You are on page 1of 233

Field Geophysics

A Bird-dog’s View

Richard L. Kirkpatrick
Contents

Preface vii

Introduction 1

1 Elastic waves and constants 3


1.1 Seismic velocities 5
1.2 Time averaging 6
1.3 Velocities in rocks 6
1.4 Reflection and refraction 7
1.5 Diffraction 9
1.6 Dispersion of seismic pulse 10
1.7 Attenuation 10
1.8 Time-distance graph 11

2 Reflection theory 13
2.1 The reflection hyperbola 13
2.2 Normal incidence time 14
2.3 Normal moveout (NMO) 15
2.4 Stacking (NMO) velocities 16
2.5 CMP stacking 18
2.6 RMS velocity 19
2.7 Average and interval velocities 21
2.8 Static correction 22
2.9 Acoustic impedance and coefficients of reflection and refraction 26
2.10 Normal incidence time section 27
2.11 Effect of dip 27
2.12 Diffraction hyperbola and Fresnel zones 28
2.13 Convolution 31

3 Seismic refraction 33
3.1 Scope 33
3.2 Spread design 33
3.3 Principal refractors 34
3.4 Critical refraction and the Snell ratio 34
3.5 Field records 36
3.6 Picking first arrivals 36
3.7 Time-distance plots 38
3.8 The refraction equation 39
3.9 Snell right triangle and crossover distance 41
3.10 Delay times and multiple-level refractions 42
3.11 Non-parallel interfaces 43
3.12 Complex refraction survey 45
3.13 The ‘mean-minus-T’ method 46

i
4 Prelude to well logging: radioactivity, electrical & electromagnetic methods 48
4.1 Introduction 48
4.2 The science of well logging 48
4.3 Radioactivity 49
4.3.1 Types of radiation 49
4.3.2 Alpha and beta particles 50
4.3.3 Gamma rays 50
4.3.4 Law of radioactivity and half-life 50
4.3.5 Radioactive decay series 51
4.3.6 Radioactive equilibrium 54
4.3.7 Natural gamma ray spectrum 54
4.3.8 Absorption of α, β and γ rays 55
4.3.9 Radioactivity of rocks 55
4.3.10 Detection of radiation and measuring units 56
4.3.11 Stripping ratios, intensity factors and absorption coefficients 57
4.3.12 Radon measurement and α particle monitors 58
4.3.13 Radioactive density measurements 59

4.4 Electrical methods 60


4.4.1 Resistivity 60
4.4.1.1 Earth resistivity 60
4.4.1.2 Ohm’s law 60
4.4.1.3 Resistivity and conductivity 61
4.4.1.4 Point and buried electrodes 62
4.4.1.5 Resistivity of rocks and minerals 64
4.4.1.6 Archie’s law 65
4.4.1.7 Apparent resistivity 65
4.4.1.8 Electrode arrays 66
4.4.1.9 Electrical sounding and mapping 70
4.4.1.10 VES field procedure 71

4.4.2 Self-potential 74
4.4.2.1 Occurrence and origins of self-potentials 74
4.4.2.2 Field procedures 75
4.4.2.3 Electrochemical mechanisms 76
4.4.2.4 Static or equilibrium potential 80

4.5 Induced polarization 81


4.5.1 Introduction 81
4.5.2 Origin of IP 81
4.5.3 Measurement of IP in time domain 82
4.5.3.1 Apparent polarizability 82
4.5.3.2 Apparent chargeability 84

4.5.4 Measurement of IP in frequency domain 85


4.5.4.1 Frequency effect 85
4.5.4.2 Metal factor 86
4.5.4.3 Phase shift 86

4.5.5 Electromagnetic coupling 87


4.5.6 Negative induced polarization 88

4.6 Electromagnetic methods 88


4.6.1 Introduction 88
4.6.2 Principles of magnetic methods 89
4.6.2.1 Magnetic field strength, flux density and permeability 89

ii
4.6.2.2 Relative permeability, susceptibility and magnetization 90
4.6.2.3 Magnetic moment, magnetic dipole, and remanence 91
4.6.2.4 Hysteresis 91
4.6.2.5 Induced magnetism 92
4.6.2.6 Susceptibility of rocks 93
4.6.2.7 Remanent magnetism 94

4.6.3 Electromagnetic induction 94


4.6.3.1 Law of induction 94
4.6.3.2 Phase 95
4.6.3.3 Single-turn coil 97

4.6.4 Elliptical polarization 99


4.6.5 Response functions 100
4.6.6 Two-coil fixed-source low-frequency systems in free space 100
4.6.6.1 Biot-Savat law 100
4.6.6.2 Horizontal cable on the ground 101
4.6.6.3 Horizontal rectangular loop 102
4.6.6.4 Field of a current dipole 102

4.6.7 Depth penetration 103


4.6.8 Near-field artificial source continuous-wave methods 105
4.6.8.1 Tilt-angle methods 106
4.6.8.2 Fixed-source systems exploiting amplitude and phase 108
4.6.8.3 Moving source-receiver systems 109

4.6.9 Transient-field methods (time-domain electromagnetism) 115


4.6.9.1 Introduction 115
4.6.9.2 Comparison of transient field to induced polarization 116
4.6.9.3 Transient field of a single-turn circular loop 116
4.6.9.4 TEM survey procedures 117
4.6.9.5 Comparison of TEM and CW methods 118

4.6.10 VLF (far-field) methods 118


4.6.10.1 VLF radio transmitters 119
4.6.10.2 VLF transmitters and the distant field in air 119
4.6.10.3 VLF field effects 120
4.6.10.4 VLF wave in the ground 122
4.6.10.5 Measurements with VLF station in strike direction 122
4.6.10.6 Measurements with station bearing perpendicular to strike 124
4.6.10.7 Comparison of VLF and conventional EM methods 125

4.6.11 Ground penetrating radar (GPR) 125


4.6.11.1 GPR fundamentals 125
4.6.11.2 Radar parameters 128
4.6.11.3 Reflection and transmission coefficients 130
4.6.11.4 Application of the GPR method 131
4.6.11.5 Distance determination 133
4.6.11.6 Migration 135

5 Well logging 136


5.1 Introduction 136
5.2 Logging tools 136
5.3 Reservoir estimates 141
5.4 Permeable zones 142
5.5 Archie’s law 142

iii
5.6 Permeability-zone logs 143
5.6.1 Self-potential 143
5.6.2 Gamma-ray logging 144

5.7 Resistivity and conductivity logs 144


5.7.1 Resistivity logs 145
5.7.2 Induction logs 146

5.8 Porosity logs 146


5.8.1 Density log 146
5.8.2 Neutron log 147
5.8.3 Sonic or acoustic log 149
5.8.3.1 Porosity from sonic travel times 149
5.8.3.2 Synthetic seismograms and low-frequency models 150

5.8.4 Electromagnetic propagation travel-time log 151

5.9 Auxiliary logs and measurements 152


5.10 Basic log interpretation procedure 152
5.11 Other geophysical methods 153
5.11.1 Borehole magnetometer 153
5.11.2 Charged-body potential method 153
5.11.3 Logging in crystalline rocks and coal fields 153
5.11.4 Geothermal methods 154
5.11.5 Geochemical prospecting 154

5.12 Well velocity surveys 155


5.12.1 Introduction 155
5.12.2 Checkshot survey 155
5.12.2.1 Checkshot QC 156
5.12.2.2 Checkshot summary 158

5.12.3 VSP surveys 159


5.12.3.1 VSP acquisition and field plots 160
5.12.3.2 VSP quality control checks 163

Appendices 167
Appendix A Seismic noise 167
Appendix B Density - velocity 170
B.1 Rock density 170
B.2 Acoustic Impedance 170

Appendix C Logging methods 171


Appendix D Resistivity 172
D.1 Pore water resistivity 172
D.2 Determination of resistivities 173
D.3 Apparent resistivities for various dipole-dipole arrays 174

Appendix E Dipole-dipole mapping 176


Appendix F Fourier analysis 176
F.1 Time domain and frequency domain 176
F.2 Fourier analysis and filtering 177
F.3 Fourier transforms 178
F.4 Fourier transform procedures 180

Appendix G Interpretation of magnetic remanence 182

iv
Appendix H Quasi-static, near and far electromagnetic fields 183
H.1 Dielectric permittivity 183
H.2 Field of an oscillating dipole in homogeneous space 184

Appendix I Logarithms and decibels 186


I.1 Natural logarithms 186
I.2 Base 10 logarithms 187
I.3 Decibels 188

Appendix J Tables of symbols, units and definitions 190


Appendix K Well velocity surveys 197
K.1 Checkshot geometry 197
K.2 Average and interval velocities 200
K.3 VSP frequency analysis 202

Indices to Figures, Tables, Examples and Equations 204

References 213

Index 215

v
Preface

W e are called bird-dogs, field QCs, field reps, client reps, seismic acquisition specialists and
various other names, but if there is commonality in these terms, it is in what we do. Our job is
basically to ensure, through good supervision, that the geophysical data acquired in the field is of the
highest quality possible, adhering to the guidelines of our clients, governmental and environmental
agencies and of our profession. Most of us have many years of experience in the different phases of
geophysical prospecting, from acquisition, through processing to interpretation. Some of us have
PhDs in geophysics; most of us do not. I have an unwavering respect for those dedicated
geophysicists in our community who have, through many years of dedication and hard work, earned
their masters degree and especially for those who have gone on to earn their PhDs. These scientists
have built and are building the foundation of our industry. They have dedicated themselves to
research to find better solutions to geophysical problems; they have written the full spectrum of
papers on every conceivable subject; and, they have become the geophysical managers of oil and
gas companies all over the world.
I have spent the last two years writing this text, however, not for the PhDs, but for the rest of us.
Nonetheless, I do hope that even they may find some use for it maybe as a desk reference for long
forgotten equations, definitions and units of measurements found in the indices at the end of the
book. In my travels, I have been fortunate to meet many fine professionals, from jug-hustlers to party
managers, from QCs to contractor geophysicists and mud-logging TAs to well-site geologists, also
dedicated to their work but lack the higher education in applied geophysics that would have
introduced them to the theories behind the various geophysical methods. That is not to say that a
doctorate in geophysics is a requisite for ensuring that geophones are planted vertically and solidly,
that noise threshold levels are maintained, that the number of bad traces is held to an established
minimum, etc. However, I do believe that the more we understand the different facets of geophysical
methods, the better we can ultimately do our job.
Speaking of PhDs, there are a couple of them I would like to acknowledge here. That I have
worked in geophysics over a span of twenty-eight years, I owe to my brother Jon Kirkpatrick, who
arranged for my first job on a Western Geophysical field crew in Guatemala in 1976. He was, at that
time, the chief geophysicist for the client company Shenandoah Oil Co. and we were in frequent
contact as he spent a lot of his time in the field. I learned a lot from Jon and enjoyed greatly our
working relationship. Evidently, Jon received the lion’s share of our father’s (Ernest Kirkpatrick,
nuclear physicist) brains and left little for his brother. Fortunately, I did genetically inherit a
measurable quantity of common sense from our mother (Olive Ina Kirkpatrick-Seif, teacher) whose
love, wisdom and patience has guided all three of us through our lives. My sister (Ann Lee Watts,
speech therapist and audiologist) was fortunate to have received both our dad’s brains and our
mother’s common sense. Jon Kirkpatrick, an example of the above-mentioned hard-working
professionals, went on to pursue a masters degree in oceanography from Scripp’s Institute of
Oceanography and a PhD in geophysics from Colorado School of Mines.
Another is John Gibson, formally geophysical manager of Union Pacific Resources (UPR) and
now in charge of research and development for multicomponent seismic data acquisition with Veritas-
DGC. If it were not for the fact that John has been doing and continues to do amazing things in
practical geophysics, I would have thought that his true calling should be in the teaching profession.
His patience and unpretentious manner of imparting his vast knowledge and skills to someone as
fortunate as myself is unparalleled. John’s proclivity for positive reinforcement and recognition of my
work only inspired me to learn even more and to look for ways to acquire even better seismic data. If

