern often associated with
layers.

The source offset needs to change in proportion to the intended maximum

investigation depth, with a conservative calculation being that the source offset is equal to

half of the maximum investigation depth (Park & Miller, 2005). The importance of

choice of optimum offset is that it controls the degree of contamination by a range of

near‐field effects. Surface wave cannot fully develop before it strikes the first geophone

in case of source positioned too close to the receiver spread. The risk of non-planar

surface waves being picked up by the geophones can be minimized by carefully choosing

the source offset (Park & Carnevale, 2010). The minimum source offset required to avoid

near-field effects depends on the longest Rayleigh wavelength that is being analyzed
41

(λmax). In most cases, plane-wave propagation of surface waves first occurs when the

source offset is greater than half of the longest Rayleigh wavelength (Park et al., 1999).

The source offset also needs to be sufficiently sized to extend to the primary depth

range of interest. This concept conforms to the common practice with Equation 3.10 and

3.11:

X1 ≥ 0.5D (3.10)

when λmax ≈ D (3.11)

where D is the length of the receiver array. However, studies have shown that this

criterion can be relaxed significantly for MASW surveys (Park et al., 1999, 2002; Park &

Shawver, 2009).

Studies have shown that this criterion can be relaxed significantly for MASW

surveys (Park et al., 1999, 2002; Park & Shawver, 2009). A long source offset, X1 ≥ D,

can potentially enhance energy for long- Rayleigh wavelength wave components, thus

increasing λmax for a receiver array of a given length (Park & Carnevale, 2010).

However, such a long source offset can result in a lack of short- Rayleigh wavelength

components, due to excessive attenuation (Park & Shawver, 2009). A long source offset,

X1≥ D, can potentially enhance energy for long- Rayleigh wavelength wave components,

thus increasing λmax for a receiver array of a given length (Park & Carnevale, 2010).

However, such a long source offset can result in a lack of short- Rayleigh wavelength

components, due to excessive attenuation (Park & Shawver, 2009). The suggested

minimum (X1, min) and maximum (X1, max) source offsets for use in practice are as

follows (Park, 2015):

X1, min = 0.2D and X1, max = D. (3.12)


42

The optimum offset (A), based on a layered ground model due to the longer

Rayleigh wavelength components of Rayleigh waves requires longer time or larger

offsets to develop into plane waves (Zhang et al., 2004).

𝜆𝑚𝑎𝑥 𝐶𝑅 𝑚𝑖𝑛
𝐴= (3.13)
4∆𝐶𝑅

where 𝜆𝑚𝑎𝑥 ,𝐶𝑅 𝑚𝑖𝑛 , and ∆𝐶𝑅 are the longest Rayleigh wavelength, the minimum

phase velocity of Rayleigh waves, and the difference between the maximum and

minimum phase velocities, respectively.

For example, the first shot was acquired with the nearest source offset of 15ft, the

longest Rayleigh wavelength of λmax = 500 ft/sec/25 Hz = 20ft, the minimum phase

velocity of CRmin =180ft/sec and the difference between the maximum and minimum

phase velocities of ΔCR = 500ft/s - 180ft/sec = 320ft/sec. Therefore, the suggested

optimum offset A would be 2.8ft.

3.2.3. Receiver Spacing. The receiver spacing (dx) is defines the shortest

wavelength that can be analyzed and the shallowest resolvable depth of MASW survey.

The minimum investigation depth λmin relates to the minimum distance between

geophones (O’Neill 2003). This assumes that the maximum depth of penetration is

approximately half of the maximum recorded Rayleigh wavelength (Park et al., 2009).

The receiver spacing is related to the shortest Rayleigh wavelength (λmin) that can be

included in a dispersion curve. In general, the receiver spacing should not be greater than

half of the shortest Rayleigh wavelength in order to avoid spatial aliasing (Xia et al.,

2009):

dx ≈ λmin ≈ Zmin
43

dx ≤ 0.5 λmin. (3.14)

Moreover, the receiver spacing acts as a guideline for determining the minimum

thickness (hmin) of the shallowest layer of the layered earth model used in the inversion

analysis (Park et al., 1999; Xia et al., 2009). In other words,

h1 ≥ hmin ≈ dx (3.15)

where h1 is the thickness of the topmost layer of the earth model.

The receiver spacing (dx) also determines the computational artifacts caused by

the aliasing impact. It sometimes generates some curved streaks in the dispersion image,

caused at the point of dispersion, where the Rayleigh wavelength becomes less than one-

half of (dx).

3.2.4. Orientation. The orientation of the profiles should best depict the

structures of the subsurface, which is as well important for the best possible

interpretation. While interpreting the seismic data, well log data and result from previous

geological surveys and studies of the region are often used in relation to the seismic data

(Sheriff & Geldart, 1995). Furthermore, the effect spacing needs to be small in all

orientation to avoid irregular surface. The subsurface coverage should be uniform with a

consistency between the contribution of different source offsets and azimuths (Bacon et

al., 2003).

Typically, there was a reasonable match between the two surveys from opposite

orientations. If a survey in one orientation produced a poor quality and a difficult-to-

extract dispersion curve, the survey in the opposite orientation generally produced a

dispersion curve of similar quality, which is largely presumed to be a result of subsurface


44

lateral heterogeneities. Processing the data surveyed in both orientations helped limit the

influence of disturbances.

To phrase it differently, both receiver spacing and source array methods can add

additional orientation effects. Subsequently, comprehensively considering the parameters

of the source array, the receiver spacing and the geometry in an acquisition system will

lay a foundation for a high quality and highly efficient seismic survey project.

3.2.5. Topographical Conditions. Topographical conditions are known to have

effect on the quality of the recorded surface wave data and therefore the quality of the

resulting dispersion curves. They can also affect the quality of the acquired surface wave

data and therefore affect the quality of the generated dispersion images (Zeng, Xia,

Miller, Tsoflias, & Wang, 2012). For optimum results, the receivers should be placed on

relatively flat terrain, suitable for Active MASW, as shown in Figure 3.8 (A) and 3.8 (B).

Figure 3.8. Topographical conditions are found to have an effect on the quality of the
recorded multichannel surface wave data. Receivers should be placed on relatively flat
terrain for optimum results (A & B). Surface reliefs greater than 0.1d and the recorded
data (C & D). Adopted from Park (2015).

Surface reliefs within the receiver spread greater than 0.1 d can have a significant

effect on the generation of surface waves (Park, 2015), as shown in Figure 3.8 (C), and
45

should therefore be avoided. The slope of the surface along the receiver spread can also

affect the quality of the surface wave records, as shown in Figure 3.8 (D).

However, topography can also interfere with surface wave propagation. Results

of numerical investigations presented by Zeng et al. (2012) indicated that the slope of the

topography along the survey profile (θ) should preferably be less than 10º. A steeper

topography (i.e., a slope angle θ > 10º) can lead to significant errors (greater than 4%) in

estimation of the Rayleigh wave dispersion characteristics (Zeng et al., 2012).

In theory, maximum investigation depth is defined by the longest surface wave

wavelength generated during the data acquisition. On practice, the maximum

investigation depth also depends on the survey site and type (strength) of seismic source.

A heavier source provides increased investigation depth. For example – the use of 20 lb.

sledgehammer typically results in 30 – 100ft. (10 – 30m) depth of investigation. Recent

studies reported that use of non-metallic plate placed in the impact point can generate

stronger energy at the lower frequency part of surface waves than a conventional metallic

plate (Cui, 2013). Also, the maximum investigation depth is determined by the longest

Rayleigh wavelength that is obtained during data acquisition.

Recording frequency of 1000 Hz (sampling interval of 1ms) is most commonly

used in active MASW surveys. The total recording time of 1 second is usually used for

the impulsive seismic source MASW surveys.

3.2.6. Resolution of MASW Data. Another important thing to consider is the

desired resolution of MASW data. Resolution is mostly a function of the receiver

spacing, which is directly related to the shortest wavelength and therefore determines the

shallowest resolvable depth of investigation. The total length of the receiver spread is
46

directly related to the longest wavelength that can be analyzed, which in turn defines the

maximum depth of investigation. The source offset distance controls the possible degree

of contamination by the near-field effects (surface waves are formed through interference

of body waves generated from reflections and refractions, so they require to propagate a

certain minimum distance from the source to fully develop, which may result in both

underestimation or overestimation of Rayleigh wave phase velocity due to the body

waves contamination near to the source). Vertical stacking with multiple impacts can

suppress ambient noise significantly, producing a high signal to noise ratio. This is good

in surveys in urban areas. Three to five vertical stacks are recommended during survey.

3.3. MULTICHANNEL ANALYSIS OF SURFACE WAVES DATA PROCESSING

Various component frequencies of Rayleigh waves involve particle motion at

various depths. For example - the lower frequencies involve particle motion at greater

depths. So the different component frequencies of Rayleigh waves exhibit different phase

velocities. The whole MASW technique is based on the relationship between the

Rayleigh wave phase velocities and the depth-range of associated particle motion.

The quality of the acquired surface wave records can be evaluated in terms of the

resolution of the phase velocity spectrum, i.e., the sharpness of the amplitude peaks

observed at each frequency, the extractable frequency range and the continuity of the

fundamental mode high-amplitude band.

The quality of the MASW data can also be affected by natural geologic conditions

that may not produce well-defined dispersion curves and cannot be used to calculate

reliable shear-wave velocity models.


47

After data acquisition, field records are analyzed using SurfSeis software

developed by Kansas Geological Survey (KGS). In order to convert seismic record to an

estimate of shear wave velocity, three crucial steps must be performed: generation of

dispersion image (overtone - a graphic representation of intensity in phase velocity and

frequency space) and extraction of the fundamental-mode dispersion curve from it. The

software calculates phase velocities for each component frequency of recorded MASW

data to generate the resultant dispersion image to extract the dispersion curve. Then, the

curve is inverted using a least–squares approach to generate a vertical shear-wave

velocity profile. Basically, software performs wavefield transformation from offset-to-

time domain into phase velocity-to-frequency domain (multichannel record, time-space

domain, to dispersion image, frequency-phase velocity domain). Then MASW data are

being inverted to generate 1-D (depth) shear wave velocity profile. Figure 3.9 displays

the raw seismic record, dispersion curve extracted from the overtone image, and the

resulting shear wave velocity profile.

3.3.1. Dispersion Analysis. The intent of dispersion analysis is to estimate one or

more dispersion curves that are in turn passed into the next step of inversion process. The

following are the most influential factors affecting the dispersion analysis:

1. Frequency range for depth of investigation,

2. Approximate phase velocity range,

3. Easy characterization of higher modes, and

4. Identification and reduction of noise events.

The influence of these factors on the analysis is highly dependent on the data

quality (signal-to-noise ratio, S/N). A “good” quality data set suggests that surface-wave
48

is the most prominent seismic event (with highest S/N), whereas a “bad” quality data set

is usually contaminated by noise. In the dispersion curve analysis, the fundamental mode

surface waves are the signal, and everything else is noise. Noise includes all higher-mode

surface waves as well as all body wave events.

Figure 3.9. (A) Acquisition seismic time series data; (B) Dispersion curves extraction
frequency and phase velocity; and (C) 1-D shear-wave velocity profiles Vs. 1-D depth
curve (http://www.kgs.ku.edu/software/surfseis/gifs/masw).

Software also performs evaluation of an approximate phase velocity range for the

surface waves. It generally ranges from as low as 650ft/sec (or 200m/sec) to as high as

2,500ft/sec (760m/sec) depending on the material type. This information is used by the

program to initiate the analysis by searching within this range for a phase velocity
49

corresponding to a certain surface wave frequency with the greatest coherence throughout

the entire range of offset and the highest signal-to-noise ratio which is indicative of a

high confidence in the acquired phase velocity (Figure 3.9). The SurfSeis software

calculates phase velocities within the specified frequency range. This calculation can be

run multiple times using different values and varieties of parameters, examining the

output curves until an optimum solution is identified. In general, the curve with the

highest signal-to-noise ratio (S/N) represents the best option. The quality of a dispersion

curve is judged according to two criteria: the signal-to-noise ratio (S/N) and the general

dispersion curve direction.

The next processing step is an extraction of a fundamental-mode dispersion curve

from the dispersion image. This extracted curve is called a "measured" dispersion curve

that is an input data to the next data analysis step (inversion). The quality of the acquired

surface wave records can be evaluated in terms of the resolution of the phase velocity

spectrum, i.e., the sharpness of the amplitude peaks observed at each frequency, the

extractable frequency range and the continuity of the fundamental mode high-amplitude

band. The quality of a dispersion curve is judged according to two criteria: the signal-to-

noise ratio (S/N) and the general dispersion curve direction. The quality of the ‘match’

between the two curves is evaluated on the root‐mean‐square error (Xia et al., 1999). The

inversion of a dispersion curve using SurfSeis Software is particularly straightforward, as

it is a fully automated process that removes any human error incurred during the

calculations.
50

3.4. MULTICHANNEL ANALYSIS OF SURFACE WAVES DATA


INTERPRETATION

Data interpretation of MASW involves the analysis of the variations of the shear

wave velocities with depth with a goal to transform the output velocity model into a

geologic model of the subsurface.

Interpretation of the subsidence features in karst terrain environment is difficult

because of the inherent complexity of the subsurface and the resolution limits of MASW

technique. Shear wave velocity models can clearly show the low-velocity zones, which

can be interpreted as sinkholes, voids, fractured bedrock, or naturally in-filled collapsed

features.

In Missouri for example, the shear wave velocity value used to determine the

depth to top of bedrock is mostly 1000ft/sec. but can vary. Again, highly stiff earth

materials have relatively high shear wave velocities compared to fractured or weak earth

materials (Table 3.1).

Reliability of output shear-wave velocity data decreases as lateral and vertical

heterogeneity of soil/rock increases. Therefore, site conditions should be considered; the

immoderate dip of the subsurface layer along the survey line (more than about 10

percent), disadvantageous topography, or known high lateral variability in soil or rock

properties may be reasons to reject field data as inconvenient for interpretation in terms

of simple vertical variation of subsurface.


51

Table 3.1. Shear wave velocity (Vs) of some earth materials (National earthquake hazards
reduction program).

3.5. ELECTRICAL RESISTIVITY TOMOGRAPHY (ERT)

3.5.1. Overview. Electrical Resistivity Tomography (ERT) is a nondestructive

geophysical technique based on the electrical resistivity method. ERT has been used for

decades as an effective environmental and geotechnical tool. In particular, ERT

methodology is widely used for determining the maximum depth of rock, acquiring

information on the elevation of soil, the top of rock, etc. This method is especially

preferred for vision characterization in karst terrain (Zhou, 1999).

The purpose of an electrical resistivity survey is to determine the subsurface

resistivity distribution in karst characteristics, such as caves, that may or may not be

easily recognizable on the subsurface. Areas where the top of rock is limestone or

gypsum, like it is in Missouri have a high probability of karst development. Karst areas

commonly lack surface water and have numerous streambeds that are dry except during
52

periods of high runoff. When an electrical resistivity tomography survey is conducted in

karst terrain, current flow is generally assumed to be electrolytic rather than electronic.

Surface electrical resistivity surveying is based on the principle that the partition

of electrical potential on the subsurface around a current-carrying electrode depends on

the soil’s materials and rock to demonstrate the variations in their electrical resistivity

because of the variations in their mineral content, permeability, fluid saturation, porosity,

etc. After that, areas of the subsurface undergoing dissolution can be differentiated from

the top of rock by measuring the resistivity of the subsurface in good resolution.