vii
Drs. Gibson and Kirkpatrick have one thing in common, it is the fact that they both have a unique
ability to explain even the most technical of problems in a way that even I can be made to
understand.
In addition to expressing my gratitude to Jon Kirkpatrick and John Gibson, I would also like to
thank R.E. Sheriff (Encyclopedic Dictionary of Exploration Geophysics) for his incredible generosity in
naming me as a contributor to the second and third editions of his dictionary. His recognition of my
minor contribution, dealing with bandpass and time-variant filtering, has not only humbled me but has
been a constant source of inspiration. As I sit at my computer terminal day after day, typing this book
and thinking about the thousands of hours that Bob Sheriff must have spent at his computer and the
invaluable contributions he has made to geophysics, I am not only inspired but also resigned to learn
more and do more for this profession that I unabashedly enjoy.
I also want to thank Spencer Collins and Jim Callaway for giving me the opportunity to ply my
trade in Guatemala. Although extremely rewarding, the work in Guatemala was difficult with
obstacles and pitfalls encountered almost daily. Despite the geological problems (karst outcrops),
topographical problems (swamps, jungles and mountains), archaeological problems (a multitude of
Mayan sites), cultural problems, environmental problems, and more, seismic operations in Guatemala
were, I feel, successful due, in no small part, to the diligence and type ‘A’ work ethic of my boss, Jim
Callaway. The support and confidence I received from Jim further strengthened my resolve to work
hard and extract the best data possible from the geophysical contractors.
Having recognized the people who have had a great influence on me and my geophysical
endeavors and to whom I dedicate this book, I must now state that this book is not for publishing nor
will it appear in any bookstore. Even though I have spent the last two years, sometimes up to 10
hours a day, writing and rewriting, typing and drawing figures in Microsoft Word, looking up
references from my very limited library and proving scores of equations, the credit for more than half
of the content of Chapter 4 and portions of Chapters 1 and 2 of this text (applied geophysical theory)
goes to Dr. D.S. Parasnis, professor emeritus, University of Luleå, Sweden (Principles of Applied
Geophysics, 1997) and to John Milsom, University College, London (Field Geophysics, 1996).
The original concepts in writing this book were threefold: 1) keep my geophysical mind active in
between seismic contracts, 2) encapsulate the highpoints of texts by Parasnis and Milsom which
were third generation (at least) copies of the original books and hard to read in places, and 3) have a
useful field reference for future field acquisition contracts. Parasnis and Milsom have both produced
excellent texts in field geophysics. Milsom’s approach is fairly straightforward and the text easy to
read. Parasnis’ text is very well presented but highly theoretical and offers students and scholars with
advanced math background (ordinary differential equations) a unique look at geophysics. My goal
therefore, was to take the most basic facets of geophysical theories and present them in a less
mathematical bias. You will find in this book, equations and proofs involving some algebra,
trigonometry and simple calculus only. Prior to inserting formulae into their appropriate section, I first
proved the equations using the aforementioned basic math levels.
Additionally, I tried to rephrase certain references to math routines that included the words
‘clearly’ and ‘obviously’ and phrases such as ‘it is obvious that…’ and ‘it can be clearly seen that…’
because what is clear and obvious to one reader, may not be so for another. Hopefully, out of
respect for their bodies of work, I have provided adequate reference to the aforementioned authors by
inserting the symbols i and ii at the end of the sub-sections in which the wording is totally or mostly
theirs.
The majority of the font used in this text is Arial 10, which I find to be easier to read in the field.
Literally hundreds of figures, tables and examples are included to make the reading more interesting.
In addition, there are a couple hundred of the most commonly used equations in the various
geophysical methods. All of the above mentioned items are indexed in the back of the book. The
appendix contains, among other things, tables and charts for decibel conversions, SI units and
equivalences, rock properties, logging methods and a brief description of the Fourier analysis.
Finally, for those few who read this book, I sincerely welcome all comments, critiques, questions
and edits and invite you to address them to rlkirkpatrick@intelnet.net.gt.

Richard L. Kirkpatrick

viii
Introduction

T he term geophysics implies the application of the principles of physics to the study of geologic
features of the earth. Being a relatively young and somewhat limited science, many universities
offer no geophysical degree. Instead, the curriculum consists of combining courses in math, physics
and geology into a degree in geosciences. The proper blend of these three disciplines will become
more evident once experience is obtained in the actual application of geophysics in the field and in
processing centers. Although the study of geophysics includes meteorology and the physics of the
atmosphere and ionosphere, only the physics of the body of the earth is considered here. In this
narrow sense then, applied geophysics is the investigation of geologic features within the earth’s crust
such as the definition of stratigraphic changes, anticlines, synclines, salt domes, geologic faults and
rock compositions through the indexing of their physical properties such as density, porosity,
permeability and elasticity. Oceanography and vulcanology are related to applied geophysics and are
offered as in-field electives in some geosciences programs.

G eophysical methods consist of magnetic, gravity, electrical, induced polarization, electromagnetic,


radioactivity, geothermology, geochemistry, ground penetrating radar, well logging and the
seismic methods of reflection and refraction. These methods have numerous applications such as in
coal, ore, salt and precious metal mining, archaeology, construction, military defense, groundwater
surveys, near surface cavern studies, buried river bed locations, fracture zones, sand deposits and oil
and gas exploration. The primary geophysical methods applied to oil and gas are seismic reflection
and refraction, well logging and gravity/magnetics.
The work of a bird-dog (field Q.C., field supervisor, seismic acquisition specialist or client rep, etc.)
may take him anywhere in the world where there is oil and gas exploration, on land, over water, in the
ground or in the air. The number of different geophysical methods is great and the theory behind them
changes rapidly, being updated frequently through constant research. However, the scope of this text
is generally limited to the seismic method with an emphasis on land acquisition by seismic reflection
and refraction methods and well logging.

T he seismic methods of geophysical exploration utilize the fact that elastic waves travel with
different velocities in different rocks. The principle is to initiate such waves at a point and
determine at a number of other points the time of arrival of the energy that is refracted or reflected by
the discontinuities between different rock formations. This then enables the position of the
discontinuities to be deduced.
The importance of the seismic methods lies above all in the fact that their data, if properly
handled, yield a more detailed interpretation than the data of any other method.
The incomparably most extensive application of seismic methods is in oil prospecting but they are
also employed for site investigations in building large-scale structures and other civil engineering
projects like determination of depth to bedrock, delineation of sand and gravel deposits, detection of
water-bearing fracture zones, etc.ii

S eismic investigations are based on different elastic properties of the rocks of the upper crust. A
seismic wave can be created on the surface by artificial seismic sources like hammer, weight-drop,
vibrator, [air cannon, asphalt compactor] or explosive charges. The wave runs through the earth with
a velocity that depends on the traversed rocks or materials. At interfaces where the seismic velocity or
density changes, seismic waves are diffracted, refracted or reflected.

1
Portions of the primary wave return to the surface after traveling different distances through the
ground. There, the remaining seismic signals are registered by a number of seismic receivers, called
geophones. Mostly, they are arrayed on a single line, but other arrays may also be used. By
evaluating the travel time between the break and the recording of a seismic signal, the seismic
velocities of strata, their location and the depth of their seismic reflectors may be inferred.
Consequently, seismic work will provide special knowledge about the thickness and extension of
layered strata and structures of the earth, which is essential to solve geological or hydrological
problems. In addition, the shape and extension of waste deposits can be comprehended.
The propagation of seismic waves follows the geometry of optical laws. The refraction of seismic
waves occurs at the boundary between two layers of different lithology where the lower layer has a
faster velocity than the upper causing the propagating ray to bend. Seismic reflection is the echo or
“bouncing” of the seismic wave at stratigraphical or structural boundaries.iii

Abeseismic wave transmits energy by the vibration of rock particles. Low-energy seismic waves can
regarded as elastic, leaving the rock mass unchanged by their passage, but rocks close to a
seismic source may be shattered and permanently distorted. Chapter 1 deals with the origin and
definition of elastic waves and the derivation of elastic constants which are the building blocks for the
determination of velocities in reflection seismology.
Chapters 2 and 3 contain discussions of the theory of seismic reflection and refraction. Chapter 4
offers basic theory in the areas of radioactivity, electrical methods, induced polarization and
electromagnetic methods leading up to well logging in Chapter 5. Ultimately, seismic acquisition will
be discussed in detail in Chapter 6 of the second edition.

2
1
Elastic waves and constants

W hen a sound wave travels in air, the molecules oscillate backwards and forwards in the direction
of energy transport. This pressure or ‘push’ wave travels as a series of compressions and
rarefactions. It has the highest velocity of any wave motion and is therefore also known as the primary
wave or simply the P-wave.
Rock particles vibrating at right angles to the direction of energy flow create an S-wave (shear,
shake or because of the relatively slow velocity, secondary wave). The S-wave velocity, which in
many consolidated rocks is roughly half that of the P-wave, depends slightly on the plane in which the
particles vibrate, but the differences are not important in small-scale surveys.
P- and S-waves are body waves, expanding within the main rock mass. At interfaces, Love
waves can be generated and at the earth’s surface Rayleigh waves can travel, with particles
following elliptical paths. Love and Rayleigh waves may carry a considerable proportion of the source
energy, but travel very slowly and are of little geophysical use. They are often lumped together as
ground roll.i

T he basis of the seismic method is the theory of elasticity. The elastic properties of substances are
characterized by elastic moduli or constants which specify the relation between the stress and the
strain. A stress is measured as force per unit area. It is a compressive (or tensile) stress if it acts
perpendicular to the area and a shear stress if it acts parallel to it. A system of compressive stresses
changes the volume but not the shape of a body; one of shear stresses changes the shape but not the
volume.
The strains in a body are deformations which produce restoring force opposed to the stresses.
Tensile and compressive stresses give rise to longitudinal and volume strains which are measured as
the change in length per unit length or change in volume per unit volume. Shear strains are measured
as angles of deformation. It is usually assumed that the strains are small and reversible, that is, a
body resumes its original shape and size when the stresses are relieved.
Hooke’s Law states that the stress is proportional to the strain, the constant of proportionality
being known as the elastic constant which is measured in Newtons per square meter.

Table 1.1 Elastic constants

Substance Bulk Modulus (κ) Shear Modulus (µ)


(Nm-2 x 10-10) (Nm-2 x 10-10)

Marble and limestone 3.7 – 5.7 2.1 – 3.0


Granites 2.7 – 3.3 1.5 – 2.4
Ohio sandstone 1.25 0.61
Iron (wrought) 16.0 7.7
Iron (cast) 9.5 5.0
Glass (crown) 5.0 2.5
Quartz (fiber) 1.5 3.0

3
The law is not strictly true and more general stress-strain relationships have also been introduced in
applied seismology, notably by Ricker (1953); yet Hooke’s law carries us a long way in the theory of
elasticity (Appendix B.2).
The two moduli of immediate interest for the study of the elastic waves in the earth are the bulk
modulus (κ) and the shear modulus (µ). Their values for some rocks and for a few common
substances will be found in Table 1.1. The bulk modulus is also referred to as the incompressibility
and the shear modulus as the rigidity. Liquids and gases offer no resistance to shear deformation so
that for them µ = 0.

I f the stress applied to an elastic medium is released suddenly, the condition of strain propagates
within the medium as an elastic wave. There are several kinds of elastic waves:

1. In the longitudinal, compressional or P-waves, the motion of the medium is in the same
direction as the direction of wave propagation. These are, in other words, ordinary sound waves.
Their velocity is given by:

Vp = [( κ + 4/3µ ) / ρ ]1/2 (1.1) ξ

where ρ is the density of the medium.

2. In the transverse, shear or S-wave, the particles of the medium move at right angles to the
direction of wave propagation and the velocity is given by:

Vs = ( µ / ρ )1/2 (1.2) ξ

It is evident that Vp > Vs. Shear waves do not propagate through liquids and gases (Vs = 0).
Shear waves can be polarized in such a way that the particles oscillate along a definite line
perpendicular to the direction of wave propagation. In a geophysical context, these polarized
waves are known as SH and SV waves depending upon whether the particle motion is horizontal
(parallel to the ground surface) or vertical (perpendicular to the surface).

3. If a medium has a free surface, there are also surface waves in addition to the above two,
which are ‘body waves’. In the Rayleigh waves, the particles describe ellipses in the vertical
plane that contains the direction of propagation. At the surface the motion of the particles is
retrograde with respect to that of the waves. The velocity of Rayleigh waves is about 0.9 Vs.
Rayleigh waves are very commonly referred to as ground roll.

4. Another type of surface waves are the Love waves. These are observed when the S-wave
velocity in the top layer of a medium is less than in the substratum. The particles oscillate
transversely to the direction of the wave and in a plane parallel to the surface. The Love waves
are thus essentially shear waves. Their velocity for short wavelengths is equal to Vs in the upper
layer and for long wavelengths Vs in the substratum.

5. In addition to the above, there are guided waves. These are confined to a particular layer (e.g.
water layer, coal seam etc.) and their velocity depends upon the wavelength and layer thickness,
as well as on density and elastic constants.

The frequency spectrum of body waves in exploration extends from about [10 Hz to 125 Hz; the
surface waves have frequencies lower than about 10 Hz]. The low-frequency cut-off of the seismic
spectrum depends on the source, and while [8 – 10 Hz] is representative of land sources, marine air-
gun arrays can generate 8 Hz or less.
In applied seismology, mainly the P-waves are of importance as exploration tools. In principle,
however, S-waves could also be used, but the difficulty is to get S-waves of sufficient energy.
Explosions, which are the common means of generating powerful elastic waves, produce
predominantly, if not exclusively, P-waves. These are converted in part to S-waves by oblique

4
reflection. Although some work has been reported with such S-waves, they have not been put to any
great use in applied geophysics.
Surface waves are incapable of giving information about structures at depth and are of no avail in
exploration geophysics. The interest attached to them, particularly the Rayleigh waves, is due to the
fact that they constitute an unwanted signal on the seismic record which has to be suppressed to
enhance the desired signals.ii

1.1 Seismic Velocities

T he seismic velocities of rocks are the velocities at which wave motions propagate. They are
quite distinct from the continually varying velocities at which individual rock particles oscillate.
Any elastic wave velocity (V) can be expressed as the square root of an elastic modulus divided by
the square root of the density. For P-waves, the elongational elasticity, j, is appropriate; for S-waves,
it is the shear modulus, µ .