3.5.2. Ohm’s Law and Resistivity. The fundamental principle behind the

combination and interpretation of electrical resistivity measurements was created in the

electrical physical theory of Ohm’s law. Ohm’s law, Equation 3.16, states that the

product of the electrical current, I, and the resistance of the wire, R, through which the

current passes is equivalent to the potential difference, V, across the conductor:

V = IR (3.16)

According to Gibson and George (2003), this relationship is best represented by

envisioning a current passing through a thin wire. The expounded application of Ohm’s

law has made this relationship a capstone concept in the study of electrical theory. Units

for electrical potential, current and resistance are volts, amperes and ohms, respectively.

As indicated, the conductor element can be tangibly described as a wire element.

The resistance of the wire is related to both the geometric shape and material

attributes of the wire. The geometry of the wire is typically cylindrical, possessing a

length and cross-sectional area, and is made of a conductive material. The total resistance
53

of the wire element, R, is the product of the material resistivity, ρ, and the ratio of the

wire length and cross-sectional area:

R = ρ (L/A). (3.17)

Considering the physical relationship between the geometry of the conductor and

the material property, Equation 3.16 can be manipulated to from Equation 3.17 to

determine the material resistivity of the conductor element.

This form states that the units for resistivity depend on the volume of the space

through which the current travels. Typical units for resistivity, ρ, include ohm-meters and

ohm-centimeters (Gibson & George, 2003).

In a similar situation, the measurement of the potential difference can be related to

the dissipation of the electrical current within an infinite, homogenous half-space. In this

instance, the application of an electrical current travels in radial fashion out from the

point of origin. During the current application, the resistance at any location away from

the source point within the homogeneous mass can be found by determining the radius

from the point of origin and the surface area of the respective hemispherical equipotential

surface. Relating this model to the original wire example, Equation 3.17 can be rewritten

using the radius, r, as the distance for which the current travels and the surface area of the

resulting equipotential surface, 2πr2.

R = ρ(r/2πr2) = (ρ/2πr) (3.18)

Equation 5.4 describes the system resistance at any point away from the source

point within the homogeneous mass. Using the resistance term from the mentioned

homogeneous earth model, Equation 3.19 relates the resistance of the earthen model to

Ohm’s law:
54

V = IR = I (ρ/2πr) (3.19)

where

U = potential, in V,

ρ = resistivity of the medium, and

r = distance from the electrode.

Likewise, the potential difference between any two points within the

homogeneous mass would be the difference between the two equipotential surfaces, as

expressed in Equation 5.5 (Gibson & George, 2003):

𝜌𝐼 𝜌𝐼 𝜌𝐼 1 1
𝑈 = 2𝜋 𝑟𝐴 − 2𝜋𝑟 𝐵 = [𝑟 𝐴 − 𝑟 𝐵 ] (3.20)
2𝜋

where

rA and rB = distances from the point to electrodes A and B

Equation 3.21 relates the applied current, I, and measured potential difference, V,

to a constant value that accounts for spatial considerations, or the way in which the

reading was acquired. This model, a concept of equipotential surfaces, and means of

measuring potential differences between various surfaces, is fundamental in the

interpretation of collected field data (Gibson & George, 2003).

ρ = (2πU/I) [1/ [(1/ra)-(1/rb)] (3.21)

Figure 3.10 explains the electric field around the two electrodes regarding

equipotential and current lines. The equipotential represents the imagery projectiles or

bowls surrounding the current electrodes where the electrical potential is equal on each

one. The current lines represent a sampling of the infinite paths followed by the current,

which are defined by the condition that they must be everywhere and normal to the

equipotential surfaces.
55

Figure 3.10. Equipotential and current lines for a pair of current electrodes, A and B, in a
homogeneous half-space.

In a homogenous condition, the measured resistivity will be equivalent to the real

value of resistivity at a given location within the media. However, the occurrence of a

homogenous condition is rare in practice, if not nonexistent. A collected reading is

considered an apparent resistivity measurement in order to account for the inherent

heterogeneity of the subsurface. Visual resistivity is the resistivity of a theoretical,

homogeneous half-space that complements the measured current and potential difference

for a particular measurement scheme (United States Corps of Engineers, 2001). Mostly,

the obvious resistivity value is an average reading of the energized soil mass engaged

during the measurement. Numerically, the obvious resistivity can be expressed by

Equation 3.22:
56

𝜌𝐼 1 1 1 1
𝑉 = 𝑈𝑀 − 𝑈𝑁 = [ 𝐴𝑀 − 𝐵𝑀 + 𝐵𝑁 − 𝐴𝑁] (3.22)
2𝜋

where

UM and UN = potentials at M and N, and

AM = distance between electrodes A and M, etc.

These distances are always the actual distances between the respective electrodes,

whether or not they lie on a line. The quantity of the brackets, S is a function only of the

various electrode spacing. The quantity is denoted by 1/K, which allows the equation to

be rewritten as:
𝜌𝐼 1
𝑉= (3.23)
2𝜋 𝐾

where

K = the array geometric factor.

Equation 3.24 can be solved for ρ to obtain


𝑉
𝜌 = 2𝜋𝐾 . (3.24)
1

The resistivity of the medium can be established from the measured values of V, I

and K, the geometric factor. The variable K is a function only of the geometry of the

electrode arrangement.

When these measurements are made over real heterogeneous earth, as

distinguished from the fictional homogeneous half-space, the symbol ρ is replaced by ρa

for apparent resistivity. The resistivity surveying problem is reduced to its essence. The

use of apparent resistivity values from field monitoring at various locations, and with

various electrode configurations, is designed to estimate the actual resistivity of several

earth materials present at a site and to locate their confines spatially below the surface of

the site.
57

An electrode array with stationary spacing is used to investigate the lateral

changes in apparent resistivity, reflecting the lateral geologic variability or localized

anomalous features. The size of the electrode array is varied as we investigate how

changes in resistivity affect depth.

The resistivity is clearly affected by a material at increasing greater depths (hence

the larger volume) as the electrode spacing is increased. Because of this impact, a plot of

apparent resistivity against electrode spacing can be used to signal vertical variations in

resistivity.

The geometric coefficient, K, varies with array types. The spacing and layout of

the current and potential electrodes impact the induced equipotential fields created within

the earthen mass. The geometric factor for a general four probe system can be derived

from Equation 3.21(Gibson & George, 2003).

3.5.3. Relationship between Geology and Resistivity. Electric current flows in

subsurface materials at shallow depths through two main methods. They are electronic

conduction and electrolytic conduction. In electronic conduction, the current flow is via

free electrons, such as in metals. In electrolytic conduction, the current flow is via the

motion of ions in groundwater. In engineering and environmental surveys, electrolytic

conduction is the more common mechanism. However, electronic conduction is

important when conductive minerals are present, such as metal sulfides and graphite in

mineral surveys.

The resistivity of common rocks, chemicals, and soil materials (Keller &

Frischknecht, 1966: Daniels & Alberty, 1966: Telford et al., 1990) is shown in Figure

3.11. Igneous and metamorphic rocks typically have elevated resistivity values. The
58

resistivity of these rocks is greatly dependent on the degree of fracturing and the

percentage of the fractures that are filled with groundwater. Thus, a given rock type can

have a large range of resistivity, from about 1000 to 10 million Ω⋅m, depending on

whether it is wet or dry. This characteristic is useful in the detection of fracture zones and

other weathering features, such as in engineering and groundwater surveys.

Sedimentary rocks, which are usually more porous and have higher water content,

normally have lower resistivity values compared to igneous and metamorphic rocks. The

resistivity values range from 10 to about 10000 Ω⋅m, with most values below 1000 Ω⋅m.

The resistivity values are dependent on the porosity of the rocks and the salinity of the

contained water.

Unconsolidated sediments generally have even lower resistivity values than

sedimentary rocks, with values ranging from about 10 to less than 1000 Ω⋅m. The

resistivity value is dependent on the porosity (assuming all of the pores are saturated) as

well as the clay content. Clay soil normally has a lower resistivity value than sandy soil.

However, one should note the overlap in the resistivity values of the different classes of

rocks and soils should be noted. This is because the resistivity of a particular rock or soil

sample depends on a number of factors, including the porosity, the grade of water

saturation, and the concentration of dissolved salts. The resistivity of groundwater varies

from 10 to 100 Ω⋅m, depending on the concentration of the dissolved salts (Loke el al.,

2011).
59

Figure 3.11. The resistivity of rocks, soils, and minerals (Keller & Frischknecht, 1966;
Daniels & Alberty, 1966; Telford et al., 1990).

3.5.4. Electrical Resistivity Array Configuration. In theory, soil resistivity

could be measured by using a single current source and receiver element. In practice, this

is not feasible due to the contact resistance between the earth and the electrode pair. To

overcome this phenomenon, four electrodes are used for measurement: two electrodes

provide current to the subsurface and two electrodes measure the potential difference

between the earth materials (Milson, 1996). Current electrodes are identified as C1 and

C2 (or A and B), and potential electrodes are identified as P1 and P2 (or M and N) (Loke,

2000), as shown in Figure 3.12.

The types of electrode arrays that are most commonly used are Wenner (Figure

3.12a, Schlumberger (Figure 3.12b), pole-dipole (Figure 3.12c) and Dipole-Dipole


60

(Figure 3.12d). Other electrode configurations are either used experimentally, on non-

geotechnical problems, or are no longer popular. Some of these include the Lee, half-

Schlumberger, Polar Dipole, and gradient arrays.

In any case, the geometric factor for any four-electrode system can be found from

Equation 3.22 and can be advanced for more complicated systems by using the rule

illustrated in Equation 3.20. It can also be seen from Equation 3.25 that the current and

potential electrodes can be interchanged without affecting the results; this property is

called reciprocity (Milson, 1996). For the purpose of this research, the discussion will

concentrate on the Dipole-Dipole array, which was used during these studies.

Unlike to the Wenner and Schlumberger arrays, the configuration of the Dipole-

Dipole array does not place the potential electrode pair inside the current electrode pair.

Current and potential electrode pairs have common interior spacing and are separated by

a distance ten times the internal spacing of the electrode pair. The Dipole-Dipole array is

ordinarily used for performing tomography survey due to the array’s ability to resolve

lateral variations. In comparison to the Wenner and Schlumberger arrays, the Dipole-

Dipole array has a weak signal and is more susceptible to the effects of ambient or

cultural, noise.

If the division between both pairs of electrodes is the same, a, and the division

between the centers of the Dipoles is restricted to a (n+1), the resistivity is given by
𝑉
ρa = πaπ(n+1) (n+2) 1 . (3.25)

3.5.5. Wenner Array. The Wenner array is described by the equal spacing

between all four electrodes. The two current electrodes, C1 and C2, are placed on the

outside of the array during which the potential electrodes, P1 and P2, reside inside of the
61

array. The potential measurements are possessed at the mid-span of the potential

electrodes at a depth of approximately 0.5 to 1.0 times the electrode spacing (also known

as the spacing). Different depth measurements are made by varying the period spacing of

the array, as shown in Figure 3.12a.

Figure 3.12. These are some commonly used electrode arrays and their geometric factors.
Note that for the multiple gradient arrays, the total array length is‘(s + 2) an’, the distance
between the center of the potential Dipole pair is P1-P2 and the center of the current pair
C1-C2 is given by ‘ma.’ 'British Geological Survey (c) NERC 2013. K = Geometric
Factor.

In an exemplified homogeneous earth model, the sensitivity of the Wenner array

provides a pattern with strong horizontal layering immediately below the potential

electrode pair. Because the focus of vertical electrical sounding (VES) is to differentiate

between horizontal layering beneath a common point, the Wenner array is a practical
62

array for this application. The high signal of the Wenner array also makes the array

suitable for use in noisier environments (Loke, 2000).

3.5.6. Schlumberger Array. The Schlumberger array is arranged with two

current electrodes on the outside of the array, set apart by a distance of at least five times

the space between the two interior potential electrodes. The potential difference

measurement is believed to show incorrect at the mid-span of the potential internal

electrodes, at a depth of approximately one-half of the length between the current

external electrodes. Similar to the Wenner array, the Schlumberger array provides a high

signal directly below the potential electrode pair, as shown in Figure 3.12b.

The Schlumberger array is preferred for VES applications due to the strong

horizontal resolution and ease of setup in the field. As compared to the Wenner array

where all four electrodes must be repositioned after each test,

The Schlumberger array is preferred for VES applications due to its strong

horizontal resolution and ease of setup in the field. As compared to the Wenner array,

where all four electrodes must be repositioned after each test, the Schlumberger array

only requires that the two exterior current electrodes be moved to acquire a new

measurement. The potential interior electrodes are moved only as the current electrodes

are spaced beyond the practical limits of the survey. That movement occurs when the

ratio between the potential electrode spacing and the space between the exterior current

electrode and positional electrodes in the mid-span is greater than 0.4 (United States

Corps of Engineers, 1995).

3.5.7. Dipole-Dipole Array. The Dipole-Dipole array is logistically the most

appropriate in the field, particularly for large spacing and for 2-D imaging. The
63

convention for the Dipole-Dipole array shown in Figure 3.12d is that the current and

voltage electrode spacing are the same, a, and the spacing against them is an integer

multiple of a. The Dipole-Dipole array is usually used for performing tomography

surveying due to the array’s ability to resolve lateral variations. For a comparison against

the Wenner and Schlumberger arrays, the Dipole-Dipole array has a weaker signal and is

very susceptible to the effects of ambient, or cultural, noise. The apparent resistivity

reading recorded using the Dipole-Dipole array represents a condition present at the mid-

span of the array length that occurs at a depth amounting to one-half the product of the

Dipole electrode spacing, a, and one plus the separation factor (n+1).

3.6. ELECTRICAL RESISTIVITY TOMOGRAPHY DATA ACQUISITION

ERT data are typically acquired with the use of resistivity meter. Resistivity meter

is connected to the electrode cables, and the electrode cables are attached to metal stakes

pounded to the ground using rubber-band. In some instances, it is essential to water the

ground in proximity of the metal electrode to ensure a good contact between the metal

stake and ground. Newer resistivity meters are equipped with multiple channels, which

allows multiple electrodes to be engaged and measurements to be taken through each

channel. For instance, the SuperSting R8 resistivity meter, produced by Advanced

Geosciences, Inc., is equipped with eight channels. Subsequently, the system engages

nine electrodes to collect eight different potential difference measurements for each

current injection (Advanced Geosciences, Inc., 2006).

Planning and preparation are a significant part of ERT survey. Desired depth of

investigation, acquisition time, data resolution must be considered prior to each ERT
64

survey. The sequencing information considers the array style and information pertaining

to the electrode locations (or electrode address) during each measuring sequence. There

are no theoretical limits to the depth of penetration. Therefore, as the electrode spacing

increases, the signal strength decreases. At a certain electrode separation distance, the

signal strength is too low to provide reliable measurements of the potential difference.

Practical limits should be instilled that consider the signal strength of the particular array

type and equipment ability. Advanced Geosciences, Inc., (2008) suggested that when

considering depth of penetration for tomography applications, practitioners can generally

assume that the depth of penetration is approximately 15 to 25 % of the length of the

electrode spread. The survey resolution is also related to electrode spacing.

Coskun (2012) recommends the commonly used Dipole-Dipole array because it

offers better lateral and vertical resolutions than the other arrays. It is also worth noting

that the data acquisition time for a Dipole-Dipole array using 72 electrodes is typically

45minutes, and around 2 hours and 3 hours for 84 electrodes, and 168 electrodes,

respectively after survey set-up. Loke (1999) shares a similar view as Coskun, adding

that the Dipole-Dipole array is more suitable for investigating karst terrain because of its

ability to detect sharp changes in bedrock topographies.

Current practices propose that the electrode spacing should not be greater than

twice the size of the object or feature to be imaged. The design of the survey (i.e., survey

run length, electrode spacing, and array type) directly impacts the depth of penetration

and resolution (Advanced Geosciences Inc., 2008). It is not always possible or practical

to image a survey line or area in one deployment of electrodes. However, this can vary

depending on the preference between the resolution and imaged depth. Typically,
65

increasing the electrode spacing will increase the imaged depth and reduce the resolution

of the section and vice versa. However, it is often recommended that if the desire is to

increase the imaged depth, more electrodes should be added to the section instead of

increasing the electrode spacing. To continue a survey after completion of the initial data

collection, electrodes may be configured to collect additional data along a common

survey line or area, using roll-along survey techniques (Loke 2000), as shown in Figure

3.13.

Figure 3.13. SuperSting R8/IP resistivity (left). Switch Box (AGI) connecting passive
cables (middle). ERT field setup (right).
66

3.7. ELECTRICAL RESISTIVITY TOMOGRAPHY DATA PROCESSING

After data acquisition, the raw data are typically transferred to the laptop for

processing. During data processing inversion software transforms the apparent resistivity

values measured from the field to true resistivity by applying backward and forward

modeling in a process called inversion. There are a variety of processing software

available these days, including RES2DINV (Geotomo), Earth Imager (AGI), ZondRes2D

(Zond), etc. In this study, Res2DInv is used for ERT data processing. The RES2DINV

program uses the cell-based method in which the model parameters are the resistivity

values of the model cells, and the data provides the measured apparent resistivity values

(Loke, 2011). The mathematical link between the model parameters and the model

response for the 2-D and 3-D resistivity models is provided by either the finite-difference

(Dey & Morrison, 1979) or finite-element methods (Silvester & Ferrari, 1990).

The ERT data processing involves two main steps. The first step is to the

inspection of the resistivity data sets for the presence of any points that have anomalously

high or low apparent resistivity values, called “bad data points”, and subsequently

removing them when necessary (Figure 3.14b). Bad data points can be a result of several

factors, such as the failure or malfunction of equipment during the survey and very poor

electrode subsurface contact due to dry soil or shorting across the cable caused by very

wet subsurface conditions.

The second step is to run an inversion of the data. The inversion involves some

iterative calculations and the generation of a 2-D resistivity image of the subsurface to

represent the actual resistivity of the sections. The final result of the inversion is a 2-D

resistivity image of the subsurface showing the distribution of resistivity across a profile.
67

Additionally, pseudo section can be displayed and shows distribution of apparent

resistivity values (Figure 3.14a).

The root-mean squared (RMS) error is used to measure the difference between

calculated and measured apparent resistivity values. In practice, the lower the error, the

better the data quality is. However, the model with the low RMS error can sometimes

show unrealistic variations in the model resistivity values, and might not be best model

from geological perspective. A careful approach is to choose the model at the iteration

after which the RMS error does not change significantly. According to Loke (1999), an

RMS error of 5% is recommended for a geologic model of good quality.

Figure 3.14. An example of a field data set with a few bad data points. The apparent
resistivity data in (a) pseudosection form and in (b) profile form. (Loke, 1999).
68

3.8. ELECTRICAL RESISTIVITY TOMOGRAPHY DATA INTERPRETATION

The interpretation is based on the inverse model generated form the inversion

software, RES2DINV. The chart showing colors with various ranges of resistivity is used

to estimate the resistivity of the imaged subsurface earth materials.

Factors such as porosity, conductivity, saturation, salinity, clay content, lithology

and temperature can affect the ability of different materials to conduct electrical current.

Accordingly, materials of the same mineral content may exhibit different resistivity

values. For example, both the top-of-rock limestone and air-filled voids typically are

characterized by high resistivity values ranging from 50 ohm-m to 107 ohm-m., which

can be seen in Table 3.2. When an air-filled void is entirely embedded in limestone, it

usually cannot be easily detected on the resistivity data because of low resistivity

contrast.

Furthermore, dry soil usually has a much higher resistivity than saturated soil. The

same situation appears with weathered and un-weathered rock. Weathered rock is usually

more porous and fractured, and it becomes more saturated with groundwater; as a result,

weathered rock has a lower resistivity than un-weathered/the top of rock.

Table 3.2. Resistivity of common earth materials (Robinson, 1988).


69

According to previous studies (Torgashov, 2012) conducted in southwest

Missouri, typical resistivity values for the subsurface materials are characterized as

follows:

 Moist clays in southwest Missouri are normally characterized by low resistivity

values, usually less than 100 ohm-m, and may vary due to different degrees of

saturation, porosity, and layer thicknesses.

 Moist soils and intensively fractured rocks intermixed with clay typically have

resistivity values between 100 and 400 ohm-m. Such variation is explained by

different porosity, saturation, clay content and layer thicknesses.

 Relatively top-of-limestone with minimal clay content is characterized by higher

resistivity values, typically more than 400 ohm-m. Resistivity values of the top of

limestone may vary due to varying layer thickness, moisture content, porosity,

saturation, and impurities.

 Air-filled cavities usually show very high resistivity values, usually more than

10,000 ohm-m, but again these are variable depending on the conductivity of the

surrounding strata and depth/size/shape of a void.

 Zones of electrical resistivity contrast are where relatively the top of rock is

surrounded by moist, loose materials (such as clay) or where air-filled voids are

embedded in relatively the top of limestone. These zones can be successfully

detected by electrical resistivity tools.


70

4. RESEARCH EXPERIMENTS

4.1. STUDY SITE

The study site is located immediately within the southwestern limits of southwest

Missouri of the Ozarks physiographic region which is characterized by undulating to

rolling plains. Elevations of the region range between ~900 to ~1,500 ft. (asl). Southwest

Missouri lies on the western side of the Ozark Uplift where the rock layers dip gently

towards the west with minor faulting and folding, regionally. It is underlain by thick

Mississippian-age limestone with high porosity and susceptible to dissolution. Uneven

dissolution of this formation has resulted in a highly irregular bedrock-overburden

interface. Therefore, karst density in the area is among Missouri’s highest (Figures 4.1

and 4.2). The details of the studies are described in the next two sections.

This research emphasizes the need to apply efficient means for investigating

hazards to urban development such as an unstable soil foundation for structures, flood

hazards, groundwater contamination, and public safety in regards to potential catastrophic

collapse.
71

Figure 4.1. Southwest Missouri map where the study was conducted.
(http://www.nationsonline.org/oneworld/map/USA/missouri_map.htm).

Figure 4.2. Map showing the location of the study area Zone A and Zone B (Courtesy of
Google Earth).
72

4.2. MULTICHANNEL ANALYSIS OF SURFACE WAVES (MASW)

4.2.1. Data Acquisition. MASW data were acquired at 34 locations parallel and

perpendicular to the ERT traverses oriented west to east. Total number of 272 MASW

profiles in both west to east and north to south directions were acquired.

Data acquisition starts by the recording software launch and arming of the trigger

geophone. Once the MASW array initialized, operator gives the command to release the

seismic source (20 lb. sledgehammer) which notifies the seismograph to start data

recording. From three to five seismic records were collected at each traverse for a

purpose of staking to minimize the environment noise.

The main goal of the study was to develop appropriate acquisition parameters for

MASW surveys conducted in karst terrain topography in order to improve the data

resolution and maximize the depth of investigation. Figure 4.3 shows employed MASW

equipment: Seistronix RAS-24 seismograph and 24 vertically polarized 4.5 Hz

geophones. Source offsets of 10 ft. (3 m) and 30 ft. (9 m) were used for data acquisition.

Receiver spacings of 2.5 ft. (0.76 m) and 5 ft. (1.5 m) were used in attempt to increase the

vertical resolution of MASW data and to minimize lateral smearing. These parameters

were chosen to achieve the investigation depth of approximately 30 ft. (9m).

For each MASW traverse, eight measurements per site were obtained (four

records with seismic arrays oriented in west-east direction and four records with seismic

arrays oriented in north-south direction (parallel and perpendicular to the direction of

corresponding ERT traverses). Table 4.1 and Figure 4.4 shows parameters used to collect

MASW data in the study.


73

Figure 4.3. Active MASW data acquisition.

Table 4.1. Survey parameters used during MASW surveys in this study.
74

Figure 4.4. Schematic diagram of MASW data acquisition setup.

4.2.2. Data Processing and Interpretation. All seismic records acquired in this

research were processed using SurfSeis4 software, developed by the Kansas Geological

Survey.

Figures 4.5 to 4.7 display examples of the multichannel seismic records with

overlay of preliminary data quality evaluation provided by the processing software. This

step of data processing is to check overall data quality. This step is required to get a

primary idea about the overall data quality, surface wave velocity range, dominant

frequency, and the relative probability of higher modes contamination. “Excellent” and

“good” quality data sets suggest that the surface waves are the most prominent seismic

events (with highest signal-to-noise ratio) in the particular seismic record. “Fair” and

“poor” quality data sets are typically contaminated by noise which can be a result of the

irregularity of the bedrock surface, presence of voids, proximity of the array to the
75

construction site, receivers malfunction, etc. These seismic records can be not suitable for

processing due to a lack of any recognizable surface wave signal.

Figure 4.5. The multichannel seismic records with overlay of preliminary data quality
evaluation provided by the processing software (SurfSeis4).

Figure 4.6. The multichannel seismic records with overlay of preliminary data quality
evaluation provided by the processing software (SurfSeis4).
76

Figure 4.7 The multichannel seismic records with overlay of preliminary data quality
evaluation provided by the processing software (SurfSeis4).

Following the visual evaluation, initial processing was initiated and included

staking multiple shots as well as applying muting. The purpose of muting application is

to remove specific seismic energy which arrives in specific time frame as a result of its

velocity of propagation characteristics (Miller, Ivanov, et al., 2001). After proper muting,

the quality and observation range of the desired mode (i.e., fundamental mode) improve

significantly. The downside of muting is that it introduces a high velocity gradient feature

at the low frequency end of the dispersion curve, which is not present before muting. This

artifact can be easily handled by using the dispersion curve from the non-muted data that

covers the same frequency range, so the artificially high velocity values can be ignored

and are thus considered harmless (Miller, Ivanov, et al., 2001).

To show the effect of muting, the muting was applied to the seismic record

collected with the source offset of 10ft. (3.04 m) and the receiver spacing of 2.5ft. (0.76)

as it shown in Figures 4.8 and 4.9 to compare their dispersion curves before and after

application of muting. It can be seen from Figure 4.8A that seismic energy associated
77

with surface waves dominates in the seismic record. Figure 4.8B displays the dispersion

image with up-warping artifact appeared at high frequencies range that can be caused by

contamination from higher modes. The fundamental mode dominates a wide frequency

range between 38 and 61 Hz (with corresponding velocities between 1,060ft/sec and

1,735ft/sec) on the overtone image. However, the resulting 1-D shear wave velocity

profile can be distorted if the extracted dispersion curve will be improperly picked from

the ambiguous overtone image.

Figure 4.9B displays the dispersion image generated from the muted seismic

record. The presence of higher mode surface waves in the original seismic record is

almost completely eliminated, and the seismic energy associated with higher modes

disappeared. The fundamental mode can be observed within a frequency range between

36 and 63 Hz (with corresponding velocities between 1,040ft/sec to 1,720ft/sec).

Their attempt to observe seismic energy associated with fundamental mode in the

range between 36 and 67 Hz by boosting the amplitude of the phase-velocity – frequency

spectrum did not bring any results. Thus, the seismic energy associated with higher

modes is significantly higher, and corresponding dispersion images inhibit the ambiguous

character of dispersion curve extraction of the fundamental mode trend below 40 Hz.
78

Figure 4.8. (A) The multichannel seismic record collected with the source offset of 10ft.
(3.04 m) and receiver spacing of 2.5ft. (0.76m); (B) Corresponding overtone image with
the fundamental and higher modes identified.
79

Figure 4.9. (A) The multichannel seismic record collected with offset of 10ft. (3.04 m)
and receiver spacing of 2.5ft. (0.76 m) with applied muting. (B) Corresponding overtone
image with the fundamental and higher modes identified.
80

Following the generation of overtone images, fundamental mode dispersion

curves were estimated. The curves then were inverted to obtain 1-D shear wave velocity

profiles (Figure 4.10).

Figure 4.10. The results of the borehole BH1. In the same graph, the 1-D shear wave
velocity profile deduced from the MASW survey along the Traverse No. 5 is shown. The
MASW method estimated with accuracy the shear wave velocity in the layers and
assisted in the layer identification of all the neighboring traverses.
81

Based on the author’s previous experience, shear wave velocity in excess of 1,500

ft/sec corresponds to top of weathered rock. The value was used for this study for the

purpose of mapping top of bedrock. To verify the interpretations, the MASW data were

compared with bore holes in terms of depth to bedrock. As seen in Figure 4.10, the depth

of interpreted top of bedrock on the MASW profile of collected by the 10 x 2.5ft. array

configuration corresponded well to the top of rock obtained by drilling.

4.3. ELECTRICAL RESISTIVITY TOMOGRAPHY (ERT)

4.3.1. Data Acquisition. ERT data were acquired along a total of sixteen

traverses oriented in W-E direction. The traverse orientations were selected to image the

dominant north-south trending joint sets in the study area. The ERT traverses were

spaced at 400ft. intervals.

The ERT data were acquired using an AGI SuperSting R8/IP, a multichannel

electrode system powered by two 12-Volts batteries (Figure 4.11). During the field set-

up, electrode cables were attached to metal stakes pounded to the ground using rubber-

band. Switch box was also used to connect the electrodes (passive electrodes) to the

resistivity meter. In an effort to ensure good quality ERT data are acquired, the field

crews routinely performed contact resistance tests. The contact resistance test was

performed prior data acquisition to ensure that all of the metal stakes are connected

properly to the cables and to the ground.

One hundred sixty eight electrodes were spaced at 5ft. intervals covering a length

of 835ft. and with expectation to image up to the depth of 100ft. As the length of the ERT

traverses exceeded the array length of 835ft., 50% roll-along overlap option was used. A
82

dipole-dipole electrode configuration was utilized in this study. This array gives a better

lateral resolution than the other types of arrays (Schlumberger and Wenner arrays) and is

more suitable for investigating karst terrain.

Figure 4.11. ERT field set up.

4.3.2. Data Processing and Interpretation. ERT data were processed using

Res2DInv software. As a first step of data processing, “bad” data points were removed

manually. The bad data points were either anomalously high or low values. Then,

“forward modeling” subroutine was used to specify the subsurface resistivity in order to

calculate the apparent resistivity that would be measured by a survey over such a

structure. As a result of data processing, 2D cross-sectional images were generated, as

shown in Figures 4.12 and 4.13.


83

Figure 4.12. Typical example of 2-D uninterpreted ERT model. Traverse No. P1 and
three overlapping ERT traverses acquired at the study site along 2,440ft.