Vp = ( j / ρ ) 1/2 and Vs = ( µ / ρ )1/2 ξ

These equations suggest that the velocity should decrease with density, but because the elastic
constants normally increase rapidly with density, the reverse is usually true.
If density and the velocities are known, all the elastic constants of a rock mass can be determined
as they are related by the equations

j = q ( 1-σ ) / ( 1+σ )( 1-2σ ) (1.3a) ξ


µ = q / 2 ( 1+σ ) (1.3b) ξ
κ = q / 3 ( 1-2σ ) (1.3c) ξ

where σ is the Poisson ratio, and is always less than 0.5; q is Young’s modulus and κ is the bulk
modulus. It follows that j = κ + ( 4µ ) / 3 and that a P-wave always travels faster than an S-wave in
the same medium.

0 1000 2000 3000 4000 5000 6000 7000 ms-1

Soil
Rippable
Sand Marginal Zone
Clay Not rippable

Coal
Shale
Sandstone Quartz
Limestone/Dolomite
Salt Chalk

Anhydrite
Basalt/diabase
Gabbro
Granite
Gneiss
Other
Air Water Ice

Figure 1.1 Ranges of P-wave velocities and


rippabilities in common rocks

5
Often only P-wave velocities can be measured, providing a rather rough guide to rock quality.
Figure1.1 shows the ranges of velocity for common rocks and also their rippabilities, the extent to
which they can be ripped apart by a spike mounted on the back of a bulldozer.

1.2 Time averaging

Clays and other weathering products have low seismic velocities and high porosities. Weathering
therefore reduces rock velocities. This fact lies behind the rippability ranges of Figure 1.1. Few
fresh, consolidated rocks have velocities of less than about 2200 m/s and rippable rocks are generally
at least partly weathered.

Example 1.1

Vp (quartz) = 5200 m/s


Vp (water) = 1500 m/s

The P-wave velocity in a sandstone, 80% quartz, 20% water is given by

1 / Vp = 0.8 / 5200 + 0.2 / 1500


= 0.000287
∴ Vp = 3480 m/s

Within fairly broad limits, the velocity of a mixture can be obtained by averaging the transit times
(not the velocities) of the constituents, weighted according to their relative proportions. The principle
can even be used when one of the constituents is a fluid (Example 1.1).
Dry rocks have air (V = 330 m s-1 ) in the pore spaces. Although time-averaging cannot be used
quantitatively for gas filled pores, dry materials generally have very low P-wave velocities. If they are
poorly consolidated and do not respond elastically, they may also strongly absorb S-waves. Wet,
poorly consolidated materials have velocities a little higher than that of water, and the water table is
often a prominent seismic interface.i

1.3 Velocities in rocks

T ypical values for the velocity of P- and S-waves are given in Table 1.2. The velocities are
generally found to be greater in igneous and crystalline rocks than in sedimentary ones. In the
sedimentary rocks, they tend to increase with depth of burial and geologic age. Many empirical
attempts have been made to represent this increase. For shales and sands, Faust (1951) finds that

V = 46.5 ( zΤ )1.6 m s-1 (1.4) ξ

where z is the depth in meters and Τ the age in years.

Table 1.2 Elastic velocities in m/s

Material Compressional Shear


Air 330
Sand 300 – 800 100 – 500
Water 1450
Glacial moraine 1500 – 2700 900 – 1300
Limestone and dolomites 3500 – 6500 1800 – 3800
Rock salt 4000 – 5500 2000 – 3200
Granites & other deep-seated rocks 4600 - 7000 2500 - 4000

6
Seismic velocities can be measured in the field as well as in the laboratory on samples of rocks.
Several methods for laboratory determination using magnetostrictive pulses, ultrasonic pulses,
resonance, etc., have been developed (Hughes and Cross, 1951; Baule, 1953; Shumway, 1956; and
Datta, 1968). Well velocity surveys (sec. 5.12) [e.g. checkshot, VSP (vertical seismic profiling),
using as energy source, vibrators, explosion or air gun] are also a common method of obtaining
information of velocities. A seismometer is lowered into a borehole and coupled with the rock, a shot
is fired close to the surface and the travel time to the seismometer is noted. The seismometer is then
lifted a short distance and another travel time is measured. From the difference in these times, the
average velocity in the material between the two positions of the seismometer can be calculated.
Seismic velocities often show anisotropy in stratified media; the velocity parallel to the strata is
generally greater than that normal to them by an amount of the order of 10 – 15%. Modern research
has indicated that anisotropy due to cracks and micro-fractures is also important besides anisotropy
from fine layering.ii

Only a small part of a wavefront is of interest in a geophysical survey as only a small part of the
energy returns to the surface at points where detectors have been placed. It is convenient to
identify the important travel paths by drawing seismic rays to which the laws of geometrical optics can
be applied. Ray-path theory works less well in seismology than in optics because useful seismic
wavelengths are generally between [10 meters and 600 meters], but field interpretation can none the
less be based on ray approximations.i

Seismic ray
Wavefront

1.4 Reflection and Refraction

T he surfaces in a medium on which the wave motion has the same phase at all points are called
wavefronts. If the wavefront surface is plane, we talk of a plane wave. The normal to a
wavefront at any point is the direction of a ray and is the instantaneous direction of wave propagation
at that point. In a medium of constant seismic velocity the rays are straight lines. In inhomogeneous
media the rays are curved. It is often more convenient to describe wave propagation by means of rays
rather than wavefronts.
When seismic rays fall on the interface between two media they may be reflected or refracted. In
addition a mode conversion occurs, that is, an incident P-wave, for example, is reflected or refracted
partly as a P-wave and partly as an S-wave (Figure 1.2). If θ is the angle made by any ray with
the normal to the interface, then Snell’s law states that sinθ / V = constant where V is the velocity of

P
P

θ2
S
θ3
θ1
Vn
Vn+1
θ5
θ4

Figure 1.2 Snell’s law S

7
the wave in question. Thus, with θ1 as the angle of incidence of the P-wave ray shown, we have:

sin θ1 / Vp1 = sin θ2 / Vp1 = sin θ3 / Vs1 = sin θ4 / Vp2 = sin θ5 / Vs2 ξ

Corresponding equations exist if the incident wave is an S-wave. At normal incidence ( θ1 = 0 ) an


incident P-wave produces no S-waves and an incident S-wave produces no P-waves.ii
As shown in Figure 1.2 some of the P-wave energy is reflected when it encounters an interface
between two different rock types ( θ1 = θ2 ), and the remainder (less signal amplitude decay due to :
geometric divergence and absorption) will continue onward however, bent towards the surface
(providing that Vn+1 > Vn) i.e., it will be refracted.

T he law of reflection is very simple; the angle of reflection is equal to the angle of incidence (Figure
1.3). Refraction is governed by Snell’s law, which relates the angle of incidence and angle of
refraction to the seismic velocities in the two media

sin i / sin r = V1 / V2 (1.5) ξ

If V2 is greater that V1, refraction will be towards the interface. If sin i equals V1/V2, the ray will be
refracted parallel to the interface and some of the energy will return to the surface as a headwave,
which leaves the interface at the original angle of incidence (Figure 1.3(b)). This is the basis of the
refraction methods discussed later on. At greater angles of incidence there can be no refracted ray as
all the energy is reflected.
When drawing accurate ray paths, allowance must be made for refraction at all shallower
interfaces. Only a normal-incidence ray, which meets all interfaces at right angles, is not affected.i

S1 S2

(a) (b) Headwaves

i i
i ic ic
V1
V2 A B
r

Figure 1.3 (a) Reflection (b) Refraction

In Figure 1.3 (b), ic is referred to as the critical angle, the angle at which according to Snell’s law,
refraction is parallel to the interface and propagation is nill. Obviously this occurs when r, the angle of
refraction, reaches 90º as indicated in the refraction proportionality equation

sin i / sin r = V1 / V2 to find refraction angle r,

sin r = V2sin i / V1

r = sin-1 ( V2sin i / V1 ) (1.6) ξ

At r = 90º the refraction is parallel to the interface and since the sine of 90º = 1,

V2sin i / V1 = 1 , V2 sin i = V1 , and sin i = V1 / V2 therefore,

8
ic = sin-1 ( V1 / V2 ) (1.7) ξ

Since ic is inversely proportional to V2 and r directly proportional to it, it can be stated that as V2
increases with respect to V1, the refraction angle r increases with its limit 90º and the critical angle
decreases approaching the limit of the angle of incidence resulting in no propagation. This however is
more theoretical than practical since in actual seismic conditions the velocity boundary differences are
generally more subtle.
In the previous discussion of wavefronts and their resultant reflection and refraction it was
assumed that the wavefront encounters a plane and level velocity boundary. However, that is not
always the case as will be seen in the following section.

1.5 Diffraction

R ays are primarily useful for describing wave propagation in homogeneous media or when waves
hit an interface whose radius of curvature is large in comparison with the wavelength λ as in
Figure 1.3 where, the interface being plane, its radius of curvature is infinite. When an obstacle in the
path of a wave has a radius of curvature that is small compared with λ one must turn to the principle
advanced by Huygens for a description of wave propagation.
Huygens’ principle states that every point on a wavefront may be considered to be a secondary
source that emits waves traveling radially outward from the point. The envelope of the wavefronts of
all such spherical waves defines the position of the primary wavefront at a later instant. Only that part
of the envelope which is in the direction of the wave advance is considered in Huygens’ construction.
Reflection, refraction and Snell’s law can be derived from Huygens’ principle but more importantly it
describes correctly the situation when a wave hits a small obstacle, an edge or any similar
discontinuity in a medium.

A T1 T2 B

E t = t0 + ∆t F

D P t = t0
C R M
N
G t = t0 + ∆t H

Figure 1.4 Diffraction of an edge. CD


is a thin layer in an otherwise
homogeneous medium.

Consider, for example, a plane wavefront impinging at time t = t0 on the horizontal layer CD that
terminates abruptly at D (Figure 1.4). The envelope, at time t = t0 + ∆t, of all spherical waves
emanating from points on CD is the reflected plane wavefront EF at this time. To the right of D the
downward-traveling wavefront has reached the position GH at t = t0 + ∆t since there is no obstacle in
its path. In addition to EF and GH there is a wavefront MFPGN that emanated from D and has now
reached this position. This represents the diffracted wave traveling outwards from D and it should be
noted that it is present even in the segment GNM which is in the geometrical shadow of the primary
wave. A point T1 on a surface AB receives two waves, the reflected one RT1 and the diffracted one
DT1. At a point T2 only the diffracted wave DT2 is received.
Since for seismic waves λ is of the order of 100 meters or more, many geological features give
rise to diffracted waves.ii

9
1.6 Dispersion of seismic pulse

A nother assumption previously made is that the elastic disturbance consists of a sinusoidal wave of
a single frequency ν with no beginning or end in time. The velocity V of this wave is obtained from
the distance travelled by any surface of constant phase (i.e. maximum compression) in unit time. This
is the phase velocity of the wave.
Particle displacement

Shot Limit of Early Early Later


equivalent direct reflections reflections
cavity arrivals

Figure 1.5 Broadening of a seismic pulse

However, when a charge is detonated on land or in water a seismic pulse of a certain shape and
finite duration (breadth) is generated. Figure 1.5 shows a sequence of events after charge is fired.
Near the shot, if the shot is on ground, the maximum stress experienced by the earth material may
exceed the reversible stress-strain relationships of Hooke’s law. The distance around the shot at
which the maximum stress falls just within such reversible relationships defines an equivalent cavity,
which, rather than the shot, is often regarded as the source of the seismic pulse. According to
Fourier’s theorem, the pulse can be considered to consist of an infinite number of sinusoidal waves
of infinitesimally close-spaced frequencies from zero to infinity. If the velocity of a wave depends upon
frequency, each of these component waves travels at its own velocity, with the result that the initial
form of the pulse is distorted as it propagates in the medium. More specifically, the pulse broadens
with increasing time, that is, with the distance traveled.
Although there is no specifiable point on the pulse (peak, trough or any other) which travels with a
constant velocity, the pulse as a whole (or rather its center of energy) travels with a constant velocity U
(provided energy is not absorbed on the way). This is the group velocity of the elastic disturbance.
However, the pulse spreads out as it travels and as U is a function of the wavelength the energy
associated with different wavelengths ultimately travels at different speeds.
The phenomenon of the distortion of pulse form during propagation due to a dependence of phase
velocity on frequency, and through it on the wavelength λ since V = νλ, is called dispersion. Between
V and U exists the relation

U = V – λ ( dV / dλ ) (1.8) ξ

The dispersion of seismic body waves (P or S) is too small to be usually significant in practice. It is,
however, appreciable for surface waves.ii

1.7 Attenuation

Seismic waves are reduced in amplitude as they are propagated through the earth due to three
factors: geometric divergence, partial transmission and reflection at acoustic boundaries, and

10
absorption of energy in the medium of transmission.
The influence of geometric divergence is well known and can be allowed for, i.e., at large
distances from the source spherical waves reduce in amplitude in inverse proportion to the distance
traveled.
The transmission and reflection coefficients of a geological interface are functions of the
elastic contrast between the layers in contact. In principle, a comparison of the amplitudes of the
reflected pulses should provide some information about such contrasts, and the attenuation suffered.
It should be noted that elasticity theory based on Hooke’s law involving reversible stresses and
strains does not admit to any attenuation due to absorption of energy. The attenuation due to
geometric divergence only means that energy is spread over larger and larger areas with distance
from the source but none of it is lost. However, it is observed in practice that seismic waves do lose
energy through absorption in rocks with the amplitude A very nearly following the empirical equation A
= A0 exp(-δνx / V) where δ is the logarithmic decrement and x the distance. Typical values of δ for
earth materials in bulk would be about 0.02 to 0.03. We see from this that high-frequency waves are
attenuated at a faster rate than low-frequency ones.
Two absorption mechanisms, viscosity and solid friction, have been suggested. Both
mechanisms are observed in rocks but is seems that solid friction generally predominates. The
presence of water apparently increases the decrement and at the same time leads to a predominance
of viscous damping. Work on the absorption of seismic waves has been reported by Datta (1968),
where also references to several other papers on the subject will be found.ii

1.8 Time-distance graph

I t will be convenient now to visualize how a seismic recording might look. Figure 1.6 shows the
graph of the arrival times of the principle waves in the sample case of a uniform overburden on top
of a substratum. The arrival time of the direct P-wave and the ground roll or Rayleigh wave (GR)
increase in direct proportion to the distance and therefore lie on the respective straight lines. The
arrivals of the wave reflected at the overburden-substratum interface lie on a curve, that is to say a
hyperbola. From section 1.4, if the angle of incidence i has a critical value ic = sin-1 ( V1 / V2 ) the
refracted wave travels along the interface. From a certain distance (marked 2h tan ic) onwards the
disturbances produced by the headwave can be observed, and they arrive earlier than the reflections.
Again from a certain distance (marked xc) the headwave overtakes the direct wave. All the refracted
arrivals also lie on a straight line since the increments in their travel time are in direct proportion to the
increments in distance.