Figure 4.13. Typical example of 2-D uninterpreted ERT model. Traverse No. P2 and
three overlapping ERT traverses acquired at the study site along 2,440ft.
84

Regarding the interpretation of the ERT data, , it was determined that moist soils

are characterized by resistivity values of less than 125 ohm-m; dry soils by resistivity

values greater than 125 ohm-m; moist weathered and/or fractured rock by resistivity

values less than 600 ohm-m; moist fractured rock with moist piped clay/soil-fill by

resistivity values less than 125 ohm-m; and drier, possibly less weathered rock by

resistivity values greater than 600 ohm-m (Anderson et al., 2006; Muchaidze, 2009; Myat

et al., 2008; Robison & Anderson, 2008). Figures 4.14 and 4.15 show the interpreted top

on two representatives ERT traverses (Traverse No. P1 and Traverse No. P2).

Figure 4.14. 2-D Interpreted versions of resistivity Traverse No. P1. Depth to top of
bedrock is around 16 ft. and corresponds to the 125 ohm-m contour interval.
85

Figure 4.15. 2-D interpreted versions of resistivity Traverse No. P2. Depth to top of
bedrock varies from 10 to 19 ft. and corresponds to the 125 ohm-m contour interval.

Boring control was used to verify the interpretations. Bore hole (BH1) was drilled

along ERT Traverse No. P1. The borehole was drilled to a depth of 98ft. (30.6m) below

subsurface. The top of rock was encountered at a depth of about 15ft. (4.6m), which was

consistent with the ERT interpretations (Figure 4.16).

The ERT data interpretations were also compared with MASW data

interpretations in terms of estimated depth to the top of bedrock (Figures. 4.17, 4.18, and

4.19). The comparison revealed a good agreement between the two.


86

Figure 4.16. Interpreted versions of resistivity Traverse No. P1. The top of rock
correlates reasonably well with the 125 ohm-m contour interval. The borehole location
has been superposed in a red dashed line. The 200ft. mark on the resistivity profile
corresponds with MASW profile No.1; the 800ft. mark corresponds with profile No. 2
location; the1, 400ft. mark corresponds with profile No.3; the1, 800ft. mark corresponds
with profile No. 4; and the 2,200ft. mark corresponds with profile No.5.
87

Figure 4.17. (A) ERT Traverse No. P1 tied with MASW profile No. 1. at the 200ft. (61m)
mark; (B) Corresponding 1-D shear-wave velocity profile. MASW depth to top of
weathered rock (“acoustic” top of rock) is identified at 15ft. (4.6m) depth. Red color line
on Figure 8.8 (B) indicates interpreted depth to top of rock.

Figure 4.18. (A) ERT Traverse No. P1 tied with MASW profile No. 2.at the 800ft.
(244m) mark; (B) Corresponding1-D shear-wave velocity profile. MASW depth to top of
weathered rock (“acoustic” top of rock) is identified on 7ft. (2.1m) depth. Red color line
on Figure 8.9 (B) indicates interpreted depth to top of rock.
88

Figure 4.19. (A) ERT Traverse No. P1 tied with MASW profile No. 5 and borehole
(BH1), at the 2,200ft. (670m) mark; (B) Corresponding 1-D shear-wave velocity profile.
MASW depth to top of weathered rock (“acoustic” top of rock) is identified at 8.5ft.
(2.6m) depth. Red color line on Figure 8.10 (B) indicates interpreted depth to top of rock.
89

5. COMPARATIVE ANALYSES

A comparative analysis is followed to analyze the performance of each MASW

configuration. Two types of comparative analyses are performed in this research:

qualitative and quantitative. The qualitative comparisons are solely based on visual

assessments of overtone images. The quantitative comparisons are based on the

assessment of the quantities calculated for each MASW configuration.

5.1. QUALITATIVE COMPARISON

The goal of qualitative analysis approach is twofold: 1) to compare overtone

images and dispersion curves; 2) to compare 1-D shear wave velocity profiles. The

comparisons are performed using different array configurations.

5.1.1. Comparison of Overtone Images and Dispersion Curves. Two MASW

profiles (Traverse No.1 and Traverse No. 2) collected by different array configurations

(geophone spacing, source offset, and the array orientation) were analyzed in attempt to

evaluate the effect of different parameters on the process of dispersion curve extraction

followed by modelling of 1-D shear wave velocity profiles. More specifically, the

MASW data results were analyzed for frequency and phase velocity ranges, interpreted

depth to top of rock, maximum depth of the investigation, signal-to-noise ratio, difference

between the theoretical and the experimental dispersion curves (root-mean-square error),

and the maximum number of iterations for inversion procedure.

5.1.2. MASW Traverse No. 1 Oriented West-East. Results from dispersion

curves and shear wave velocity profiles for all the tested parameters are presented in

Table 5.1.
90

Table 5.1. Results from all the dispersion curves and generated 1-D shear wave velocity
profiles for MASW Traverse No. 1. X1 - source offset; dx - geophone spacing; D -
receiver spread length; 𝜆𝑚𝑎𝑥 maximum resolvable Rayleigh wave wavelength; 𝜆𝑚𝑖𝑛
minimum resolvable Rayleigh wave wavelength; Z max penetration depth; S/N - signal-
to-noise ratio; RMS -root-mean-square error.

Figure 5.1A displays the MASW data collected along Traverse No.1 oriented

west-east with source offset of 10ft. and geophone spacing of 2.5ft. Phase velocities were

manually picked using 11 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.1B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.1C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 31 Hz to 𝑓𝑚𝑎𝑥 = 65 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,541ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 895ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 14ft. and 𝜆𝑚𝑎𝑥 = 55ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 18ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 98 % and an RMS error of 6.5 % with nine iterations. The
91

fundamental mode is clearly observable in the absence of a higher mode signs. The

dispersion curve is continuous and distinguishable at frequencies higher than 30 Hz. The

theoretical dispersion curve matches the experimental curve at higher frequencies. These

results can be due to the employed combination of the receiver spread length and the

source offset.

Figure 5.1. MASW survey data collected along Traverse No.1 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.
92

Figure 5.2A displays the MASW data collected along Traverse No.1 oriented

west-east with source offset of 10ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.2B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.2C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 30 Hz to 𝑓𝑚𝑎𝑥 = 53 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,696 ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 917ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 17ft. and 𝜆𝑚𝑎𝑥 = 56ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 19ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 88 % and an RMS error of 9.2 % with eight iterations. The

fundamental mode lacks separation from the higher mode (this overlap might be due to

the similarities in phase velocities in that frequency ranges). The dispersion curve is non-

coherent at frequencies higher than 30 Hz. The theoretical dispersion curve fairly

matches the experimental curve at higher frequencies. Low quality survey results can be

due to the employed combination of the receiver spread length and the source offset.
93

Figure 5.2. MASW survey data collected along Traverse No. 1 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

Figures 5.3 and 5.4 display the MASW data collected along Traverse No. 1

oriented west-east on the basis of the comparison between the MASW survey output for

the employed source offset of 30 ft. (9.14m) with a receiver spacing of 2.5ft. (0.76m) and

the source offset of 30ft. (9.14m) with a receiver spacing of 5ft. (1.52m), respectively.

Figures 5.3 and 5.4 show how the various combinations of source offset and geophone
94

spacing can affect the resulting overtone images and extracted dispersion curves for the

fundamental and higher modes.

Figure 5.3A displays the MASW data collected along Traverse No.1 oriented

west-east with source offset of 30ft. and geophone spacing of 2.5ft. Phase velocities were

manually picked using 8 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.3B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.3C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 31 Hz to 𝑓𝑚𝑎𝑥 = 54 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,533 ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 859 ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 16ft. and 𝜆𝑚𝑎𝑥 = 50ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 20ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 85 % and an RMS error of 13.2 % with six iterations. The

fundamental mode is hardly distinguishable and very limited in the presence of a higher

mode. The dispersion curve is continuous and distinguishable at frequencies higher than

30 Hz. The theoretical dispersion curve fairly matches the experimental curve at higher

frequencies. These results can be due to the employed combination of the receiver spread

length and the source offset.


95

Figure 5.3. MASW survey data collected along Traverse No.1 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

Figure 5.4A displays the MASW data collected along Traverse No.1 oriented

west-east with source offset of 30ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 8 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.4B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.4C). The dispersion analysis shows

that the data possesses dominant frequencies at approximately 𝑓𝑚𝑖𝑛 = 38 Hz to 𝑓𝑚𝑎𝑥 = 53


96

Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥 = 1,744ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 995ft/sec,

corresponding to the Rayleigh waves wavelength of 𝜆𝑚𝑖𝑛 = 19ft. and 𝜆𝑚𝑎𝑥 = 46ft. The

picked frequency and phase velocity ranges fall within the determined fundamental

mode. With these values and the traverse length, the minimum and maximum penetration

depths is 𝑍𝑚𝑎𝑥 = 19ft. These parameters yielded the inverted Vs model that was obtained

with the high signal-to-noise ratio (S/N) of 92 % and an RMS error of 12.6 % with six

iterations. The fundamental mode lacks separation from the higher mode (this overlap

might be due to the similarities in phase velocities in that frequency ranges). The

dispersion curve is non-coherent at frequencies higher than 30 Hz. The theoretical

dispersion curve fairly matches the experimental curve at higher frequencies. Low quality

survey results can be due to the employed combination of the receiver spread length and

the source offset.


97

Figure 5.4. MASW survey data collected along Traverse No.1 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

5.1.2. MASW Traverse No. 1 Oriented North to South. Figure 5.5A displays

the MASW data collected along Traverse No. 1 oriented west-east with source offset of

10ft. and geophone spacing of 2.5ft. Phase velocities were manually picked using 11

equally spaced points from the fundamental-mode dispersion curve in Figure 5.5B

(represented by white dots) and used for an inversion process to generate a 1-D shear

wave velocity model (Figure 5.5C).


98

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 27 Hz to 𝑓𝑚𝑎𝑥 = 47 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,740ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 850ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 18ft. and 𝜆𝑚𝑎𝑥 = 64ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 21.5ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 98 % and an RMS error of 4.8 % with nine iterations. The

fundamental mode is clearly observable in the absence of a higher mode signs. The

dispersion curve is continuous and distinguishable at frequencies higher than 30 Hz. The

theoretical dispersion curve matches the experimental curve at higher frequencies. These

results can be due to the employed combination of the receiver spread length and the

source offset.
99

Figure 5.5. MASW survey data collected along Traverse No. 1 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

Figure 5.6A displays the MASW data collected along Traverse No. 1 oriented

west-east with source offset of 10ft. and geophone spacing of 2.5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.6B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.6C).


100

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 28 Hz to 𝑓𝑚𝑎𝑥 = 47 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,515 ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 900ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 19ft. and 𝜆𝑚𝑎𝑥 = 54ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 17ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 70 % and an RMS error of 4 % with twelve iterations, the

lowest of the eight trials. The fundamental mode is observable and can be distinguished

from the higher mode. The dispersion curve is continuous at frequencies higher than 30

Hz. The theoretical dispersion curve fairly matches the experimental curve at higher

frequencies. These results can be due to the employed combination of the receiver spread

length and the source offset.


101

Figure 5.6. MASW survey data collected along Traverse No. 1 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

Figures 5.7 and 5.8 display the MASW data collected along Traverse No. 1

oriented north-south on the basis of the comparison between the MASW survey output

for the employed source offset of 30ft. (9.14m) with a receiver spacing of 2.5ft. (0.76m)

and the source offset of 30ft. (9.14m) with a receiver spacing of 5ft. (1.52m),

respectively. Figures 5.7 and 5.8 show how the various combinations of source offset and
102

geophone spacing can affect the resulting overtone images and extracted dispersion

curves for the fundamental and higher modes.

Figure 5.7A displays the MASW data collected along Traverse No. 1 oriented

north-south with source offset of 30ft. and geophone spacing of 2.5ft. Phase velocities

were manually picked using 9 equally spaced points from the fundamental-mode

dispersion curve in Figure 5.7B (represented by white dots) and used for an inversion

process to generate a 1-D shear wave velocity model (Figure 5.8C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 31 Hz to 𝑓𝑚𝑎𝑥 = 55 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,609 ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 805ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 15ft. and 𝜆𝑚𝑎𝑥 = 52ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 18ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 88 % and an RMS error of 19 % with six iterations. The

fundamental mode is hardly distinguishable and limited in the presence of a higher mode.

The dispersion curve is continuous and distinguishable at frequencies higher than 30 Hz.

The theoretical dispersion curve fairly matches the experimental curve at higher

frequencies. Low quality survey results can be due to the employed combination of the

receiver spread length and the source offset.


103

Figure 5.7 MASW survey data collected along Traverse No. 1 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

Figure 5.8A displays the MASW data collected along Traverse No. 1 oriented

north-south with source offset of 30ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 7 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.8B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.8C).


104

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 30 Hz to 𝑓𝑚𝑎𝑥 = 47 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,688ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 874ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 19ft. and 𝜆𝑚𝑎𝑥 = 56ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 18ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 76 % and an RMS error of 12 % with ten iterations. The

fundamental mode is hardly distinguishable and limited in the presence of a higher mode.

The dispersion curve is continuous and distinguishable at frequencies higher than 30 Hz.

The theoretical dispersion curve fairly matches the experimental curve at higher

frequencies. Low quality survey results can be due to the employed combination of the

receiver spread length and the source offset.


105

Figure 5.8. MASW survey data collected along Traverse No. 1 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency) where the fundamental
mode is quite clear. (C) The 1-D shear wave velocity profile, deduced from the inversion
technique. 1) Risk of higher mode; 2) The signal-to-noise ratio (S/N); 3) R.M.S. error; 4)
Iterations.

Eight configurations of MASW survey setup were tested. An average dispersion

curve was obtained by combining all 8 records using the frequency range from 27 to 54

Hz and the phase velocity range from 805ft/sec to 1,696ft/sec. This process was repeated

for all of the receiver array types, with the dominant frequency of the surface waves

varied from 30 to 53 Hz for the various array types. An excellent signal-to-noise ratio

was obtained for all of the records until a satisfactory match was reached between the
106

experimental and the theoretical dispersion curves with a relative error of less than 10%

to 15%.

The following conditions were common for all three inversions: 1) the 1-D shear-

wave velocity profile geometry was defined by ten model blocks (layers) with fixed

thicknesses increasing by 5% with depth, plus a model half-space; 2) the depth to the

average model half-space was determined as 18.6ft., corresponding to approximately

one-third of the maximum resolvable wavelength; 3) a maximum of 8 iterations is

possible.

5.1.3. MASW Traverse No. 2 Oriented West-East. Results from dispersion

curves and shear wave velocity profiles for all the tested parameters are presented in

Table 5.2.

Table 5.2. Results from all the dispersion curves and generated 1- D shear wave velocity
profiles for MASW Traverse No. 2. X1 - source offset; dx - geophone spacing; D -
receiver spread length; 𝜆𝑚𝑎𝑥 maximum resolvable Rayleigh wave wavelength; 𝜆𝑚𝑖𝑛
minimum resolvable Rayleigh wave wavelength; Z max penetration depth; S/N - signal-
to-noise ratio; RMS -root-mean-square error.

Figure 5.9A displays the MASW data collected along Traverse No. 2 oriented

west-east with source offset of 10ft. and geophone spacing of 2.5ft. Phase velocities were
107

manually picked from the fundamental-mode dispersion curve in Figure 5.9B

(represented by white dots) and used for an inversion process to generate a 1-D shear

wave velocity model (Figure 5.9C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 31 Hz to 𝑓𝑚𝑎𝑥 = 55 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,228ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 839ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 13.7ft. and 𝜆𝑚𝑎𝑥 = 40ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 13ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 95 % and an RMS error of 5.2 % with eight iterations. The

fundamental mode is clearly observable in the absence of higher modes. The dispersion

curve is continuous and distinguishable at frequencies higher than 30 Hz. The theoretical

dispersion curve matches the experimental curve at higher frequencies. These results can

be due to the employed combination of the receiver spread length and the source offset.
108

Figure 5.9. MASW survey data collected along Traverse No. 2 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Figure 5.10A displays the MASW data collected along Traverse No. 2 oriented

west-east with source offset of 10ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.10B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.10C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 31 Hz to 𝑓𝑚𝑎𝑥 = 55 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥
109

= 1,247ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 928ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 17ft. and 𝜆𝑚𝑎𝑥 = 40ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 13ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 85 % and an RMS error of 9 % with eight iterations. The

fundamental mode is observable in the presence of a higher mode. The dispersion curve

is continuous and distinguishable at frequencies higher than 30 Hz. The theoretical

dispersion curve matches the experimental curve at higher frequencies. Low quality

survey results can be due to the employed combination of the receiver spread length and

the source offset.


110

Figure 5.10. MASW survey data collected along Traverse No. 2 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Figures 5.11 and 5.12 display the MASW data collected along Traverse No. 2

oriented west-east on the basis of the comparison between the MASW survey output for

the employed source offset of 30ft. (9.14m) with a receiver spacing of 2.5ft. (0.76m) and

the source offset of 30ft. (9.14m) with a receiver spacing of 5ft. (1.52m), respectively.

Figures 5.11 and 5.12 show how the various combinations of source offset and geophone

spacing can affect the resulting overtone images and extracted dispersion curves for the

fundamental and higher modes.


111

Figure 5.11A displays the MASW data collected along Traverse No. 2 oriented

west-east with source offset of 30ft. and geophone spacing of 2.5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.11B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.11C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 25 Hz to 𝑓𝑚𝑎𝑥 = 53 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,670ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 899ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 15ft. and 𝜆𝑚𝑎𝑥 = 43ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths are 𝑍𝑚𝑎𝑥 = 14ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 98 % and an RMS error of 9.5 % with ten iterations. The

fundamental mode lacks separation from the higher mode (this overlap might be due to

the similarities in phase velocities in that frequency ranges). The dispersion curve is

moderately distinguishable at frequencies higher than 30 Hz. The theoretical dispersion

curve fairly matches the experimental curve at higher frequencies. Low quality survey

results can be due to the employed combination of the receiver spread length and the

source offset.
112

Figure 5.11. MASW survey data collected along Traverse No. 2 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Figure 5.12A displays the MASW data collected along Traverse No. 2 oriented

west-east with source offset of 30ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.12B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.12C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 29 Hz to 𝑓𝑚𝑎𝑥 = 58 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥
113

= 1,221ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 913ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 15ft. and 𝜆𝑚𝑎𝑥 = 43ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths are 𝑍𝑚𝑎𝑥 = 14ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 94 % and an RMS error of 2.9 % with ten iterations. The

fundamental mode is visible and can be separated from the higher mode. The dispersion

curve is moderately distinguishable at frequencies higher than 30 Hz. The theoretical

dispersion curve fairly matches the experimental curve at higher frequencies. Low quality

survey results can be due to the employed combination of the receiver spread length and

the source offset.


114

Figure 5.12. MASW survey data collected along Traverse No. 2 oriented west-east array.
(A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

5.1.4. MASW Traverse No. 2 Oriented North-South. Figure 5.13A displays the

MASW data collected along Traverse No.2 oriented north-south with source offset of

10ft. and geophone spacing of 2.5ft. Phase velocities were manually picked using 10

equally spaced points from the fundamental-mode dispersion curve in Figure 5.13B

(represented by white dots) and used for an inversion process to generate a 1-D shear

wave velocity model (Figure 5.13C).


115

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 29 Hz to 𝑓𝑚𝑎𝑥 = 62 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,312ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 834ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 13ft. and 𝜆𝑚𝑎𝑥 = 55ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 15ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 98 % and an RMS error of 2.7 % with ten iterations. The

fundamental mode is clearly observable in the absence of higher mods. The dispersion

curve is continuous and distinguishable at frequencies higher than 30 Hz. The theoretical

dispersion curve matches the experimental curve at higher frequencies. These results can

be due to the employed combination of the receiver spread length and the source offset.
116

Figure 5.13. MASW survey data collected along Traverse No.2 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Figure 5.14A displays the MASW data collected along Traverse No.2 oriented

north-south with source offset of 10ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.14B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.14C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 30 Hz to 𝑓𝑚𝑎𝑥 = 58 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥
117

= 1,515ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 900ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 16ft. and 𝜆𝑚𝑎𝑥 = 51ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 17ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 95 % and an RMS error of 10 % with ten iterations, the

lowest of the eight trials. The fundamental mode is clearly observable in the absence of

higher modes. The dispersion curve is continuous and distinguishable at frequencies

higher than 30 Hz. The theoretical dispersion curve matches the experimental curve at

higher frequencies. These results can be due to the employed combination of the receiver

spread length and the source offset.


118

Figure 5.14. MASW survey data collected along Traverse No.2 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Figures 5.15 and 5.16 display the MASW data collected along Traverse No.2

oriented north-south on the basis of the comparison between the MASW survey output

for the employed source offset of 30ft. (9.14m) with a receiver spacing of 2.5ft. (0.76 m)

and the source offset of 30ft. (9.14m) with a receiver spacing of 5ft. (1.52m),

respectively. Figures 5.15 and 5.16 show how the various combinations of source offset
119

and geophone spacing can affect the resulting overtone images and extracted dispersion

curves for the fundamental and higher modes.

Figure 5.15A displays the MASW data collected along Traverse No.2 oriented

north-south with source offset of 30ft. and geophone spacing of 2.5ft. Phase velocities

were manually picked using 10 equally spaced points from the fundamental-mode

dispersion curve in Figure 5.15B (represented by white dots) and used for an inversion

process to generate a 1-D shear wave velocity model (Figure 5.15C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 30 Hz to 𝑓𝑚𝑎𝑥 = 62 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥

= 1,346ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 854ft/sec, corresponding to the Rayleigh waves wavelength

of 𝜆𝑚𝑖𝑛 = 14ft. and 𝜆𝑚𝑎𝑥 = 45ft. The picked frequency and phase velocity ranges fall

within the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 15ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 99 % and an RMS error of 11 % with ten iterations. The

fundamental mode is clearly observable in the absence of higher modes. The dispersion

curve is continuous and distinguishable at frequencies higher than 30 Hz. The theoretical

dispersion curve matches the experimental curve at higher frequencies. These results can

be due to the employed combination of the receiver spread length and the source offset.
120

Figure 5.15. MASW survey data collected along Traverse No.2 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Figure 5.16A displays the MASW data collected along Traverse No.2 oriented

north-south with source offset of 30ft. and geophone spacing of 5ft. Phase velocities were

manually picked using 9 equally spaced points from the fundamental-mode dispersion

curve in Figure 5.16B (represented by white dots) and used for an inversion process to

generate a 1-D shear wave velocity model (Figure 5.16C).

The dispersion analysis shows that the data possesses dominant frequencies at

approximately 𝑓𝑚𝑖𝑛 = 29 Hz to 𝑓𝑚𝑎𝑥 = 62 Hz. The phase velocity varied between 𝑉𝑝ℎ𝑚𝑎𝑥
121

= 1,184ft/sec and 𝑉𝑝ℎ𝑚𝑖𝑛 = 913ft/sec, corresponding to the Rayleigh waves wavelength of

𝜆𝑚𝑖𝑛 = 15ft. and 𝜆𝑚𝑎𝑥 = 41ft. The picked frequency and phase velocity ranges fall within

the determined fundamental mode. With these values and the traverse length, the

minimum and maximum penetration depths is 𝑍𝑚𝑎𝑥 = 14ft.

These parameters yielded the inverted Vs model that was obtained with the high

signal-to-noise ratio (S/N) of 91 % and an RMS error of 17 % with ten iterations. The

fundamental mode lacks separation from the higher mode (this overlap might be due to

the similarities in phase velocities in that frequency ranges). The dispersion curve is non-

coherent at frequencies higher than 30 Hz. The theoretical dispersion curve fairly

matches the experimental curve at higher frequencies. Low quality survey results can be

due to the employed combination of the receiver spread length and the source offset.
122

Figure 5.16. MASW survey data collected along Traverse No.2 oriented north to south
array. (A) The 24 channel record (shot gather); (B) The dispersion image (overtone) with
superposed dispersion curve (phase velocity versus frequency); (C) The 1-D shear wave
velocity profile, deduced from the inversion technique. 1) Risk of higher mode; 2) The
signal-to-noise ratio (S/N); 3) R.M.S. error; 4) Iterations.

Eight configurations of MASW survey setup were tested. An average dispersion

curve was obtained by combining all 8 records using the frequency range from 29 to 62

Hz and the phase velocity range from 834 ft/s to 1,541 ft/s. This process was repeated for

all of the receiver array types, with the dominant frequency of the surface waves varied

from 30 to 53 Hz for the various array types. An excellent signal-to-noise ratio was

obtained for all of the records until a satisfactory match was reached between the
123

experimental and the theoretical dispersion curves with a relative error of less than 10%

to 15%.

The following conditions were common for all three inversions: 1) the 1-D shear-

wave velocity profile geometry was defined by ten model blocks (layers) with fixed

thicknesses increasing by 5% with depth, plus a model half-space; 2) the depth to the

average model half-space was determined as 15ft., corresponding to approximately one-

third of the maximum resolvable wavelength; 3) a maximum of 9.7 iterations is possible.

When a satisfactory match is reached between the experimental and the numerical

dispersion curves (usually with a relative error of less than 10% to 15%), then the Vs

profile has been found and the researcher can proceed with determining the seismic

subsurface classification.

Generally, velocity values of less than 1,200ft/sec (366m/sec) are interpreted as

sand and unconsolidated material, while velocity values greater than 1,200ft/s (366m/s)

are interpreted as soft, weathered limestone.

5.2. DISPERSION OF RESOLUTION CURVE

Resolution of dispersion curve is another important factor in determining quality

of MASW data. The resolution along the velocity axis defines the ability to discriminate

a particular phase velocity from other velocities for a given frequency. The resolution

along the frequency axis determines the ability to discriminate a particular frequency

from other frequencies for a given velocity. The dispersion curve width is related to the

relationship between the frequency and the velocity of the surface waves.
124

To determine the resolution of the dispersion image in the frequency-velocity

domain (f-v), the length of the receiver spread needs to be increased (Park et al., 1998).

Forbriger (2003a) provided an analytical result to assess the resolution of the dispersion

image based on the Equation (5.1).

Δd = 1 / f D (5.1)

where Δd is the half-width between the neighboring minimum of dispersion

energy in the (f-v) domain, f is the frequency, and D is the length of the receiver array.

However, then understand that the resolution of the dispersion image could vary with the

algorithms that were used to generate the dispersion image in the frequency-velocity

domain. A current comparison of several different algorithms can be found in (Dal Moro

et al., 2003).

Three to five impacts generated by a 20 lb. sledgehammer were vertically stacked

at each MASW location using SurfSeis software developed by KGS. A record length of

2,048 milliseconds at a 1-millisecond sampling rate interval was employed.

The influence of these factors on the analysis is highly dependent on the data

quality such as high or low the signal-to-noise ratio, the confidence of dispersion curve

extraction, the R.M.S. error and the maximum number of iterations. A “good” quality

data set suggests the surface-wave energy is the most prominent seismic event in the

seismic record (i.e., the highest S/N), whereas a “bad” quality data sets are usually

contaminated by noise.

The quality of the acquired surface wave records can be evaluated in terms of the

resolution of the phase velocity spectrum, i.e., the sharpness of the amplitude peaks
125

observed at each frequency, the extractable frequency range and the continuity of the

fundamental mode high-amplitude band.

The quality of a dispersion curve is judged according to two criteria: the signal-to-

noise ratio and the general dispersion curve direction. A high S/N indicates a high

confidence in the obtained phase velocity. An S/N higher than 0.5 or more than 55% is

considered acceptable.

The quality of the MASW data can also be affected by natural geologic conditions

that may not produce well-defined dispersion curves and cannot be used to calculate

reliable shear-wave velocity models, only on the basis of the comparison between the

nearest source offset and variable receiver spacings. After that, the comparison between

the farther source offset and variable receiver spacings.

Two configurations of MASW geophone array results were generated with the

data Figure 5.17. The first nearest source offset was (X1) = 10ft. with a receiver spacing

of (dx) = 2.5ft., and the second nearest source offset was (X1) = 10ft. with a receiver

spacing of (dx) = 5ft.

The output of two different MASW array configurations are shown in Figure

5.17. Figure 5.17A and 5.17B show seismic field records collected using geophone

spacing of 2.5ft. (0.76m) with corresponding dispersive image; Figures 5.17 C and 5.17D

show seismic field records collected using geophone spacing of 5ft. (1.52m) with

corresponding dispersive image. For the estimation of dispersion curve resolution, a

double-ended arrow at frequencies of 30 Hz and 50 Hz is superposed over the dispersion

image. The double-ended arrow at 30 Hz is twice as long as that at 50 Hz. This is

determined by Equation (5.1) that indicates the resolution of the dispersion image is one-
126

half at 30 Hz compared with that at 50 Hz. For Figures 5.17C and 5.17D, the resolution

reduction of one-half (Δd is doubled) as shown in Figure 5.17D was expected. The

double-ended arrows in Figures 5.17A and 5.17B at 30 and 50 Hz are longer than 60%,

not 100% like the arrows in Figure 5.17C and 5.17D at 30 and 50 Hz, respectively.

Figure 5.17. The multichannel seismic record; (B) Corresponding overtone image in
which the fundamental mode dispersion is identified; (C) The multichannel seismic
record collected with doubled receiver spacing; (D) Corresponding overtone image in
which the fundamental mode dispersion is identified with noticeable discrepancies.

The third and fourth MASW array configurations are shown in Figure 5.18. The

third farthest source offset was 30 ft. (9.14m) with a receiver spacing of 2.5 ft. (0.76m),

and the fourth farthest source offset was 30 ft. (9.14m) with a receiver spacing of 5ft.

(1.52m).

The output of the third and fourth MASW array configurations are shown in

Figure 5.18. Figure 5.18A and 5.18B show seismic field records collected using

geophone spacing of 2.5ft. (0.76m) with corresponding dispersive image; Figures 5.18C
127

and 5.18D show seismic field records collected using geophone spacing of 5ft. (1.52m)

with corresponding dispersive image. For the estimation of dispersion curve resolution, a

double-ended arrow at frequencies of 30 Hz and 50 Hz is superposed over the dispersion

image. The double-ended arrow at 30 Hz is twice as long as that at 50 Hz. Both double-

ended arrows in Figure 5.18C and Figure 5.18D at 30 and 50 Hz were the same length

lengths, which also supports Equation (5.1) in that the resolution of the dispersion image

is determined by the length of the receiver array and the receiver spacing. Also, the

resolution (indicated by a double-ended arrow around the frequency of 30 Hz) of the

dispersive energy from the data with the farther source offset of 30ft. (9.14m) is better

than that of the data generated from seismic record collected with the nearest source

offset of 10ft. (3.04m).