Ground roll
-1
500 m s
Time

Refracted
Reflected wave
-1
wave 5000 m s

Direct
arrival
-1
2500 m s

0,0 2h tan ic
. .
xc Distance

Figure 1.6 Basic time-distance graph

11
In plotting Figure 1.6 the S-waves were disregarded. Similarly, waves reflected repeatedly
between overburden base and ground surface are also disregarded. In an actual situation there will,
in general, be several layers and these may in addition be inhomogeneous. These and other
complications will discussed later but it is worth noting here that the time-distance graphs then
obtained, although much more involved, nevertheless show a basic similarity with the graph in Figure
1.6.ii

12
2
Reflection theory

2.1 The reflection hyperbola

T he depth to an interface between rock formations can be determined by measuring the travel time
of a seismic wave generated at the surface and reflected back from the interface. The energy of P
as well as S-waves is reflected partly as P and partly as S-waves. If the reflected and the incident
waves are of the same kind (both P or both S) the ordinary law of reflection applies, namely, angle of
incidence = angle of reflection as shown in section 1.4.
It is generally assumed that the observed reflections are only P – P reflections. This assumption
is justified as a rule since near the shot point most of the explosive energy is transmitted as P-waves.
Figure 2.1 (a) shows a simple example of a reflected ray from shot S to detector G traveling
through one layer with a uniform velocity V. Applying the Pythagorean theorem we can derive the time
it takes for the reflected wave to complete the trip from S to G having been reflected at A.

x
S G SA2 = h2 + ( x / 2 )2

SA = h2 + x2 / 4 , In terms of time-velocity:
h SAG = tV therefore tV = 2 h2 + x2 / 4 where t is
V
total travel time (two way time) from S to G
through A. The one way time from S to A is
t / 2. Therefore:

t = (2/V) h2 + x2 / 4 (2.1a) ξ
A
and, h = (1/2) V2t2 – x2 (2.2) ξ
Figure 2.1 (a) Reflection path

Either equation shows that the x versus t curve is a hyperbola


convex towards the x axis and with the t axis as the axis of
symmetry. The line through x = 0, t = 0 with slope 1 / V is its
asymptote. The time t in equation (2.1a) is referred to as the
t two-way time (Ttwo).

rewriting (2.1a), t2 = (2 / V )2 ( h2 + x2 / 4)

Asymptote = (4h2 / V2) + (x2 / V2)

t2 = (4h2 + x2) / V2 (2.1b) ξ


x

(b) One-sided hyperbola

13
2.2 Normal-incidence time

T he normal-incidence time (vertical or zero offset time) t0 is the two-way time (Ttwo) from S to the
interface and back to S along the normal to the interface. Putting h in terms of time-velocity (since
V is uniform and the interface is parallel to the surface)

h = (t0V ) / 2 and

t0 = (2h) / V (2.3) ξ

Since t0 = (2h) / V and t02 = 4h2 / V2 , the equation t2 = (4h2 + x2) / V2 can be rewritten

t2 = t02 + x2 / V2 (2.4a) ξ

or, t2 – t02 = x2 / V2 (2.4b) ξ

which is the hyperbolic equation for the two-way time with an offset x linking the travel time t and the
normal-incidence time t0.
For small offsets this equation can be replaced by the parabolic approximation form

t – t0 = x2 / 2t0V2 (2.5) ξ

which gives the normal moveout (NMO), t – t0, directly as a function of velocity, reflection time and
offset. As the velocity usually increases with depth and the travel time always increases with depth,
the NMO decreases with depth.

Direct
arrivals

Refractions

Ground roll

Air wave
Reflections

Figure 2.2 Seismic time section showing airwave, ground roll,


direct arrivals, refractions and reflection hyperbola

Curved alignments of reflection events can be seen on many multi-channel records (Figure 2.2). If
the source is at the center of the geophone spread, the curves obtained over horizontal interfaces are
symmetrical. Curvature is the most reliable way of distinguishing shallow reflections from refractions,
but for deep horizons the critical distance for refraction will usually lie beyond the furthest geophone.
Only one branch of the hyperbola is shown in Figures 1.6, 2.1 and 2.2. The other branch will
correspond of course, to the arrivals at detectors left of S. Almost without exception time-distance
graphs in reflection seismics are plotted with the time axis downward. [This stands to reason since
actual seismic prospecting is downward looking]. Using this convention and adding complementary
halves to the graphs in Figure 1.6, we obtain Figure 2.3 (a) an actual two-sided time section and

14
Figure 2.3 (b) its corresponding t – x graph. In both time sections 2.2 and 2.3 (a) the various seismic
events i.e., ground roll, direct wave, refracted wave and reflections, are readily seen.

Figure 2.3 (a) Time section showing both sides of the shot

x Geophones SP Geophones x

Time
Time

Figure 2.3 (b) t – x graph representing the actual time section

The reflection stretch in Figure 2.3 is considered to extend on either side of the source to a
maximum distance of xm = 2h tan ic . Over this stretch there are no refracted arrivals. As shown in
equation (1.7) sin ic = V1 / V2 so that

xm = 2(V1 / V2) h / [1 – (V1 / V2)2]1/2 (2.16) ξ

2.3 Normal move-out (NMO)

A lthough source receiver separations (offsets) are small compared with reflector depths they are
not (and cannot be) zero as required by the concept of vertical incidence (section 2.1). If the
hyperbolic equation (2.4a) for two-way travel time with offset x is written as

t= t02 + (x2 / V2) (2.6) ξ

where t0 =(2h / V) is the vertical or zero-offset two-way time, Ttwo, then the change in time

∆t = t - t0 (2.7) ξ

∆t = t02 + (x2 / V2) – t0 (2.8) ξ

15
is the difference between Tx, two at the detector and the zero-offset T0, two. This is known as normal
move-out or NMO (Figure 2.4).

NMO = Tx – T0 (2.9) ξ

From the definition of ∆t it is obvious that subtracting ∆t from t gives the vertical T0 two at a given
detector. If x << Vt0 from equations (2.5 and 2.7) we get the aforementioned approximate parabolic
expression

∆t ≅ x2 / 2t0V2 (2.10) ξ

while the equation (2.8), although geometrically exact, ignores waveform and noise considerations.

0 x S x R
0
T0 . t0
tx
V = stacking
Tx – T0 = NMO velocity
t0
Tx
. tx

tx ∆tnmo = tx – t0
2 2 2
= t0 + x / V - t0

Figure 2.4(a) Definition of NMO (b) Relationship of Vs to NMO

If ∆t is calculated for each sample on a trace and the sample value placed at t – ∆t, then instead of
a hyperbola as in Figure 2.3 we shall obtain a straight line through the vertex of the hyperbola,
provided V is correctly chosen. Stated quite simply, NMO corrections by virtue of a selected velocity
are used to remove the non-vertical component of ∆t.
The main problem in NMO corrections is thus to estimate V. In practice V is, in effect, chosen so
that the reflection hyperbolae are transformed into horizontal straight lines. A method of determining V
that is theoretically exact is to plot t2 against x2. This gives a straight line whose slope is 1/V2 and
whose intercept is t0 as the hyperbolic equation (2.6) t = [ t02 + (x2 / V2)]1/2 shows.
The velocity that gives a satisfactory straight-line alignment of the reflections seen on a CMP
(common mid-point) gather is called the stacking velocity. The velocity is different for different
hyperbolae on the same gather.
All NMO corrections involve some interpretation of data (e.g. sorting out diffraction hyperbolae,
multiples, overlapping events, etc.) and visual checks are always needed even if automated schemes
are used to estimate stacking velocities.
Since ∆t is a function of the record time, and varies with the Ttwo at a sample, it is also called the
dynamic correction (or simply stated a non-static correction).

2.4 Stacking (NMO) Velocities

STtacking velocity (V ) is an empirical value that yields a satisfactory NMO relationship between
nmo
and T when used in the expression (equation (2.4a))
x 0

Tx2 = T02 + (x / Vnmo)2

Figure 2.5 shows the gather of two traces whose reflections coincide at point A with source and
receiver positions as indicated. If the datum surface and reflector are both flat and if the seismic
velocity between them is constant, then the dashed lines show the raypaths for the two reflections.
T0 is the observed reflection time for the vertical raypath and Tx is the total time along the slant
path. The two travel paths and their projections on the datum surface form two right triangles, each
having sides x / 2, VT0 / 2, and VTx / 2.

16
S2
S1 x/2 R2 x/2 R1

VT0
VTx VTx
2 2

A
Figure 2.5

Raypath S2,R2 – A – S2,R2 is convenient for developing equations related to time and velocity as x = 0
however improbable in practice due to the extreme source noise and potential damage to the receiver.
From Figure 2.5 we can compute

(VTx / 2)2 = (x / 2)2 + (VT0 / 2)2

V2Tx2 = x2 + V2T02

Tx2 – T02 = x 2 / V 2

and V 2 = x 2 / (Tx2 – T02)

√(Tx (2.11) ξ
2
therefore V=x/ – T02)

recalling that Tx – T0 = NMO, (2.11) can be expressed

V = x / √ [ 2T0(NMO) + (NMO)2 ] (2.11a) ξ

In a perfect geophysical world a seismic survey designed as in Figure 2.5 could be employed and
would be much easier and more cost effective than actual modern surveys.
Actually the raw seismic record is inherently contaminated by unwanted signals besides the
reflections that are the target of the investigations. All unwanted signal is called noise. Noise is
classified into coherent (or correlated) noise and random (incoherent or uncorrelated) noise.
Coherent noise is so called because it is more or less similar (and predictable) from one detector
to another. Examples of coherent noise are multiple reflections sometimes called signal-generated
noise such as ringing and reverberation in marine work, and surface waves such as ground roll and
airwave (Appendix A).
Random noise is completely unpredictable from one detector to another. Ambient noise such as
wind and rain, static electricity, micro-tremors, and vehicle and foot traffic are examples of incoherent
noise. In addition there is induced noise such as 50 or 60 cycle high line and to a lesser degree noise
created by the recording instruments and line equipment such as, cross-talk, harmonic distortion,
analog to digital error, generator noise, to name a few (Appendix A).
To compensate for both types of noise, multiple shots are taken to specially designed geophone
arrays resulting in geometrically coincident or common reflection points. Then, after static effects to
the traces (shot to up-hole phone distance, weathering layer velocity and elevation changes across
the spread) are removed, all of the traces common to a single reflection point are gathered together
and stacked. This point on the reflection horizon is called CMP (common mid-point) or CRP (common
reflection point). Formally it was referred to as the CDP (common depth point) but changed in
recognition of the fact that the reflection points of the gathered traces do not coincide unless all
reflectors are flat.
Analytical estimates of velocity are used to determine stacking velocity (Vnmo) which is an optimum
velocity determined empirically from the seismic data, independent of borehole measurements.
Velocity analysis of seismic data is made possible by the redundancy of information in a CMP gather,
(Figure 2.6), and by the assumption that NMO curves are hyperbolas. Stacking velocities for a

17
CMP gather can be determined by three different methods: best fit hyperbola method, the T2, X2
method, or the constant velocity stack method. These will be discussed later on.

2.5 CMP stacking

CMP stacking was originally suggested by Mayne (1962) to enhance weak reflections in relation to
background noise, but it has also proved to be useful for reducing the signals from multiple
reflections.ii The principle of CMP stacking can be seen in Figure 2.6 where 6 shots are fired to 6
geophone arrays with their raypaths reflecting at point A where A represents (A1, A2, A3, A4, A5, A6).
S6 S5 S4 S3 S2 S1 R1 R2 R3 R4 R5 R6

A (Common Mid-point)

Figure 2.6 CMP gather of 6 traces

Assuming a horizontal reflector in Figure 2.6, trace 2 from shot record 1 has the same middle point as
trace 6 in shot record 2, trace 10 in shot record 3 and so forth. The number of traces in a CMP gather
is the fold of a stack. In the above example the fold is 6. However in modern two-dimensional (2D)
seismic surveys up to 280 channels can be recorded with folds commonly of up to 70. The formula for
calculating fold in a multi-channel recording is

fold = (channel spacing x number of channels) / 2 ξ


shot spacing

Example 2.1

Suppose we assume Figure 2.6 has a single, flat homogeneous layer with a velocity, V of 5000 ft/s,
geophone group interval of 100 ft and depth of 100 ft, then the reflection times calculated from
equation (2.6) of t1, t2, ….,t6 are approximately 0.0567, 0.089, 0.1265, 0.165, 0.204 and 0.243
seconds respectively. They would appear on a statics corrected CMP gather as shown in Figure 2.7.