As demonstrated by previous examples, Equation (5.1) might work well as the

frequency changes when the length of the receiver array is fixed. Furthermore, it is

usually necessary to double the length of the receiver array for a given frequency, and the

resolution increase is normally less than 90%. A comparison of the resolution of

dispersion image generated from seismic record collected with the father source offset of

30 ft. (9.14m) at both Figures 5.18c and 5.18d is much better than in Figures 5.18a and

5.18B, and the double-ended arrow is 10% shorter than on the data acquired with the

nearest source offset of 10ft. (3.04m) as it shown in Figure 5.18B and 5.18D.

Analytically, it can be shown that the width of a dispersion curve depends on the

difference between the true phase velocity and the assumed phase velocity. If the

integrative energy rapidly decreases with the difference in velocity, the bandwidth would

be relatively narrow, giving a relatively well-defined velocity range that can be regarded
128

as the phase velocity. The relationship between the energy bandwidth and the phase range

of the harmonic waves facilitates an understanding of how the bandwidth varies with

phase velocity, frequency and source distance.

Figure 5.18. (A) The multichannel seismic record; (B) Corresponding overtone image in
which the fundamental mode dispersion is identified; (C) The multichannel seismic
record collected with doubled receiver spacing; (D) Corresponding overtone image in
which the fundamental mode dispersion is identified with visible distortion.

Figure 5.19 displays 4 dispersion images (overtones) deduced from the surface

wave records collected using various MASW array configurations. Figure 5.19A displays

dispersion curve extracted from fundamental mode of overtone image generated for the

seismic record collected with source offset of 10ft. (3.05m) and a receiver spacing of 2.5

ft. (0.76m), the longest Rayleigh wave wavelength of λmax = 1,663ft/sec/26 Hz = 64ft.,

the minimum phase velocity of CRmin = 827ft/sec and a difference between the

maximum and minimum phase velocities of ΔCR = 1,663ft/sec - 827ft/sec = 836ft/sec.

Therefore, the suggested optimum offset A would be 16ft.


129

Figure 5.19B displays dispersion curve extracted from fundamental mode of

dispersion image generated for the seismic record collected with source offset of 10ft.

(3.05m) and a receiver spacing of 5ft., the longest Rayleigh wavelength of λmax =

1,220ft/sec/31 Hz = 39ft., the minimum phase velocity of CRmin = 850 ft/sec and a

difference between the maximum and minimum phase velocities of ΔCR = 1,220ft/sec -

850ft/sec = 370ft/sec. Therefore, the suggested optimum offset A would be 22.3ft.

High level of Rayleigh waves coherency was noticed during comparison of

Figures 5.19a and 5.19b. It was observed that the data acquired with a configuration of

the first nearest offset of 10ft. (3.04m) and a receiver spacing of 2.5ft. (0.61m) considered

a better quality than that from the second nearest offset of 10ft. and a receiver spacing of

5ft. (1.52m). The continuity is even clearer for data in a rectangular window (especially

for frequencies lower than 30 Hz).

Figure 5.19C displays dispersion curve extracted from fundamental mode of

dispersion image generated for the seismic record collected with source offset of 30ft.

(9.14m) and a receiver spacing of 2.5ft. (0.61m), the longest Rayleigh wavelength of

λmax = 1,699ft/sec/25 Hz = 68ft. (20.72m), a minimum phase velocity of CRmin =

855ft/sec and a difference between the maximum and minimum phase velocities of ΔCR =

1,699ft/sec - 855ft/sec = 844ft/sec. Therefore, the suggested optimum offset A would be

17.2ft. (5.18m).

Figure 5.19D displays dispersion curve extracted from fundamental mode of

dispersion image generated for the seismic record collected with source offset of 5ft.

(1.52m), the longest Rayleigh wavelength of λmax = 1,546ft/sec/26 Hz = 59ft. (17.98 m),

a minimum phase velocity of CRmin = 904ft/sec and a difference between the maximum
130

and minimum phase velocities of ΔCR = 1,546ft/sec - 904ft/sec = 642ft/sec. Therefore, the

suggested optimum offset A would be 20.7ft. (6.01m).

After comparing all of the results, high level of Rayleigh waves coherency was

noticed in the data acquired from the nearest offset of 10ft. and a receiver spacing of

2.5ft. (0.61m) and the farthest offset of 30ft. (9.14m) and a receiver spacing of 2.5ft.

(0.61m) than with the farther offset of = 30ft. (9.14m) This is the most reliable method of

dispersion imaging with the lowest number of computational artifacts.

Figure 5.19. Dispersion images (overtones) deduced from the surface wave records
collected using various MASW array configurations (north to south orientation).
131

Similar analysis was performed for W-E oriented arrays (Figure 5.20). It was also

noticed that the overal quality of overtone images obtained by using west to east oriented

arrays are slightly poorer in terms of continuity of dispersion curve. This can be

attributed to the irregularity of the depth to top of rock. As the dominant solutuon

widening joints trend north-south in the study area, qulity of MASW data acquired

perpendicular to the structural trends is porrer due to the significant relief along the

geophone array.

Figure 5.20. Dispersion images (overtones) deduced from the surface wave records
collected using various MASW array configurations (west to east orientation).
132

5.3. COMPARISON OF 1-D SHEAR-WAVE VELOCITY PROFILES

The variable increases in shear-wave velocity with depth were attributed to

fracturing of the limestone bedrock. Generally, high shear-wave velocities are indicative

of limestone bedrock. Fracture zones and highly weathered bedrock are typically

attributed to the shear-wave velocities between 1,200 and 2,500ft/sec. Burlington-Keokuk

limestone in the study area has shear-wave velocities between 2,000 and 2,500 ft/sec, so

observed velocities may also be relevant to another limestone type with similar lithology.

Data collected using four MASW array configurations were analyzed. The first

and second data sets were acquired using the source offsets of 10ft. (3.04m) with a

receiver spacing of 2.5ft. (0.76m) and 5ft. (1.52m), respectively. The third and fourth data

sets were acquired using the source offset of 30 ft. (9.14m) with a receiver spacing of

2.5ft. (0.76m) and 5ft. (1.52m), respectively.

Figures 5.21 and 5.22 display the comparison of 1-D shear-wave velocity profiles,

deduced using the inversion technique from the seismic data sets collected with various

configurations of active MASW array. Table 5.3 lists the average Vs values for the top

4ft. (1.22m), 12.5ft. (3.81m), and 27.5ft. (8.38m) depths for one site.

Vs profiles were generated from seismic records acquired with different source

offsets and receiver spacings that were oriented from west to east. The 1-D shear-wave

velocity profiles obtained with different types of configuration of MASW array

configurations are shown in Figure 5.21. The variation of the Vs velocities at all depths

seems to be fairly consistent for all array configurations. The depth to the top of rock can

be identified at a depth of around 12.5ft. (3.81m) from the data collected with all array

configurations.
133

Table 5.3. Average Vs at ten 1-D shear-wave velocity profiles. No.1.

A comparison of 1-D shear wave velocity profiles generated up to a depth of

32.5ft. (9.9m) shows the following results:

a) All of the profiles start with a thin layer (from 1 to 4ft. [0.3 – 1.22m] thick) and

are characterized by average shear wave velocities between 1,070 and 1,240ft./s)

for the data collected with all array configurations. This high-velocity layer is

most probably observed due to the combination of stiff subsurface materials

(reddish-brown clay with admixed chert fragments, alternating from fine to coarse

grained with occasional stylolites and fossils).

b) There is a low velocity layer observed on depths of 11.5 - 12.5ft. (3.5 – 3.81m)

with corresponding shear wave velocities range between 1,230 and 1,250ft./s

which is confidently determined in the most profiles. This layer is attributed to the

top of rock (represented by Burlington-Keokuk limestone).

c) At a depth of approximately 27.5ft. (8.38m), corresponding 1-D shear wave

velocity values increase gradually until the depth of 32.5ft. (9.9m) from

1370ft/sec to 1729ft/sec in all MASW profiles.


134

d) It can also be observed from the Vs profiles that there is a slight decrease in the

velocities for the data sets collected using the farthest source offset of 30ft.

(9.14m) and longer receiver spacing of 5ft. (1.52m) at a depth of approximately

27.5ft. (8.38m). Decrease in velocities from 1,729ft/sec to 2,066ft/sec were

observed for the data sets collected using the nearest offset of 10ft. (3.04m) for

the data sets collected in west-east direction and approximately 1,370ft/sec to

1,404ft/sec for the data sets collected in north-south direction.

e) This decrease in velocity values may correspond to the presence of either clay or

saprolite formation above the bedrock. Also, the highest velocity found in the

inversion was around 1,200 to 2,500ft/sec which are indicative of bedrock, which

is characterized by calcareous shale with some limestone interbeds. A reasonable

correlation between the MASW data sets collected with opposite array orientation

was observed. Generally, extraction of dispersion curves from the overtone

images generated from seismic records collected in west-east or north-south

direction provided results of similar quality. That can be explained by the

heterogeneity of subsurface.
135

Figure 5.21. Comparison of 1-D shear wave velocity profiles, deduced using the
inversion technique from the seismic data sets collected with various configurations of
active MASW oriented west to east array.

Figure 5.22. Comparison of 1-D shear wave velocity profiles, deduced using the
inversion technique from the seismic data sets collected with various configurations of
active MASW oriented north to south array.
136

5.4. QUANTITATIVE ANALYSES

As a first step of quantitative analysis, three criterias were chosen, and include

continuity of dispersion curve, maximum depth of investigation, and reliability of the top

of rock estimate. The rating categories are briefly summarized in Table 5.4.

Table 5.4. Encoded categories for three criterias.

The continuity was analyzed on the basis of visual assessment of overtone images

and dispersion curves. The quality of the acquired surface wave records were evaluated in

terms of the resolution of the phase velocity spectrum, i.e., the sharpness of the amplitude

peaks observed at each frequency, the extractable frequency range and the continuity of

the fundamental mode high-amplitude band. The continuity ratings “Good”, “Fair”, and

“Poor” were assigned on the basis of visual assessment of overtone images and

dispersion curves, as shown in Fig. 5.23. To ease quantitative data analysis, each rating

was encoded numerically. Thus, number 30 was assigned to “good”, number 15 was

assigned to “fair”, and number 0 was assigned to “poor”.


137

Figure 5.23 Overtone images obtained from MASW data of (A) good quality, (B) fair
quality, and (C) poor quality.

Regarding to the maximum depth of MASW profiles, number 0 was assigned to

profiles that do not extend to a depth of 20ft., number 5 was assigned to profiles that do

extend to a depth of 30ft., and number 10 was assigned to profiles that extend to a depth

greater than 30ft.

The analysis of reliability of the top of rock estimate was based on the comparison

of ERT- and MASW-determined depth to top of bedrock. The difference between the two

was calculated in percent. Number Zero was assigned to profiles that gave the difference

more than 30%, number 5 was assigned to profiles that gave the difference in the interval

of 20 to 30%, and number of 10 was assigned to profiles that had difference less than

20%.
138

The quantitative data analysis was performed by summing up the ratings of tree

criterias. Thus, a number of 50 is the maximum possible number, and a number of Zero is

the minimum number. The data quality categories were developed based on the numbers

after summation. The categories are briefly summarized in Table 5.5. Using this

approach, all MASW profiles were evaluated in order to assign the data quality ratings.

Table 5.5. Data quality categories.

In an attempt to identify the MASW array configuration that gives the better data

quality, statistical analysis was performed. During the analysis, mean values were

calculated for each group of MASW array configuration. Table 5.6 summarizes statistics

for west to east direction arrays, and Table 5.7 summarizes statistics for north to south

direction arrays. As seen in Table 5.6, the highest mean value (30.73) was observed for

10ft. source offset and 2.5ft.geophone spacing array. This implies that this configuration

gave the better quality of MASW data. The lowest mean value (21.91) was observed for

30 ft. source offset and 5ft. receiver spacing array. It is interesting to note that according

to the statistical analysis, the shortest geophone spread (10 x 2.5ft.) gave the better data

quality, whereas the longer geophone spread (30 x 5ft.) gave the poorer quality. This
139

supports the previous statement about the effect of irregularity of the depth to top of rock

along short and long geophone arrays. According to Table 5.7, mean values calculated for

configurations direction north to south are slightly higher than those calculated for

configurations direction west to east (Table 5.6).

Table 5.6. Quantitative analysis performed for 136 MASW profiles acquired in west to
east direction.

Table 5.7. Quantitative analysis performed for 136 MASW profiles acquired in north to
south direction.

To analyze the frequency of occurrence of MASW profiles with good, fair, poor,

and severe quality, Figures 5.24 and 5.25 were generated. Figure 5.24 shows distribution

of MASW data acquired by the array oriented in west to east direction and Figure 5.25

shows MASW data acquired by the array oriented in north to south direction. According
140

to Figure 5.24, the largest number of good quality MASW profiles was acquired using

10ft. source offset and 2.5ft. receiver spacing. In contrast, the smallest number of good

quality MASW profiles was acquired by the configuration with 30 ft. source offset and

5ft. receiver spacing.

Figure 5.24 Histogram showing how often MASW data of good, fair, poor, and severe
quality were acquired using a specific array configuration (west to east direction).

Regarding to the north to south oriented arrays, the largest number of MASW

profiles with good data quality was acquired by the 10ft. source offset and 2.5ft. receiver

spacing array. It is interesting to note that no MASW profiles with good data quality were

acquired using the longest (30 x 5ft.) geophone spread. In contrast, this configuration (30

x 5ft.) gave the largest number of MASW profiles with severe data quality.
141

Figure 5.25 Histogram showing how often MASW data of good, fair, poor, and severe
quality were acquired using a specific array configuration (north to south direction ).
142

6. CONCLUSIONS

The main goal of this research was to develop optimum MASW configuration that

can be used in karst terrain to image subsurface to a depth of 30ft. (9m). To accomplish

the goal, MASW data were acquired with different parameters and in different directions.

Qualitative and quantitative analyses were performed in this study to assess the

performance of each configuration.

The following conclusions can be made on the basis of qualitative analysis:

 Continuity of dispersion curve must be the primary criteria in evaluating MASW

data quality. The dispersion is more continuous over a broad range of frequency

when MASW data are acquired parallel to the solution widened joints.

 As the solution widened joints trend north to south, MASW data acquired by the

north to south oriented arrays are of a better quality.

The following conclusions can be made on the basis of quantitative analysis:

 In addition to the continuity of dispersion curve, two criterias – maximum depth

of investigation and reliability to top of rock estimate – can be used to evaluate

overall quality of MASW data. Using three criterias, it was determined that

MASW data acquired by 10ft. source offset and 2.5ft. receiver spacing array

configurations are of a better quality.


143

7. RECOMMENDATIONS

This research has enhanced the understanding of acquisition of MASW data in

karst terrain. However, further studies are still required to improve acquisition,

processing and interpretation of seismic data in karst terrain. The following studies are

recommended:

 Choice of the method should be guided by the anticipated depth and size of

target(s), the nature of the background materials or the maximum depth of rock

surrounding the target(s), the reason for delineating the target(s), the desired

resolution of the target(s), and the size of the investigation area and the sources of

cultural interference in the investigation area.

 The values that are determined for optimum data acquisition and processing

should be taken as guidelines for all other parameters (i.e., it is important to

choose processing parameters that are appropriate for the depth target of the

investigation).

 For impact sources, such as the drop weight or vibrator should be used to enhance

the quality of the active MASW data in complex karst terrain.

 A quantitative evaluation of the resolution of surface wave methods should be

conducted to clearly delineate the examination capabilities not only for

researchers, but as well to associate engineers avoid error application of surface

wave methods.

 In addition, calculation efforts should be made to improve the efficiency of

intensive inversion algorithms using parallel programming and computing on

GPS or GIS.
144

BIBLIOGRAPHY

Abbiss, C.P. (1981). Shear wave measurements of the elasticity of the ground.
Geotechnique, 31, (1), pp91-104.

Advanced Geosciences, Incorporated. "Instruction Manual for the Super Sting with Swift
Automatic Resistivity and IP System." Austin, January 2006.

Aki, K., & Richards, P. G. (1980). Quantitative seismology: Theory and methods (Vol.
1). San Francisco, CA: W. H. Freeman and Co.

Anderson, N. (2010). Lectures in: Ground Penetrating Radar Applications in Engineering


and Environmental Studies. GE482. Rolla, Missouri, USA: Missouri University of
Science and Technology.

Anderson, N., 2012, Determination of Optimum “Multi-Channel Surface Wave Method”


Field Parameters: NUTC R292.

Auld, B. A., Acoustic Fields and Waves in Solids (Stanford University Press, 1973).
EWING, W.M., JARDETZKY, W. S., and PRESS, F., Elastic Waves in Layered
Media (McGraw-Hill, London, 1957), 380 pp.

Bacon, M., Simm, R and Redshaw, T. (2003): 3-D Seismic Interpretation; Cambridge
University press.

Baechle, G.T., G.P. Eberli, R.J. Weger, and J.L. Massafero, 2009, Changes in dynamic
shear moduli of carbonate rocks with fluid substitution: Geophysics, 74, E135-
E147.

Ballard, R.F. & McLean, F.G. 1975. Seismic field methods for in situ moduli. Proc. Conf.
on In Situ Measurement of Soil Properties. Spec. Conf. Geotech. Eng. Div.,
A.S.C.E., Vol 1, pp121-150.

Beaty, K. S., Schmitt, D. R., & Sacchi, M. (2002). Simulated annealing inversion of
multimode Rayleigh wave dispersion curves for geological structure. Geophysical
Journal International, 151(2), 622-631.

Boore, DM (2004). Estimating VS (30) (or NEHRP Site Classes) from shallow
velocity models (depths < 30m), Bull. Seismo. Soc. Am., 94(2):591–597.
http://parkseismic.com/SSC-Main.html.

Bolt, B. A. (1976). Nuclear Explosions and Earthquakes: The Parted Vail. San Francisco,
CA: W. H. Freeman and Co.
145

British Geological Survey (c) NERC 2013.

Buettner, M., Ramirez, A. and Daily, W. “Electrical Resistance Tomography for Imaging
the Spatial Distribution of Moisture in Pavement Sections” Structural Materials
Technology: An NDT Conference, San Diego, Feb. 20-23 1996 [abstract], 1996,
342-347.

Bullen, K.E., 1963, an Introduction to the Theory of Seismology: Cambridge University


Press, 381 p.

Cai, Y., Sangghaleh, A., & Pan, E. (April 01, 2015). Effect of anisotropic base/interlayer
on the mechanistic responses of layered pavements. Computers and Geotechnics,
65, 2, 250-257.

Chambers, J. E., Wilkinson, P. B., Penn, S., Meldrum, P. I., Kuras, O., Loke, M. H., &
Gunn, D. A. (June 01, 2013). River terrace sand and gravel deposit reserve
estimation using three-dimensional electrical resistivity tomography for bedrock
surface detection. Journal of Applied Geophysics, 93, 1, 25-32.

Choon B. Park, Richard D. Miller, and Hidetoshi Miura (2002). Optimum Field
Parameters of an MASW Survey.

Choon B. Park, Richard D. Miller, Jianghai Xia, and Julian Ivanov (2007) “Multichannel
Analysis of Surface Waves (MASW)-Active and Passive Methods”. The Leading
Edge.

Coskun, N. “The effectiveness of electrical resistivity imaging in sinkhole


investigations,” International Journal of Physical Sciences, 7(15), 2398 - 2405.

Coots, T. “Greene County Comprehensive Plan: Greene County Planning and Zoning,”
http://www.greenecountymo.org/web/Public_Information/files/Comp.pdf, 2007.

Choon B. Park, Richard D. Miller, and Hidetoshi Miura (2002) “Optimum Field
Parameters of an MASW Survey”.

Cornou, C., M. Ohrnberger, D. M. Boore, K. Kudo, and P.-Y. Bard, 2006, Derivation of
structural models from ambient vibration array recordings: Results from an
international blind test: Proceedings of the 3rd International Symposium on the
Effects of Surface Geology on Seismic Motion, paper NBT.

Daniels F. and Alberty R.A., 1966. Physical Chemistry. John Wiley and Sons, Inc. de
Groot-Hedlin, C. and Constable, S., 1990. Occam's inversion to generate smooth,
two dimensional models form magneto telluric data. Geophysics, 55, 1613-1624.
146

Dal Moro, G., Pipan, M., Forte, E., & Finetti, I. (2003). Determination of
Rayleigh wave dispersion curves for near surface applications in unconsolidated
sediments. In SEG International Exposition and Seventy-Third Annual
Meeting, 24-31 October 2003, Dallas, Texas (Vol. 22, pp. 1247–1250).

Dey A. and Morrison H.F. Resistivity modeling for arbitrary shaped two-dimensional
structures. Geophysical Prospecting 27, 1979, 106-136.

Denizman, C., 2003, Morphometric and spatial distribution parameters of karstic


depressions, lower Suwannee River basin, Florida: Journal of Cave and Karst
Studies, 65, 29-35.

Donohue, S., Forristal, D., & Donohue, L. A. (2013). Detection of soil compaction using
seismic surface waves. Soil and Tillage Research, 128, 54–60.

Doolittle, J.A., and M.E. Collins, 1998, Comparison of EM induction and GPR methods
in areas of karst: Geoderma, 85, 83-102.

Duffy, B.G., 2008, Development of multichannel analysis of surface waves (MASW) for
characterizing the internal structure of active fault zones as a predictive method of
identifying to distribution of ground deformation, M.Sc. Thesis, University of
Canterbury.

Elkrry, A., Nwokebuihe, S., Torgashov, E., Dera, A., Alotaibi, A., Anderson, N. (2015).
Site Specific Pavement Condition Assessment. EEGS SAGEEP 28 (Environmental
and Engineering Geophysical Society, the Symposium on the Application of
Geophysics to Engineering and Environmental Problems). Austin, TX. March,
2015.

Everett, M. E. (2013). Near-Surface Applied Geophysics. Cambridge: Cambridge


University Press.

Fellow, L. D., “Geology of Galloway Quadrangle Greene County Missouri,” Missouri


Geological Survey and Water Resources, 1970.

Forbriger, T., 2003a, Inversion of shallow-seismic wavefields: I. Wavefield


transformation: Geophys. J. Int., 153, 720-734.

Foti, S. (2000). Multistation methods for geotechnical characterisation using surface


waves. Ph.D. Thesis, Politecnico di Torino. 229p.

Gazetas, G. 1982. Vibrational characteristics of soil deposits with variable velocity. Int.
Jour. Numerical and Analytical Methods in Geomech., Vol 6, pp 1-20.

Gibson, Paul J, and Dorothy M George. Environmental Applications of Geophysical


Surveying Techniques. Hauppauge: Nova Science Publishers, 2003.
147

Gucunski, N., Woods, R.D., 1991. Instrumentation for SASW testing. In: Bhatia, S.K.,
Blaney, G.W. (Eds.), Recent Advances in Instrumentation, Data Acquisition and
Testing in Soil Dynamics, Geotechnical Special Publication, vol. 29. American
Society of Civil Engineers, pp. 1– 16.

Heisey, J. S., Stokoe II, K. H., Hudson, W. R., and Meyer, A. H. (1982), Determination
of in situ Shearwave Velocity from Spectral Analysis of Surface Waves, Research
Report No. 256-2, Center for Transportation Research. The University of Texas at
Austin, December, 277 pp.

Ivanov, J., Park, C. B., et al. (2001). Modal separation before dispersion curve extraction
by MASW method. SAGEEP Denver, Colorado.

Ivanov, J., Park, C. B., et al. (2005). "Analyzing and filtering surface wave energy by
muting shot gathers." Journal of Environmental and Engineering Geophysics 10:
307-322.

Ivanov, J., R.D. Miller, P. Lacombe, C.D. Johnson, and J.W. Jr. Lane, 2006, Delineating
a shallow fault zone and dipping bedrock strata using multichannel analysis of
surface waves with a land streamer, Geophysics, 71, 5, p. A39-A42. doi:
10.1190/1.2227521.

Ivanov, Julian, Choon Park, and Jianghai Xia. "MASW/Surfseis 2 Workshop." Fort
Worth: Kansas Geologic Survey, March 28, 2009.

Ivanov, J., R.D. Miller, S. Peterie, C. Zeng., J. Xia, and T. Schwenk, 2011, Multi-channel
analysis of surface MASW (MASW) of models with high shear-wave velocity
contrast, 2011 SEG Annual Meeting, September 18 - 23, 2011 , San Antonio,
Texas.

Jones, R.B. 1958 In-situ measurement of the dynamic properties of soil by vibration
methods. Geotechnique, vol. 8 (1), pp. 1-21.

Kearey, Philip, Micheal Brooks, and Ian Hill. An Introduction to Geophysical


Exploration. Williston: Blackwell Publishing, 2002.

Kansas Geological Survey “Multichannel Analysis of Surface Waves”, 2013


http:/www.masw.com/index.html

Keller G.V. and Frischknecht F.C., 1966. Electrical methods in geophysical prospecting.
Pergamon Press Inc., Oxford.

Kramer, S. L. (1996). Geotechnical Earthquake Engineering. Upper Saddle River, NJ:


Prentice-Hall.
148

Lin, C.-P., Chang, C.-C., & Chang, T.-S. (2004), the use of MASW method in the
assessment of soil liquefaction potential. Soil Dynamics and Earthquake
Engineering, 24 (9–10), 689–698.

Loke, M.H., “Electrical imaging surveys for environmental and engineering studies,” in
http://www.abem.se/support/downloads/case-studies/practical-guide-to-2d-
3dsurveys. 1999.

Loke, M.H. "Electrical Imaging Survey for Environmental and Engineering Studies."
2000.

Loke, M.H., 2011. Electrical resistivity surveys and data interpretation, in Gupta, H (ed.),
Solid Earth Geophysics Encyclopaedia (2nd Edition) “Electrical &
Electromagnetic” Springer-Verlag, 276-283.

Macfarlane, P.A., Healey, J.M., and Wilson, B.W., 2005 THE SOUTHEAST KANSAS
OZARK AQUIFER WATER SUPPLY PROGRAM PHASE 1 PROJECT
RESULTS, Kansas Geological Survey Open File Report 2005-15 15p.
(http://www.kgs.ku.edu/Hydro/Publications/2005/OFR05_15/aquifer.pdf. (Date
accessed: May 13, 2008).

Miller, R.D., Xia, J., and Park, C.B., 1999, Estimation of near-surface shear-wave
velocity by inversion of Rayleigh waves: Geophysics, v. 64, no. 3, p. 691-700.
(XIA-99-04)

Miller, R., Xia, J., Park, C., Davis, J., Shefchik, W., and Moore, L., 1999a, Seismic
techniques to delineate dissolution features in the upper 1000 ft at a power plant;
Technical Program with biographies, SEG, 69th Annual Meeting, Houston, Texas,
492-495.

Miller, R., Xia, J., Park, C., Ivanov, J., and Williams, E., 1999b, Using MASW to map
bedrock in Olathe, Kansas; Technical Program with biographies, SEG, 69th
Annual Meeting, Houston, Texas, 874-877.

Miller, R.D., J. Xia, and C.B. Park, 2005, Seismic techniques to detect dissolution
features (karst) at a proposed power-plant site, in, D.K. Butler, ed., near surface
geophysics: Society of Exploration Geophysicists, 663-679.

Milson, John. Field Geophysics. New York: John Wiley & Sons, 1996.

Missouri Department of Conservation 2005. Wildlife Code of Missouri, Rules of the


Conservation Commission. Issued March 1, 2005. Missouri Department of
Conservation, Jefferson City, MO.

Monroe, Watson H. A Glossary Karst Terminology. Geological Survey Water Paper


1899-K, 1970.
149

Muchaidze, I., “Imaging Karst Terrain Using Electrical Resistivity Tomography.” In


https://mospace.umsystem.edu/xmlui/bitstream/handle/10355/26344/Muchaidze_
2008. Pdf? Sequence =1. 2008.

Muchaidze, I., 2009, Imaging in Karst Terrain Using Electrical Resistivity Tomography.
Master’s Thesis, Missouri Science and Technology University, 2009.

Myat, M., Wamweya, A., Kovin, O., Anderson, N., and Robison, J., 2008a, Application
of Electrical Resistivity Method in Steeply Dipping Karst Terrain: 11th
Multidisciplinary Conference on Sinkholes and the Engineering and
Environmental Impacts of Karst: Integrating Science and Engineering to Solve
Karst Problems, September 22-26, 2008, Ramada Conference Center Tallahassee,
Tallahassee, FL. 224.

Myat, M., Muchaidze, Y., Wamweya, A., Anderson, N., and Robison, J., 2008b,
Assessment of Karst Activity at Springfield Route 60 Study Site: 11th
Multidisciplinary Conference on Sinkholes and the Engineering and
Environmental Impacts of Karst: Integrating Science and Engineering to Solve

Nakano, M., Fuji, K., and Takeuchi, S. (1988), Rayleigh wave scatteringat wedge corners
with major wedge angles. In Scattering and Attenuation of Seismic Waves, Part I
(eds. Aki, K. and Wu R. S.), Pure Appl. Geophys. 128, 119–132.

Neidell, N.S., and Taner, M.T., 1971, Semblance and other coherency measures for
multichannel data: Geophysics, 36, 482–497.

Odum, J.K., R.A. Williams, W.J. Stephenson, D.M. Worley, C. von Hillebrandt-
Andrade, E. Asencio, H. Irizarry, and A. Cameron, 2007, Near-surface shear wave
velocity versus depth profiles, Vs 30, and NEHRP classifications for 27 sites in
Puerto Rico: USGS Open-File Report 2007-1174.

Okada, H. 2003. The microtremor survey method. Geophysical Monograph Series. No.
12. Society of Exploration Geophysicists, Tulsa, Oklahoma, USA.

O’Neill, A., 2003, Full-waveform reflectivity for modeling, inversion and appraisal of
seismic surface wave dispersion in shallow site investigations, Ph.D. thesis, The
University of Western Australia, Australia.

Park, C.B., Miller, R.D., and Xia, J., 1997, Multi-channel analysis of surface waves
(MASW) “A summary report of technical aspects, experimental results, and
perspective”: Kansas Geological Survey Open-file Report 97-10.

Park, C. B., R. D. Miller, and J. Xia, 1998, imaging dispersion curves of surface waves
on multi-channel record: 68th Annual International Meeting, SEG, Expanded
Abstracts, 1377–1380.
150

Park, C.B., XIA, J. and Miller, R.D. (1998b) Imaging dispersion curves of surface waves
on multichannel record: 68th Ann.