Trace 2-S1 6-S2 10-S3 14-S4 18-S5 22-S6


0.0

.05

.10

.15

.20

.25

.30
sec.

Figure 2.7 (a) Uncorrected CMP Gather (b) Faux stack

18
If the traces were stacked at this point, the summation would produce one trace that bears very little
resemblance to any of the six original components Figure 2.7 (b).
One further calculation reveals that the normal-incidence time t0 = .04 sec. Then if the six traces
were corrected for NMO, the result would appear as in Figure 2.7(c) and stacking would produce a
single trace whose signal to noise ratio has been improved by

where n is the number of traces stacked. For example, a stack of nine common mid-point traces
would improve the signal to noise ratio (assuming the noise is random) by a factor of 3. An additional
nine traces would improve the ratio by a factor of 2 .

Trace 2-S1 6-S2 10-S3 14-S4 18-S5 22-S6

.05

.10

.15

.20

.25

.30

Figure 2.7(c) NMO corrected CMP gather (d) Stacked trace

Remembering that the stacking velocity Vnmo is chosen such that the relationship t2 = (x2 / V2) +t02
is satisfied (section 2.5), it is obvious that in the case of Example 2.1, in our one layer model, the
velocity that best describes that relationship is the one that transforms the reflection hyperbola,
t = [(x2 / V2) +t02]1/2 into a straight line, t0, which is of course V = 5000 ft/s, the original and only
velocity. In a multi-layer situation, RMS velocities are used in conjunction with stacking velocities.

2.5 RMS velocity

If there are two or more reflecting horizons that


x separate layers with different velocities the rays will
suffer refraction at the interfaces. For example, if
S G
there are two layers with velocities V1 and V2, the
actual path to and from the reflection point B (Figure
h 2.8) will not be SBG as was the case in the single
layer example but SABCG. Equation (2.4a) must
then be modified. Without going into detail, we can
say that if the layer thicknesses and velocities are
V1 such that the travel times in the two layers are equal,
A C the modification consists of replacing V by the root-
V2 mean-square (RMS) velocity, (V2RMS = (V12t1 +
V22t22…..+Vn2tn) / Tn where tn is the transit time
B through the nth layer and Tn is the total transit time to
the base of the nth layer) which is expressed as (V12
+ V22)1/2 / √ 2.
Figure 2.8 Reflected and refracted If the two travel times are different, say ∆t1 and
rays in a multi-layer medium ∆t2, we have instead (Sheriff and Geldart, 1982):

19
Σi=1 ∆ti)
2 2 1/2
Vrms = (Σi=1Vi2∆ti / (2.12) ξ

This equation for RMS velocity as applied here for two layers can be extended to n number of layers.

F igure 2.9 (a) shows an n layer example where all layers are flat and parallel to the surface with n
raypaths normal to the horizons. The thickness of each layer is V1∆t1, V2∆2,….Vn∆n respectively.

S x
R S R

V1∆t1 ∆Z1 V1

V2∆t2 ∆Z2 V2

Vn∆tn ∆Zn Vn

Straight line approximation

Figure 2.9 (a) (b)

In each layer, t = t0 and NMO = 0, therefore the velocity through n layers would be the average
velocity
Vavg = Σk=1Vk∆tk / Σk=1 ∆tk
n n

For a multi-layer example in which the raypaths follow a hyperbolic curve through n flat layers as
shown in Figure 2.9 (b), RMS velocity Vrms can be used to estimate NMO correction. Maclaurin’s
(infinite) series

Tx,n2 = T0,n2 + AnX2 + BnX4 + CnX6 + …..

relates Tx2 to X 2 for a reflection from the base of the n th layer. The coefficient of the second term is

Σk=1 Vk∆Zk
n n
An = (Σk=1 ∆Zk / Vk) /

and the coefficient of the third term is

(Σk=1 ∆Zk / Vk) Σk=1 (Vk3∆Zk)] / 16 Σk=1 (Vk∆Zk)4


n n n n
Bn = [Σk=1 (Vk∆Zk)2 -

In most actual cases, the values of Bn and coefficients of higher powers of X2 are so small that the
series can be truncated to two terms so that

/ Σk=1 Vk∆Zk
n n
Tx,n2 ≅ T0,n2 + X2 [(Σk=1 ∆Zk / Vk) (2.13) ξ

For abnormally long offsets, it may be necessary to use more terms. Rewriting equation (2.13) in

20
terms of velocity yields

/ Σk=1 ∆Zk / Vk
n n
V2 = (Σk=1 Vk∆Zk)

and since ∆Zk = Vk ∆tk, then

n
Vrms,n = [(Σk=1 Vk2∆tk) / tn]1/2
which is identical to equation (2.12). The NMO then for the reflection path, assuming flat velocity
layers and reflectors can be estimated from

Tx,n2 = T0,n2 + X2 / (Vrms,n)2 (2.14) ξ

The geometrical implication of this discussion is that the set of n layers can be approximated by a
single layer having velocity Vrms,n. The actual zigzag Snell path can then be replaced by the dashed
straight line path shown in Figure 2.9 (b).
The terms “RMS” and “stacking” velocity are sometimes used interchangeably, but it should be
remembered that RMS velocity is calculated from borehole velocity measurements and stacking
velocity is found empirically from the reflection seismic data.
The whole point of using the concepts of stacking or RMS velocity is that they enable the
estimation of NMO corrections by a single calculation. Ultimately, whatever velocity value best flattens
reflections within a CMP gather is the best velocity to use for NMO corrections.

2.7 Average and interval velocities

One-way vertical time below datum (sec.)


0 0.2 0.4 0.6 0.8 1.0 1.2

2000
Slope = Vavg to 5501’ = 10,038 ft/s
Depth below datum (ft)

Slope = Vavg to 6671’ = 10,597 ft/s


4000

Depth = 5501’
Time = .548 s
Slope of line segment
6000
between 5501’ and 6671’ = Depth = 6671’
Vint = 14,356 ft/s Time = .6295 s

8000

10000

Figure 2.10 Time-depth graph from well velocity survey (checkshot) with Vavg and Vint

21
T wo other useful velocities are average velocity and interval velocity. Interval velocity is the
average speed of a wavefront between two points, measured perpendicular to the velocity layers,
which are assumed to be parallel. Interval velocities can be calculated from RMS velocities using the
Dix formula

V2DIX = (Vb2Tb – Va2Ta) / (Tb – Ta) (2.15) ξ

The subscripts a and b denote, respectively, the top and bottom of the nth layer. RMS velocities are
normally slightly higher than true average velocities. Significant errors can arise if they are used
directly to make depth estimates, but the main source of error generally lies in the use of NMO data to
estimate velocity and Dix conversion may not help very much.
Average velocity is the ratio of vertical depth to the travel time of a wavefront from its source to
that depth.
Well velocity surveys (section 5.12) are the most accurate method for determining vertical travel
times to the depths at which down-hole times are recorded. It is therefore the best method of
measuring average velocities to those depths and the interval velocities between them. In Figure 2.10
a time-depth graph from a well velocity survey shows average velocities and interval velocities.

2.8 Static correction

I n section 2.4 it was mentioned that prior to NMO stacking, static effects must be removed from the
returned signal. Reflections are recorded at the earth’s surface, which may vary in elevation, after
they have passed through the weathered layer often referred to as the LVL or low velocity layer. The
weathered layer may vary in both thickness and velocity.
Static corrections are made for two principal reasons: so that changes in reflection time across a
stacked record section can be attributed wholly to subsurface effects and also that individual traces in
a CMP gather are properly aligned to preserve reflected signals when they are stacked. Two distinct
corrections are required: a weathering correction and an elevation correction.
A weathering correction replaces the actual travel time through the weathered layer by a
computed travel time. The computed travel time would result if the weathered layer (LVL) were
replaced by an equal thickness of the underlying higher velocity rock. Although the weathered layer
does not vary in thickness, its approximate thickness is usually known from previous surveys in the
prospect area. If the thickness of the weathered layer is not known, then it can be determined by one
or more uphole surveys.
Uphole time (msec)
Uphole geophone
5 10 15 20 25 30 35 40 45 50
0
Vw Vw = 2500 ft/s
Shot w 25 Indicated
weathering
50 depth = 38 ft

Shot 4
Shot depth (ft)

75 Velocity break

100 Vsw = 6250 ft/s

Shot 3 125
Vsw 150
Shot 2 175
200

Shot 1 225

Figure 2.11 (a) Uphole survey of 4 shots (b) Time –depth plot showing the computed
in sub – weathering and 1 in weathering V w, Vsw and the depth of the weathering

22
A n uphole survey requires a shot hole drilled to a depth of perhaps several hundred feet. An
uphole geophone is used to record shots from varying depths in the hole. The first shot is fired at
the bottom of the hole, and later shots are fired at fixed intervals up to the near-surface. Small
charges (between four and sixteen ounces for example) are used and shot depth intervals are
calculated to minimize damage to the shot hole and prevent sympathetic detonation of the other
charges. A plot of the uphole times versus shot depths usually shows the velocity break of the
weathered layer. (Figure 2-11).
The effect of the low-velocity layer can be seen in Figure 2.11 (b) by directly calculating velocity
(Vavg) from a given uphole time in the sub-weathering. Using shot 3 as an example Vavg = shot depth /
uphole time or in this case Vavg = 125 ft / 33 msec or 3788 ft/s, considerably less than the actual
interval velocity in the sub-weathering of 6250 ft/s calculated from the slope of the line through the four
plotted times or

Vint = ∆d / ∆t

where ∆d is the difference between shot depths in sub-weathering and ∆t is the difference of the
respective uphole times.
There are two static components: the source static correction which can sometimes be negated in
seismic explosive surveys by drilling the shot hole to a depth that would place the charge well below
the LVL and the geophone static correction since the returning signal must still pass through the
weathered layer.

Uphole geophone
The simplest, and often the best, method of
computing the weathering correction at a
t = dw / V w Vw geophone station is to use the uphole time at a
dw nearby shotpoint or a previous shot at that same
station. Figure 2.12 shows a common shot and
geophone station. The uphole time for a shot
Uphole
fired at depth ds is
signal
t=
ds - dw tuh = [(ds – dw) / Vsw] + (dw / Vw) ξ
ds - dw

Vsw
Vsw that is the uphole time = travel time in sub-
weathering plus the travel time in weathering so
that

tuh = [(ds – dw)Vw + dwVsw] / VswVw


ds Shot
tuh = (ds / Vsw) – dw[(Vw+Vsw) / VwVsw] (2.17) ξ
Figure 2.12 Calculating geophone static
correction from the uphole time = (ds / Vsw) - dw[(1/Vsw) + (1/Vw)]

Since the average velocity of the shot is Vavg = ds / tuh then uphole time can be expressed as the
average time (ds / Vsw) less the weathering correction (Wc), that is

tuh = (ds / Vsw) - Wc (2.18) ξ

then the weathering correction from (2.17) is

Wc = - dw[(Vsw – Vw) / VswVw] x

or Wc = - (dw / Vw) + (dw / Vsw) (2.19) ξ

and from (2.18) Wc = (ds / Vsw) – tuh (2.20) ξ

23
The correction velocity is usually fairly constant over large areas (except areas of complex near-
surface geology). This velocity may be known from previous experience, measured in uphole surveys
or determined from first arrivals refracted along the base of the weathered layer.
Since there is usually a shot hole located at every second or fourth geophone station, the normal
routine is to compute a weathering correction at every shot hole and interpolate for the other stations.
Computing weathering corrections from uphole times requires a shot below the weathered layer.
When a surface source such as Vibroseis, weight-drop or detcord is used, weathering corrections are
sometimes found by statistical methods or they may be computed from first arrivals. If the energy
source is above the base of the LVL, weathering corrections should be applied at both ends of the
reflection path. There is usually no weathered layer at the sea bottom. Along shorelines, however,
there are sometimes very thick sedimentary deposits of low velocity material.

T he effects of topography on a trace’s reflection times are removed by applying elevation


corrections which, in effect, move both source and receiver vertically to a preselected datum
plane (Figure 2.13).

ground surface
R

Er
Es
S’ R’ datum
plane
Ed
sea
level

Figure 2.13 Elevation corrections have the effect of moving S to S’ and G to G’.

The parts of reflection wavepaths near the surface are normally almost vertical unless the plane of
the reflector is quite steep. This is true for substantial offset distances, because wave velocity tends to
increase with depth. In Figure 2.13 the elevation corrections for source and receiver are

Ec,s = (Ed – Es) / Vsw (2.21) ξ

and Ec,r = (Ed – Er) / Vsw (2.22) ξ

A trace’s total static correction is the algebraic sum of its weathering and elevation correction as
applied to both the source and the receiver. It is applied by shifting the trace’s entire array of data
samples the required number of digital time intervals. A trace’s reflection amplitudes that appear
higher in a time section compared to adjacent statically correct traces require a positive correction to
shift the entire trace down in time and a negative correction to shift the trace up (Figure 2.14 (a)).

Tc = Wc + Ec

Wc = (ds / Vsw) – tuh , Ec,r = (Ed – Er) / Vsw

Tc,r = [(ds / Vsw) – tuh] + [(Ed – Er) / Vsw]

Tc,r = (ds – tuhVsw + Ed –Er) / Vsw (2.23) ξ

24
Vw

Vsw
Datum
plane

Reflector

Figure 2.14 (a) Before static and NMO correction showing what the 6 traces would look like in a
CMP gather at their relative arrival times. The trace on the right shows how they might look if they
were stacked and plotted in variable area at this point.

Datum
plane

Reflector

Figure 2.14 (b) The same 6 traces after static correction now aligned on an NMO curve

Datum
plane

Reflector

Figure 2.14 (c) After both static and NMO correction the traces in the CMP gather all align
horizontally with respect to the datum and stacking results in a trace indicating a single reflector.

25
2.9 Acoustic impedance and coefficients of reflection and transmission

W hile an elastic disturbance is passing over a point in a homogeneous medium, the stress at the
point is varying with time and the point itself is moving from its equilibrium position with a certain
velocity. The magnitude of the ratio stress/point velocity at any instant is a constant characteristic of
the medium and is called its acoustic impedance for the disturbance concerned (P, S, etc.). It is
shown in Appendix B.2 that the acoustic impedance, denoted by I is the product of the density ρ of the
medium and the phase velocity V (section 1.6) of the disturbance.
Consider now a plane P-wave in medium 1 (acoustic impedance I1) falling at normal incidence on
the interface with medium 2 (acoustic impedance I2). The wave is then partly reflected and partly
transmitted. At normal incidence there is not change of direction of the transmitted wave, that is, no
refraction (see equation (1.5)). If the displacement amplitude of the incident wave is 1.0 in the
direction of the incident wave, the displacement amplitude of the reflected wave is

Rc = (I2 – I1) / (I2 + I1) = (ρ2V2 – ρ1V1) / (ρ2V2 + ρ1V1) 2.24 ξ

in the direction of the reflected wave. The displacement amplitude of the transmitted wave is

Tc = 1-Rc = 2I1 / (I2 + I1) 2.25 ξ

where Rc is the reflection coefficient and Tc is the transmission coefficient. Together they are
known as partition coefficients since they determine the partitioning of the energy of the incident wave,
although the coefficients as given by equations (2.24) and (2.25) are for amplitude and not energy.
It should be noted that for reflection to occur ρ2V2 and ρ1V1 must differ from each other, while for
refraction it is sufficient that V2 and V1 differ. Rc is negative if I2 < I1. In such a case an incident
compression is reflected as dilatation (a negative pulse returned phase reversed from positive
transmitted pulse) and vice versa. If Rc is positive a compression is reflected as a compression (a
positive pulse returned from a positive transmitted pulse) and a dilatation as a dilatation.
In general, |Rc| < 1, but if I1 = 0, we have Rc = 1, in which case all the incident energy is reflected.
This is almost the case, for example, for a wave falling on the surface of the sea from a source in air.
If I2 = 0, then Rc = -1 and again all energy is reflected, but a compression as a dilatation and vice
versa. An example is a wave arriving at the surface of the sea from a source in the water.

Table 2.1 Typical reflection coefficients Rc

Interface Approximate Rc
Air over sea 1.0
Sea over limestone 0.65
Sea over boulder clay 0.45
Sea over recent sand/clay 0.3
Clays over gas sand, 500m - 0.3
Sea-bed multiples 0.25
Sand/shale over limestone, 1500m 0.20
10% change in acoustic impedance ± 0.05

Table 2.1 gives some typical reflection coefficients. Many factors like porosity, water content,
degree of lithification, etc., affect the acoustic impedance of rocks. In a sedimentary column, ρ and V
generally increase with depth due to compaction and the acoustic impedances in various layers tend
to become more like each other so that Rc decreases, with the result that the reflections from deeper
interfaces are generally weaker than those from shallower ones. In addition there are transmission
losses in the pulse energy due to geometrical divergence and transmission through interfaces.ii
As mentioned above the equation for Rc is based on a wave falling on the interface at normal
incidence. Actually the proportion of energy reflected increases as the angle of incidence increases.

26
2.10 Normal-incidence time section

T he incident and reflected rays SB and BG in Figure 2.8 on page 19 can be considered to be
almost vertical provided the detector is very close to the source. If S and G are coincident the two
rays will be exactly normal to the reflector. If we now associate a source with each of the detectors
and plot the respective two-way times below each detector, the graph will be a horizontal straight line
instead of the hyperbola in Figure 2.3 on page 15. If the reflector, instead of being horizontal, is
dipping or curved only the reflected ray corresponding to an incident ray that hits an interface normally
will arrive back at the detector. Whatever the path (straight, broken or curved) that the incident ray
takes in reaching the point of reflection, it will bounce back and retrace exactly the same path if the
incidence at the reflector is normal.

S1 S2 Sx

B1 Normal-incidence
section
B2 Bx
B3

A1 Ax

A2
Reflector A3

Figure 2.15 Construction of a normal-incidence time section over a curved reflectorii

The concept of normal incidence enables us to construct time sections when the geological
structure is more complicated than in Figure 2.8. The principle is illustrated in Figure 2.15 where a
curved reflector is shown and the material above it is assumed to have constant seismic velocity. In
this case the two-way time Ttwo of of a normal-incidence ray such as S1A1 will be proportional to the
path length 2S1A1. It is plotted at B1 below S1, the length S1B1 representing Ttwo on some suitable
scale. For a source-detector combination at S2 there will be two normally incident rays, namely, S2A2
and S2A3. Their Ttwos are plotted at B2 and B3. Continuing in this way we obtain the normal-incidence
time section shown. Since the diagram is only intended to illustrate the principle, reflections other than
those at the curved reflector have been disregarded. For a complicated multi-layer situation with
varying velocity the refraction at each interface must be taken into account to trace the path of a ray
and the Ttwo for it must be calculated with due regard to the path length and velocity in each layer.ii

2.11 Effect of dip

I n sections 2.3 and 2.4 the development of Vnmo from total travel times at offset x is based on the
assumption that velocity layers and reflectors are flat. Equation (2.4a) indicates that Tx is a
hyperbolic function of x. This hyperbolic time-distance condition is satisfied if velocity layers and
reflectors are flat, and in that case Vnmo and Vrms are virtually equal. Furthermore, it turns out that NMO
curves are still approximately hyperbolic if velocity layers and reflectors are parallel but not flat. In that
case

Tx2 = T02 + x2 / (Vrms / cos α)2 (2.26) ξ

and Vnmo = Vrms / cos α (2.27) ξ

where α is the dip angle of velocity layers and reflectors.

27
Figure 2.16 shows the derivation of the general expression relating Tx to x, where velocity layers and
reflectors dip at angle α.

S x/2 x/2 R
α α
(x / 2) sin α

(x / 2) sin α A
(Vt0) / 2

α
(Vt0) / 2 – ( x sin α) / 2
i
i

Vtx

(Vt0) / 2 + ( x sin α) / 2

From right triangle AS’R

(Vtx)2 = (Vt0)2 + (x cos α)2

S’ tx2 = t02 + (x2cos2 α) / V2

Source image point tx2 = t02 + x2 / (V / cos α)2 ξ

Figure 2.16 Derivation of expression relating tx to x with dipping reflector

In the above figure, rays are reflected from the dipping interface as if derived from the image point
S’ at a depth of 2d cos α below the surface where d is the perpendicular distance from the shotpoint to
the interface (d = (Vt0 + x sin α) / 2).
If velocity layers and reflectors are not parallel, then NMO curves are not hyperbolas and cannot
be closely approximated by simple expressions using Vrms or Vnmo. For small departures from the
parallel layer case, the conventional treatment of NMO is usually good enough. Where velocity
layering is very complex, it may be necessary to use special techniques involving Snell-path ray
tracing to compute NMO corrections.

2.12 Diffraction hyperbola and Fresnel zones

I t was shown in section 1.5 that a point of abrupt change to seismic properties of the ground is a
secondary source of waves. Referring to Figure 2.17, if S is a primary source exactly above the
edge of D, it follows that in addition to the reflected arrivals to the right of S, which give the hyperbolic
equation (2.1a), there will also be diffracted arrivals, whereas to the left of S there will be diffracted

28
arrivals but no reflections. The arrival time td of the diffracted wave is is

td = (h/V) + (x2 + h2)1/2 / V (2.28) ξ

in which the first term is the time for the primary wave to go from S to D, where it is diffracted, and the
second term is the time for the diffracted wave to go from D to G. Equation (2.28) is also an equation
for a hyperbola, termed the diffraction hyperbola.

0s
2h / V = t0

Time section

S x G

Diffracted ray h

D A

Figure 2.17 Diffraction


hyperbola (dashed curve) S’

The length of the path traveled by the diffracted wave is SD + DG = S’D + DG, if S’ is the image of
S in the reflector, while the path length for the reflected wave is SA + AG = S’A + AG = SAG, and S’D
+ DG > S’G. Hence td is always greater than the reflection time t given by equation (2.1a) except at
the source (x = 0) where td = t = t0.
Now consider a coincident source-geophone combination placed at G. The two-way time for the
diffracted wave will be 2(GD) / V and for this combination the diffraction hyperbola will be given by

td = 2(x2 + h2)1/2 / V (2.29) ξ

where x is the horizontal distance of the diffraction edge from the source-receiver point. The reflection
‘curve’ will in this case be simply t0 = 2h / V. The time section is shown in Figure 2.17 where the
dashed line represents diffracted arrivals and the full line the reflected ones.
From simple ray theory it might appear as if the energy received at the detector when a ray is
reflected comes from the point of incidence only. Actually it comes from within a certain area around
the point whose size we can calculate as follows.
Suppose we have a plane reflector at at distance r from the detector G, which is also the position
of the shot in the normal incidence concept. We imagine successive shperical wavefronts of radii r, r
+ λ / 2, r + λ, …, r + nλ / 2 where λ is the wavelength. These intersect the reflector and divide it into
(circular) zones as shown in Figure 2.18. Intermediate wavefronts intersect at intermediate points.
The zone division is such that the distances of any two adjacent circles from G differ by λ / 2. Then
the radius of a circle, the periphery of which is at distance r + nλ / 2 from G, the center being P, is

[(r + nλ / 2)2 – r2]1/2

and the area of the annular zone between circles with indices n+1 and n will be given by

π[r + (n + 1)λ / 2]2 – π[r + nλ / 2]2 ≅ πrλ

29
assuming λ << r. In reflection seismics λ is of the order of 100 m while r may be more than 1 km so
that this approximation has validity. It becomes evident that with this approximation all Fresnel zones
have the same area and the radius of the innermost zone, which is a circle and not an annulus like
each of the other zones, is (rλ).

Reflector
r P r+λ r + 4λ/2 r + nλ/2
r + λ/2 r + 3λ/2
Figure 2.18 Fresnel zones

By Huygen’s principle each intersection point gives rise to a spherical wavefront that produces an
effect at G inversely proportional to its distance while the total contribution of each zone is proportional
to the area πrλ. The zones are so narrow, if λ << r, that the distance of all the points in the nth zone
may be taken as r + nλ/2. The contribution of the nth zone at the detector G may therefore be written
as

K (πrλ) / [r + (nλ/2)]

where K is the constant of proportionality. The effect of the innermost zone n=0, circle of radius 1(rλ),
is Kπλ.
Wave theory tells us that if the path lengths of two waves arriving at a point differ by one whole
wavelength, the waves reinforce each other, while if the path difference is one-half wavelength, the
waves tend to cancel each other. Hence, the effects of adjacent zones at G are of opposite signs and
the total effect is given by the infinite series

Kπλ – (Kπrλ) / [r + (λ/2)] + (Kπrλ) / [r + (2λ/2)] - (Kπrλ) / [r + (3λ/2)] + …

= Kπλ[1 – 1 / (1 + λ/(2r)) + 1 / (1 + 2λ/(2r)) - 1 / (1 + 3λ/(2r)) + …]

In spite of the simple appearance of the series in the large parentheses, a closed expression for it is
not easy to obtain. It is given in fact by the integral
1


0
du / (1+uλ / 2r)

With the approximation λ << r we may put λ / 2r = 0 in the denominator of the integrand, and the
integral then evaluates immediately to ½ so that the total effect at G is found to be simply Kπλ / 2.
However, the effect of that part of the central zone which is within a radius a = rλ / 2 is exactly this
because its area is πrλ / 2 and we can say that the reflected energy at G comes mainly from within a

30
circle of radius a around the point at which a ray is incident. Now if t is the two-way time to the
detector then, at normal incidence, r = Vt / 2 where V is the average velocity down to the reflector.
Next, λ = V / ν if ν is the frequency of the wave. Hence, we finally get for the radius of the effective
area of reflection, the expression

a = (V / 2) t/ν (2.30) ξ

Example 2.2

If there is a reflection at t = 2.0 sec. and we have V = 2500 ms-1, a frequency component of 40 Hz will
have been reflected largely from a circular area of radius 279.5 m.