Park, C.B., R.D. Miller, and J. Xia, 1999, Multichannel Analysis of Surface Waves:
Geophysics, 64, 800-808.

Park, C.B., Miller, R.D., and Xia, J., 1999a, Multi-channel analysis of surface waves:
Geophysics, 64(3), 800-808.

Park, C.B., Miller, R.D., and Xia, J., 1999b, Detection of near-surface voids using surface
wave: Proceedings of the Symposium on the Application of Geophysics to
Engineering and Environmental Problems (SAGEEP 99), Oakland, CA, March 14-
8, 281-286.

Park, C.B., R.D. Miller, and J. Xia, 2000, Detection of higher mode surface waves over
unconsolidated sediments by the MASW method: Proceedings of the Symposium
on the Application of Geophysics to Engineering and Environmental Problems
(SAGEEP 2000), Arlington, Va., February 20-24, p. 1-9. (PAR-00-01).

Park, Choon B, Richard D Miller, Jianghai Xia, and Julian Ivanov. "Multichannel
Seismic Surface-Wave Methods for Geotechnical Applications." First International
Conference on the Application of Geophysical Methodologies to Transportation
Facilities and Infrastructure. St. Louis, 2000.

Park, C. B., Miller, R. D., & Xia, J. (2001). Offset and resolution of dispersion curve in
multichannel analysis of surface waves (MASW). In Proceedings of the
Symposium on the Application of Geophysics to Engineering and Environmental
Problems 2001 (p. SSM4). Denver, CO: Environment and Engineering
Geophysical Society.

Park, C.B., Miller, R.D., and Miura, H., 2002, Optimum field parameters of an MASW
survey [Exp. Abs.]: SEG-J, Tokyo, May 22-23, 2002. (PAR-02-03).

Park, Choon B. MASW - Horizontal Resolution in 2D Shear-Velocity (Vs) Mapping.


Open-File Report, Lawrence: Kansas Geologic Survey, 2005.

Park, C. B., Miller, R. D., Ryden, N., Xia, J., and Ivanov, J., 2005, Combined use of
active and passive surface waves: Journal of Engineering and Environmental
Geophysics (JEEG), 10, (3), 323-334.

Park, C. B., Miller, R. D., Xia, J., & Ivanov, J. (2007). Multichannel analysis of surface
waves (MASW) active and passive methods. The Leading Edge, 26 (1), 60–64.
151

Park, C. B., & Shawver, J. B. (2009). MASW Survey Using Multiple Sources Offsets. In
D. K. Butler (Ed.), Proceedings of the Symposium on the Application of
Geophysics to Engineering and Environmental Problems 2009 (pp. 15–19).
Denver, CO: Environment and Engineering Geophysical Society.

Park, C. B., & Carnevale, M. (2010). Optimum MASW survey—Revisit after a Decade
of Use. In D. O. Fratta, A. J. Puppala, & B. Muhunthan (Eds.), GeoFlorida 2010:
Advances in Analysis, Modeling & Design (pp. 1303–1312). Reston, VA:
American Society of Civil Engineers.

Park, C., 2013, MASW for geotechnical site investigation: The Leading Edge, June 2013,
v. 32, p. 656-662.

Park, C. B. (2015). Data acquisition. Retrieved 23 June, 2015, from


http://www.masw.com/DataAcquisition.html.

Press, F., 1966, Seismic velocities: in Handbook of physical constants, revised edition,
Clark, S.P., Jr. (ed.), Geological Society of America Memoir, 97, 97–173.

Richart, F. E., Hall, J. R., and Woods, R. D., 1970, Vibrations of soils and foundations:
Prentice-Hall, Inc.

Rix, G.J., and Leipski, A.E., 1991, Accuracy and resolution of surface wave inversion:
Recent Advances in Instrumentation, Data Acquisition and Testing in Soil
Dynamics: Geotechnical Special Publication No. 29, ASCE, 17-23.

Robinson, E.S. & Çoruh, C. (1988) Basic Exploration Geophysics. Wiley, New York.

Robison, J. and Anderson, N., Geophysical Investigation of the Delaware Avenue


Sinkhole, Nixa, Missouri: 11th Multidisciplinary Conference on Sinkholes and the
Engineering and Environmental Impacts of Karst: Integrating Science and
Engineering to Solve Karst Problems, September 22-26, 2008, Ramada
Conference Center Tallahassee, Tallahassee, FL, 2008.

Rosenblad, B. L. and Cheng‐ Hsuan, L. (2011). Influence of Poisson's Ratio on Surface


Wave Near‐ Field Effects. Geo‐ Frontiers 2011: Advances in Geotechnical
Engineering, ASCE.

Ryden, N. and Mooney, M. (2009). Analysis of surface waves from the light weight
deflectometer: Soil Dynamics and Earthquake Engineering, 29: 1134-1142.

Rene P, Maros N, Ivan D, David K (2012) “Determination of Cavities Using Electrical


Resistivity Tomography”. Contributions to Geophysics and Geodesy, 42:201-211.

Socco, L.V., and Strobia, C., 2004, Surface-wave method for near-surface
characterization: a tutorial: Near Surface Geophysics, 2, 165-185.
152

Socco, L. V., Foti, S., & Boiero, D. (2010). Surface-wave analysis for building near-
surface velocity models - Established approaches and new perspectives.
Geophysics, 75 (5), 75A83–75A102.

Stokoe II KH, Wright GW, James AB, Jose MR. Characterization of geotechnical sites
by SASW method. In: Woods RD, Ed. Geophysical characterization of sites, A. A.
Balkema/Rotterdam, 1994. p. 15-25.

Shade, K.A., 2003, Temporal analysis of floodplain deposition using urban pollution 153
stratigraphy, Wilson Creek, SW Missouri. Masters of Science Thesis, Department
of Geography, Geology, and Planning, Southwest Missouri State University,
Springfield, Missouri.

Silvester, P.P. and Ferrari R.L. Finite elements for electrical engmeers. 2nd. ed.
Cambridge University Press, 1990.

Sudha, K., M., Mittal, S., & Rai, J. (January 01, 2009). Soil characterization using
electrical resistivity tomography and geotechnical investigations. Journal of
Applied Geophysics, 67, 1, 74-79.

Sheriff, R. E., & Geldart, L. P. (1995): Exploration Seismology, Cambridge University


Press.

Shishay T. Kidanu, Evgeniy V. Torgashov, Aleksandra V. Varnavina, Neil L. Anderson


(May 23, 2016). ERT-based Investigation of a Sinkhole in Greene County,
Missouri. Journal of Geosciences, 2 (2): 99-115.

Schrott, L. and O. Sass, 2008, Application of field geophysics in geomorphology:


Advances and limitations exemplified by case studies: Geomorphology, 93, 55-73.

Shuang X, Lung, and Jianghal (2004). The Selection of Field Acquisition Parameters for
Dispersion Images from Multichannel Surface Wave Data: Pure and Applied
Geophysics, 161 (2004) 185–201.

Steeples, Don. "Near-Surface Seismology: A Short Course." Lawrence: Society of


Exploration Geophysicist, 1998.

Steeples, D. W., Baker, G. S., Schmeissner, C., 1999. Toward the Autojuggie: Planting
72 Geophones in 2 Sec. Geophysical Research Letters, 26(8): 1085–1088.

Surf-Seis. Multi-Channel Analysis of Surface Waves. 2006. Retrieved February 3, 2010


from: http://www.kgs.ku.edu/software/surfseis/masw.html. Steeples, Don. "Near-
Surface Seismology: A Short Course." Lawrence: Society of Exploration
Geophysicist, 1998.
153

Telford, W.M., Geldart, L.P., and Sheriff, R.E., 1990, Applied Geophysics, second
edition: Cambridge, Cambridge University Press, 770 p.

Tian, G. (April 01, 2003). Multichannel analysis of surface wave method with the
autojuggie. Soil Dynamics and Earthquake Engineering, 23, 3, 61-65.

Torgashov, E., Anderson, N., Li, M., Ismail, A., & Elkrry, A. (2012). Imaging in Karst
Terrain Using Electrical Resistivity Tomography and Subsurface Wave Method;
EEGS SAGEEP 25 (Environmental and Engineering Geophysical Society, the
Symposium on the Application of Geophysics to Engineering and Environmental
Problems. Tucson, AZ, 2012.

Tseng, Y. H., Kang, S. C., Chang, J. R., & Lee, C. H. (December 01, 2011). Strategies for
autonomous robots to inspect pavement distresses. Automation in Construction,
20, 8, 1156-1172.

Thierry, P., N. Debeblia, and A. Bitri, 2005, Geophysical and Geological characterization
of karst hazards in urban environments: application to Orleans (France): Bulletin
of Engineering Geology and the Environment, 64, 139-150.

Unklesbay, A.G. and Jerry D. Vineyard. Missouri Geology: Three Billion Years of
Volcanoes, Seas, Sediments, and Erosion. University of Missouri Press Columbia
and London, 1992.

United States Corps of Engineers. "EM 1110-1-1802: Geophysical Exploration for


Engineering and Environmental Investigations." Engineering Manual, Washington
D.C., 1995.

United States Corps of Engineers. "Geotechnical Investigations." Engineering Manual


No. 1110-1-1804, Washington D.C., 2001.

Vandike, James E., 1993, Groundwater Level Data for Missouri, Water Year 1991-
1992: Missouri Division of Geology and Land Survey, Water Resources Report No. 42,
96 p.

Vandike, James. Karst in Missouri, an Overview. Part 1. Missouri Department of Natural


Resources, 1997.

Veni, G., DuChene, H., Crawford, N.C., Groves, C.G., Huppert, G.H., Kastning, E.H.,
Olson, R. & Wheeler, B.J., 2001, Living with karst: a fragile foundation.
Environmental awareness series, American Geological Institute, 64 pp. + 1 pl.

Veni, George. Revising the karst map of the United States. Journal of Cave and Karst
Studies 64(1), 45-50, 2001.
154

Wadati, K. (1933). On the Travel Time of Earthquake Waves, Part II. Geophysical
Magazine, 7, 101–111.

What is Karst? http://water.usgs.gov/ogw/karst/pages/whatiskarst, 2012.

Xia, J., R.D. Miller, and C.B. Park, 1998, Construction of vertical seismic section of
near-surface shear wave velocity from ground roll [Exp. Abs.]: Soc. Explor.
Geophys./AEGE/CPS, Beijing, 29-33.

Xia, J., Miller, R.D., and Park, C.B., 1999, Estimation of near-surface shear-wave
velocity by inversion of Rayleigh wave: Geophysics, 64, 691-700.

Xia, J., Miller, R.D., and Park, C.B., 2000a, Advantage of calculating shear-wave
velocity from surface waves with higher modes: Technical Program with
Biographies, SEG, 70th Annual Meeting, Calgary, Canada, 1295-1298.

Xia, J., Miller, R.D., Park, C.B., Hunter, J.A., and Harris, J.B., 2000b, Comparing shear-
wave velocity profiles from MASW technique with borehole measurements in
unconsolidated sediments of the Fraser River Delta: Journal of Environmental and
Engineering Geophysics, v. 5, no. 3, 1-13.

Xia, J., Miller, R. D., Park, C. B., Hunter, J. A., Harris, J. B., & Ivanov, J. (2002).
Comparing shear-wave velocity profiles inverted from multichannel surface wave
with borehole measurements. Soil Dynamics and Earthquake Engineering, 22 (3),
181–190.

Xia, J., Miller, R. D., Park, C. B., & Tian, G. (2003). Inversion of high frequency surface
waves with fundamental and higher modes. Journal of Applied Geophysics, 52 (1),
45–57.

Xia, J., Chen, C., Li, P.H., and Lewis, M.J., 2004a, Delineation of a collapse feature in a
noisy environment using a multichannel surface wave technique: Geotechnique,
54(1), 17-27.

Xia, J. (2014). Estimation of near-surface shear-wave velocities and quality factors using
multichannel analysis of surface-wave methods. Journal of Applied Geophysics,
103, 140–151.

Xu, Y., J. Xia, and R. D. Miller, 2006, Quantitative estimation of minimum offset for
multichannel surface-wave survey with actively exciting source: Journal of
Applied Geophysics, 59, 117–125.

Yuan, D., & Nazarian, S. (1993). Automated Surface Wave Method: Inversion
Technique. Journal of Geotechnical Engineering, 119 (7), 1112–1126.
155

Yilmaz, Ö. (1987). Seismic Data Processing. Tulsa, OK: Society of Exploration


Geophysicists.

Yoon, S. and Rix, G.J. (2009). Near-field effects on array-based surface wave methods
with active sources: J. Geotech. Geoenviron. Eng., 135(3), 399-406.

Yixian Xu a, Jianghai Xia b, Richard D. Miller (2006). Quantitative estimation of


minimum offset for multichannel surface-wave survey with actively exciting
source. Journal of Applied Geophysics 59 (2006) 117–125.

Zeng, C., Xia, J., Miller, R. D., Tsoflias, G. P., & Wang, Z. (2012). Numerical
investigation of MASW applications in presence of surface topography. Journal of
Applied Geophysics, 84, 52–60.

Zhang, S. X., & Chan, L. S. (2003). Possible effects of misidentified mode number on
Rayleigh wave inversion. Journal of Applied Geophysics, 53 (1), 17–29.

Zhang, S.X., Chan, L.S., and Xia, J. (2004). The selection of field acquisition parameters
for dispersion images from multichannel surface wave data: Pure and Applied
Geophysics, 161: 1–17.

Zhou W, Beck BF, Stephenson JB (1999), “Investigation of ground water flow in karst
areas using component separation natural potential measurements”. Journal of
Environmental Geology, 37: 19-25.
156

VITA

Ghassan Alsulaimani was born in 1976, in Taif city, Saudi Arabia. Mr.

Alsulaimani began his collegiate studies in 1994 and received a Bachelor of Science in

Geological Engineering in 1998 from King Abdul-Aziz University (KAU) in Saudi

Arabia. He has worked for the Saudi Geological Survey since 1999 as an Engineer

Geologist. Meanwhile, he was involved in some projects related in the exploration of

industrial minerals and rocks. He also worked in the geohazard department for more than

four years on a project involving the stability analysis of rock slopes and rock falls.

In May 2014 he received his Master of Science from the Geological Engineering

from Missouri University of Science and Technology. Mr. Alsulaimani started the Ph.D.

program in Geological Engineering with an emphasis in Geophysics at Missouri

University of Science and Technology during the fall semester of 2014.

During his Ph.D. program, Mr. Alsulaimani worked as a research assistant. He

has been involved in several subsurface shallow geophysical projects. The geophysical

methods he has used include resistivity, seismic, electromagnetic, GPR, and metal

detector. Ghassan is interested in shallow-applied geophysics. He has presented at a

number of conferences as a primary author and as a co-author, and he has published two

scientific papers. He has also been involved in professional societies such as SEG, GSA,

AEG, EEGS and SCE. Mr. Alsulaimani has held several elected positions in student

organizations at Missouri University of Science and Technology, including vice president

of the Saudi Student Association (SSA) in 2013-2016, and ICRM in 2014.

In May 2017 Mr. Alsulaimani received his Ph.D. in Geological Engineering from

Missouri University of Science and Technology.

You might also like