The bearing of Fresnel zones on seismic time sections is that, if a reflector is greater in areal
extent than the circle of radius given by equation (2.30), the reflection shows the shape of the reflector.
If the extent is less than this radius, a diffraction hyperbola predominates in the seismic expression
and the reflector’s shape can be difficult to discern.ii

2.13 Convolution (superposition)

A seismic record is primarily a record of the amplitude of the wave disturbance arriving at the
detector at different times. Consider a subsurface made of reflecting interfaces spaced so that the
travel time between successive reflectors is a constant, τ. Suppose a seismic source produces a very
sharp pulse (Figure 2.19). Such a pulse is often called a spike. The reflection record will then look as
in Figure 2.20, where the relative heights are proportional to the original spike height and to the
various reflection coefficients. A medium that exhibits this proportionality is said to be linear.
However, the ideal situation of Figure 2.19 is never obtained in reality not least because the seismic
source is not a spike.
amplitude

1.0

time
τ

Figure 2.19 A spike pulse Figure 2.20 Reflections of a spike


from various reflectors

Now consider a more realistic source pulse shown in Figure 2.21(a). It will be sufficiently accurate
for our purpose to characterize this pulse by a series of uniformly spaced ordinates 0, -4, -12, -8, 40,
16, -24, -8, 0 in arbitrary units. In other words, we consider the pulse to be made up of a succession
of spikes or impulses following each other at equal intervals of time. If the pulse is reflected from an
interface with an ideal reflection coefficient 1, the ground disturbance as a function of time will be
exactly a replica of the pulse (assuming the ground is a linear medium and the pulse undergoes no
change of shape in propagation). Suppose there is an interface (1) with reflection coefficient 0.5. On
reflection from this, the arriving pulse will produce half as much disturbance, which is written as the
series 0, -2, -6, -4, 20, 8, -12, -4, 0 in Figure 2.21(a).
The ground disturbance due to a reflection from another interface (2), one-half time unit ‘deeper’
than interface (1) that is, V / 2 depth units apart from (1), will start one time unit later since the
outward-going pulse has to travel to and back from interface (2) in being reflected from it. Suppose
the reflection coefficient of interface (2) is 0.25. A series of ordinates 0, -1, -3, -2, 10, 4, -6, -2, 0
displaced one time unit in relation to the earlier series 0, -2, -6, … will represent the ground motion due
to this reflection. Similarly a third reflector one time unit deeper than interface (2), and having a
reflection coefficient 0.5, will produce a disturbance displaced two units in time. Adding
(superimposing) the disturbances we see that the resultant ground disturbance will be represented by

31
the pulse in Figure 2.21(b) with sampled ordinates 0, -2, -7, -7, 16, 12, -12, 10, 6, -12, -4, 0 (or more
stringently expressed, by the ordinate series which we may use to reconstruct a first approximation of
the pulse form).

40
40 20

20 10
16

Time Time
0 0
0 0
-4
-8 -8
-12
(a) (b)
-24

0 -4 -12 -8 40 16 -24 -8 0

0 -2 -6 -4 20 8 -12 -4 0
0 -1 -3 -2 10 4 -6 -2 0
0 -2 -6 -4 20 8 -12 -4 0

0 -2 -7 -7 16 12 -12 10 6 -12 -4 0

Figure 2.21 Superposition of a series of reflection coefficients on an original source pulse

Exactly the same result can be obtained if, Reflector (1) (3)
instead of superposition, we use the following
procedure. The string of reflection coefficients (1), (2)
(2), and (3) is folded back on itself (convolved) and
slid past in discrete steps of one time unit, across
the series 0, -4, -12, … . For each position of the Coefficient 1/2
1/2 1/4
string, we multiply each of the three reflection
coefficients by the pulse ordinate of this series
found directly above that coefficient, and then add
the three results together. We obtain again the (Folding)
series 0, -2, -7, … . This procedure is known as
convolution and, as seen, its outcome is exactly
the same as that of superposition.
The procedure of convolution is given in Appendix A , equation (A.1). In that equation, convolution
of b with a is expressed as b ∗ a. The same series cp is obtained whether b is convolved with a, or a
with b by folding b, so that

b∗a=a∗b ξ

An important condition for the application of equation (A.1) is that the sampling interval must be
uniform and the same for both a and b. It will be noticed that in the example of Figure 2.21, the
interval is not uniform in the series of reflection coefficients. Before applying equation (A.1) it is
therefore necessary in this case to insert a coefficient a2 = 0 between reflectors (2) and (3) so that the
series a in this case becomes the evenly spaced series 0.5, 0.25, 0.0, 0.5. A reflector of coefficient 0
midway between reflectors (2) and (3) is implicit in the calculation of Figure 2.21(a) for it will merely
produce a row of zeros.

32
3
Seismic refraction

3.1 Scope

R efraction, the original method of seismic investigation, was widely used in the 30s and 40s for oil
and gas exploration but later replaced with the more technically sophisticated reflection method.
Today, refraction is used as an auxiliary procedure for obtaining velocity data and depth of
weathering layer for determining near surface corrections for the deeper reflection data. Now as then
it is also employed in civil engineering for bedrock investigations in connection with dam sites,
hydroelectric power stations and large scale building construction. It also used in detection of fracture
zones in hard rocks for groundwater prospecting, hazardous waste disposal projects etc.
In contrast to the reflection method, the maximum shot to receiver distances in the refraction
method are much larger than the depths of the interfaces of interest. In addition, travel times are
usually only a few tens of milliseconds and there is little separation between the arrivals of different
wave types. Generally, only the first P-wave arrivals, also called first breaks, can be picked with
sufficient confidence for use in depth calculations.
Another difference is in the design and employment of the surface detectors. For reflection work,
the geophones are laid out in an array such that unwanted signals such as airwave, ground roll and
direct arrivals are attenuated in favor of the deeper reflections. As the head wave (Figure 3.1) arrives
at the surface, each disturbance is detected by each geophone sequentially and tends to be cancelled
by the summation process of the geophone array. In refraction, the spread of geophones are usually
grouped, or podded at each station in order to enhance these signals.

S G G

ic ic

Figure 3.1 Cancellation of headwave by a nine detector geophone array

3.2 Spread design

A s previously mentioned, the length of the receiver spread for refraction studies is much greater
than the depth of the horizons of interest. Sufficient information on the direct wave and
reasonable coverage of the refractor is obtained if the length of the spread is about three times the
critical distance. A simple but not always accurate rule of thumb states that the spread length should
be at least eight times the expected depth of interest.
Often, refraction surveys are shot as a double-ended walk away. That is to say the shots are taken
from both ends of the receiver spread and then distanced progressively away from the spread. The
near shots provide information about direct and reflection arrival times and near surface velocities.
The far shots are placed at sufficient distance so that the first arrivals observed will have come by

33
way of the refractor.
In some surveys, center shots, which are shots within the spread, are taken. They are especially
useful if there are considerable differences in interpretation at opposite ends of the spread and if these
differences imply a different number of refractors. They may make it possible to obtain a more reliable
estimate of the velocity along an intermediate refractor and to monitor the thinning of an intermediate
layer that is hidden at one end of the spread by refractions from greater depths. An additional and
more reliable depth estimate is obtained which does not depend on assumptions about the ways in
which the thickness of the various layers vary along the spread. Center shots also provide extra data
on the direct wave velocity. Although shooting within the spread implies an additional cost, it is
generally worth the investment for the increased reliability of the data in complex near surface velocity
layers.

3.3 Principal refractors

P -wave velocities for some common rocks were given in Figure 1.1 on page 5. In shallow refraction
work, it is often sufficient to describe the ground in terms of dry and wet overburden and fresh and
weathered bedrock. Dealing with more than three refracting interfaces considerably complicates the
interpretation as do dip and non-parallel interfaces. Ideally, the interfaces studied in a small refraction
survey should be shallow, roughly planar and dip at less than 15º. The velocity must increase with
depth at each interface, as the first arrivals will then come from successively deeper interfaces as the
distance from the shot-point increases.
The P-wave velocity of dry overburden may be as low as 1100 to 1200 feet per second, which is
close to the velocity of sound in air, and rarely exceeds 2800 fps. There is usually a slow almost
imperceptible increase with depth due to the increase of the weight of the overburden that is followed
by an abrupt increase to 5000 – 6000 fps at the water table.
Fresh bedrock often has a P-wave velocity of more than 8000 fps but is likely to be overlain by a
transitional weathered layer where the velocity may be less than 6500 fps and usually increases
steadily with depth and the accompanying reduction in weathering.

3.4 Critical refraction and the Snell ratio

S nell’s law (section 1.4) implies that if, as in Figure 1.3, V is greater than V and if sin i = V /V , the
2 1 1 2
refracted ray will travel parallel to the interface at velocity V . After critical refraction, some of the
2
energy will return to the ground surface as a head wave represented by rays which leave the interface
at the critical angle. The head wave travels through the upper layer at velocity V1, however, because
of its inclination, appears to move across the ground at the V2 velocity with which the wave front
expands below the interface. It will therefore eventually overtake the direct wave, despite the longer
travel path.
Consider a situation where a layer of thickness h in which the waves have a velocity V1 is
underlain by another with velocity V2. Then by Snell’s law,

V1 / V2 = sin i / sin r

where i and r are the angles of incidence and refraction for the seismic ray (Figure 1.3) initiated at S1
and traveling through A. The ray S2B, which describes a critical angle of incidence ic = sin-1 (V1 / V2),
is refracted so that r = 90º and travels along the boundary between the two media. This is only
possible if V2 > V1.
Figure 3.2 illustrates a refraction shot in which ray S1A intersects the interface at critical angle ic.
Owing to this ray, the interface is subjected to oscillatory stress and each point on it sends out
secondary waves, and rays such as BG3, CG4 etc. emerge in the top layer at the angle ic to reach the
respective geophones G3 and G4 etc. It should be noted that a ray intersecting the horizon at the
critical angle is partly reflected back to the surface and partly refracted along the horizon. However, it
will be shown that the geometry indicated in Figure 3.2 does not allow for a coincident reflection-
refraction since critical point A occurs prior to the point at which a reflection could be detected at the
surface geophone station. Conversely, up to a certain minimum distance xm from the shot, there are
no refracted arrivals.

34
Let a represent the surface projection of the line segment N1A. Since tan ic = a / h then

a = h tan ic (3.1) ξ

The minimum distance at which a refraction could be recorded at the surface would be double that
allowing for two way time. Therefore

xm = 2h tan ic (3.2) ξ

Within this distance, there are no refracted arrivals.

S2 S1 G1 G2 G3 G4 G5 G6

V1 ic ic ic

V2 N1 R11 R12 A B C D E

Figure 3.2 Critical refractions

Example 3.1

Given velocities V1 = 3500‘/s, V2 = 5000‘/s, depth to horizon h = 150’, and group intervals S1-G1, G1-
G2, G2-G3 etc.= 100‘, estimate (1) the critical angle ic, (2) minimum distance xm, (3) direct arrival
times, (4) refraction arrival times and (5) reflection times S1R11G1 and S1R12G2.

(1) From equation (1.7), ic = sin-1 (V1 / V2)


= sin-1 (3500 / 5000)
ic ≈ 44.427º

(2) From equation (3.2), xm = 2h tan ic


= 2h tan [sin-1(V1 / V)]
= 2(150’) tan [sin-1(3500 / 5000)]
xm ≈ 294.0588 ft

Note that the distance from S1 to G3, x = 300‘ and assuming parallel and planar surfaces, that if
there were to be a reflection recorded at that distance x, it would have been reflected at a point below
one half x or 150 ‘. Since a = h tan ic ≈ 147.0294, the minimum distance at which refractions occur,
then there can be no reflection recorded at G3 from a shot taken at S1 as (x / 2) > h tan ic.

(3) If x is the distance from shot to geophone, then the direct arrival time is simply

t = x / V1 (3.3) ξ

tS1G1 = 100‘ / 3500‘/s tS1G1 ≈ 0.028571 sec


tS1G2 = 200’ / 3500’/s tS1G2 ≈ 0.057143 sec

continuing, tS1G3 ≈ 0.085714 sec


tS1G4 ≈ 0.114286 sec
tS1G5 ≈ 0.142857 sec
tS1G6 ≈ 0.171429 sec

35
(4) Since xm ≈ 294 feet then the first geophone station to record a refraction arrival will be G3 at a
shot-receiver distance of 300 feet. Ray S1A travels from S1 to A in velocity medium V1 as does ray
BG3, CG4 etc., while the refracted ray traveling from A to B does so in V2. The refraction arrival time to
G3 can then be directly approximated. TS1ABG3 = (S1A / V1) + (AB / V2) + (BG3 / V1). Now that a and h
are known, the distance S1A can be calculated as the square root of a2 + h2 (Pythagorean theorem) or
since ic is also known, cos ic = h / S1A thus S1A = h / cos ic. In this example where the horizon and the
ground surface are planar and parallel, S1A = BG3, and AB = x – 2a, where x is the distance from the
shot at S1 to G3. Therefore,

TS1ABG3 = (S1A / V1) + (AB / V2) + (BG3 / V1)


= 2(S1A / V1) + (x – 2a) / V2
= 2((h / cos ic) / V1) + (x – 2(h tan ic)) / V2
= 2h / V1cos ic + (x – 2h tan ic) / V2
= 2h / V1 cos sin-1(V1/V2)) + (x – 2h tan sin-1(V1/V2))) / V2

≈ 0.121212 sec

Similarly, by augmenting each subsequent travel segment in V2 by one group interval such that AC =
AB + 100 feet, AD = AB + 200 feet etc., then the remaining travel times TS1ACG4 = (S1A / V1) + (AC /
V2) + (BG4 / V1), TS1ADG5 = (S1A / V1) + (AD / V2) + (BG5 / V1), TS1AEG6 = (S1A / V1) + (AE / V2) + (BG6 /
V1), can be approximated.

TS1ACG4 ≈ 0.141212 sec


TS1ADG5 ≈ 0.161212 sec
TS1AEG6 ≈ 0.181212 sec

(5) The two-way reflection time given by equation (2.1a) is

t = (2 / V1) h2 + x2 / 4

Therefore, TS1R11G1 = (2 / 3500) 1502 + 1002/4


TS1R11G1 ≈ 0.09035
TS1R12G2 ≈ 0.103016

3.5 Field records

S ince the beginning, hard copy records have been produced in the field, originally by
photographically recording the analog signal and more recently by digitally converting the signal
and reproducing it on dot matrix or ink-jet printers. On some refraction crews, hard copies have given-
way to instant analysis and interpretation on computer screens. In a day’s work which includes
repeats, checks and tests as well as the shooting of a number of different spreads, several dozen
records may be produced and must be carefully annotated if confusion is to be avoided. Annotation
should include the date, time and name of the observer, along with the survey location and spread
number. Orientation should be noted, and the position of geophone station 1 should be defined.
Unless the geophone spacing is uniform, a sketch showing the shot and geophone locations should
be added to the documentation. Additionally, information about the type of instruments, filter and gain
settings and the use of special equipment such as S-wave geophones should be provided.

3.6 Picking first arrivals

P icking first arrivals on refraction records may be difficult at times due to poor signal-to-noise ratio at
far offsets and by coincident time arrivals by different types of waves. Figure 3.3 represents a
refraction record of one shot to six geophones over a time of 400 milliseconds. Notice the coincidence
of a reflection on trace 1 with the airwave and the proximity of a refraction to the direct arrivals on trace
6. These effects tend to distort the amplitudes of the arrivals making interpretation difficult.

36
Since high frequencies are selectively absorbed in the ground, the distance between the first
break and any later peak gradually increases from the source. Furthermore, the trace beyond the first
break is affected by many other arrivals as well as by later parts of the primary wave train, and these
will modify the peak and trough locations.

G1 G2 G3 G4 G5 G6 G1 G2 G3 G4 G5 G6 G1 G2 G3 G4 G5 G6
0.000 0.000 0.000

0.020 0.020 0.020

0.040 0.040 Direct arrivals 0.040

0.060 0.060 0.060

0.080 0.080 0.080


Reflections
0.100 0.100 0.100
Refractions
0.120 0.120 0.120

0.140 0.140 0.140

0.160 0.160 0.160

0.180 0.180 0.180

0.200 0.200 0.200

0.220 0.220 Ground roll 0.220

0.240 0.240 0.240

0.260 0.260 0.260

0.280 0.280 0.280

0.300 0.300 0.300


Air wave
0.320 0.320 0.320

0.340 0.340 0.340

0.360 0.360 0.360

0.380 0.380 0.380

0.400 0.400 0.400

Figure 3.3 (a) (b) (c)

Analog refraction Identification of the Time picks


record wave time arrivals

The first arrival on traces 1 - 6 is the direct arrival having traveled the shortest distance at the
fastest velocity. Having picked the first arrivals, a best-fit line is drawn connecting the points of origin
of the first troughs (the industry standard recognizes a down break as the first break which implies
negative voltage). The first arrival of direct wave, airwave and ground roll can be estimated with a little
knowledge of the near surface geology. It can be seen on the record that the direct wave line and the
refraction line are near convergence in fact, if there were a seventh trace, it would indicate that the two
waves arrived at nearly the same time.
Since the velocity of sound in air is very consistent, varying only slightly with changes in
temperature, barometric pressure and humidity, the arrival times to all of the traces along the spread
can be estimated fairly accurately. Ground roll and the direct wave are a bit more difficult; however,
information on near surface velocities in the area is generally easily obtainable. The most difficult of all
the arrivals to pick are the reflections. Their amplitudes are generally masked by the more stronger
airwave and ground roll. Figure 3.3 shows that the first reflection break on trace 1 is coincident with
the first arrival of the airwave and that on trace 2, the reflection and ground roll coincide.

37
Example 3.2 (1) From Figure 3.2(c), pick the times of all the waves at the point where the best-fit
lines intersect the geophone traces. (2) Using any two time picks from from the air wave, ground roll
and direct wave picks estimate their velocities. (S1G1 = G1G2 … = 100’)

Wave type G1 G2 G3 G4 G5 G6 [(y2 – y1) / (x2 – x1)]-1


Airwave .091 .182 .273 .364 .455 .545 [(.364s - .091s) / (400’ – 100’)-1 = 1098’/s
Ground roll .050 .100 .150 .200 .250 .300 [(.15s - .1s) / (300’ – 200’)]-1 = 2000’/s
Direct wave .029 .057 .086 .114 .143 .171 [(.057s - .029s) / (200’ – 100’)]-1 = 3571’/s
Refraction .121 .141 .161 .181
Reflection .090 .103

3.7 Time-distance plots

T he data extracted from a refraction survey consist of sets of times measured at geophones at
various distances from the source position. These are plotted against vertical time axes and
horizontal distance axes (Figure 3.4). The gradient of any line is thus equal to the reciprocal of a
velocity and steep slopes correspond to slow velocities.

0.320

0.300 ~
0.280 Ground roll
> 2000'/s
0.260 Airwave
1100'/s ~
0.240

0.220
Reflected
0.200 ~
0.180 > +
0.160 +
~
0.140 + ^ Refracted
0.120 +
^
0.100 *
~ Direct wave
0.080
*> ^ 3500’/s

0.060
^
~
0.040

0.020 ^ xm
0.000
G1 G2 G3 G4 G5 G6 xc

Figure 3.4 Time-distance plot from refraction survey

The velocities of the airwave, ground roll and direct wave can be estimated directly from the
inverse of the slopes of their respective lines. The time axis intercept of the reflection hyperbola is the
zero-offset time obtained from equation (2.1a) by setting x = 0. The minimum distance for recording a
refraction is shown as xm, which is defined in equation (3.2) as being equal to 2h tan ic. The distance
xc is the crossover distance, that is the distance at which the refracted wave overtakes the direct wave
and will be proven in the section 3.8.

38
3.8 The refraction equation

I f G is near to the shot S (which is assumed to be on the surface) as is the case in Example 3.1, the
first wave to arrive at the geophone will be the direct wave along SG. However, if SG is sufficiently
great the first arrival will correspond to the wave SABG which will have overtaken the direct wave
because of a higher velocity along the path AB.
If x is the shot-geophone distance, the travel time for the direct wave t is equal to x / V1 and its
plot against x will be a straight line through the origin, that is the coordinate of the shotpoint, with slope
1 / V1 (Figure 3.7).
The travel time for a refraction ray S1ABG in Figure 3.5 can be derived from the Snell’s law
equality (section 1.4)

sin ic = V1 / V2

S1 x G
a

h h
m1
ic ic
V1
N1 A B V2

Figure 3.5 Refraction equation

Let m1 represent the distance from S1 to A


a represent the surface projection of N1A
h the depth to the refracting horizon
and x the horizontal distance from S1 to G

From Snell’s law (equation (1.7)), sin ic = V1 / V2 and in Figure 3.5 sin ic = a / m1

therefore, V1 / V2 = a / m1 , m1 = aV2 / V1 and also m1 = (a2 + h2)1/2

so that, a2V22 / V12 = a2 + h2

a2(V22 – V12) = h2V12

then, a2 = h2V12 / (V22-V12) ξ

From the Pythagorean theorem, m12 = a2 + h2

substituting for a2 m12 = h2V12 / (V22 – V12) + h2

m12(V22 – V12) = h2V12 + h2(V22 – V12)

m12 = h2V22 / (V22 – V12)

m1 = hV2 / (V22 – V12)1/2 ξ

Let tSA be the travel time along m1 and substitute the expression (V22 – V12)1/2 with z

then, tSA = m1 / V1 = (hV2 / z) / V1

tSA = hV2 / V1z

39
Assuming that the ground surface and refracting interface are planar and parallel, and letting tm be the
total time from S1 to A and from B to G, then

tm = 2hV2 / V1z ξ

If x is the distance from shot to receiver and a is the surface projection of N1A, then the distance from
A to B is

AB = x – 2a since a = hV1 / (V22 – V12)1/2, then

AB = x – 2hV1 / z

Therefore the travel time from A to B is

tAB = (x – 2hV1 / z) / V2 ξ

and the total refraction time T along SABG is

T = 2hV2 / V1z + (x – 2hV1 / z) / V2

= [2hV22 + V1xz – (2hV12z) / z] / V1V2z

= [V1xz / V1V2z] + [2h(V22 – V12) / V1V2z]

= (x / V2) + 2h(V22 – V12) / V1V2z

Finally, replacing z with the expression (V22 – V12)1/2

T = (x / V2) + 2h(V22 – V12)1/2 / V1V2 (3.4) ξ

which describes a straight line with slope x / V2 and an intercept of the time axis given by the second
term of the equation (Figure 3.7).

Example 3.3

Verify the refraction time along S1ABG3 directly estimated in Examples 3.1 and 3.2 with that using the
refraction equation (3.4) using the same parameters x = 300’, h = 150’, V1 = 3500’/s and V2 = 5000’/s

T = (300/5000) + 2 • 150(50002 – 35002)1/2 / 3500 • 5000

T ≈ 0.121212243 sec

which agrees with the previous direct estimates.

Example 3.4

Use equation (3.4) to find the slope and intercept of the refraction line.

slope = x / V2

slope = 300 / 5000 = 0.06

40
intercept = 2h(V22 – V12)1/2 / V1V2

= 2 • 150(50002 – 35002)1/2 / 3500 • 5000

intercept ≈ 0.061212

3.9 Snell right triangle and crossover distance


V1

I mplicit in Snell’s law, sin ic = V1 / V2, is the


construction of a right triangle (Figure 3.6) in
which V1 is the side opposite the critical angle ic
h
V2
and V2 is the hypotenuse. Letting h be the side ic
adjacent to angle ic, the refraction equation (3.4)
can be simplified and rewritten in terms of the
cosine of ic.
Figure 3.6 Snell right triangle

The second term of the right-hand side of equation (3.4)

2h(V22 – V12)1/2 / V1V2 = (2h / V1) ((V22 – V12)1/2 / V2)

from Figure 3.6, cos ic = h / V2 and h = (V22 – V12)1/2

then, cos ic = (V22 – V12)1/2 / V2

substituting this into the second term of the refraction equation yields

T = (x / V2) + 2h cos ic / V1 (3.5) ξ

which is the same straight line with slope x / V2 and time axis intercept 2h cos ic / V1. The second
term is called the delay time ∆c in the layer.

A t a certain distance from the source, the refracted wave will overtake the direct wavefront and
arrive first at the detector. The point xc at which this occurs is often called the break point and the
distance is called the crossover distance. Also called the critical distance, it can be found by
setting the equation for the time of a direct arrival (equation (3.3)) equal to the equation for refraction
time (equation (3.4)) and solving for x.

x / V1 = (x / V2) + 2h(V22 – V12)1/2 / V1V2

x / V1 - x / V2 = 2h(V22 – V12)1/2 / V1V2

x(V2 – V1) = 2h(V22 – V12)1/2

x = 2h(V22 – V12)1/2 / (V2 – V1)

x = 2h(V2 + V1)1/2(V2 – V1)1/2 / (V2 – V1)

Therefore, xc = 2h[(V2 + V1) / V2 – V1)]1/2 (3.6) ξ

Thus, a time-distance graph of the first arrivals to geophones planted at various distances from the
shot will show two intersecting straight lines whose slopes give V1 and V2 (Figure 3.7(a)).

41
3.10 Delay times and multiple-level refractions

U sing equation (3.6) the coordinate xc of the point of intersection yields the thickness h1 of the top
layer. Alternatively reading off the delay time ∆c, h1 is defined as

h1 = V1∆c / 2cos ic = V2∆c / 2cot ic (3.7) ξ

No refracted
arrivals

Slope 1/V3

Slope 1/V2

ti3
Time

Slope 1/V2

Time
ti2
Slope 1/V1

Slope 1/V1

Distance from S xm xc xc xc2

S R G S Gm Gn

i2 i2
M ic ic
i ic V1
ic ic
V1 C A F V2
B
T r
A B V2 ic2
ic ic2 ic2
V2
(a)
D E V3
(b)

Figure 3.7 Single and multi-layer refraction time-distance graphs

The delay time ∆c is composed of two parts; the first part is the delay at S, which is the difference
between the time for the actual path SA traveled with velocity V1 and the virtual path TA along which
the velocity is V2, and the second part is the delay at Gm. The delay at S is SA / V1 – TA / V2.
However, SA = SM + MA = h1cos ic + TA sin ic, while TA / V2 = (TA sin ic) / V1 since sin ic = V1 / V2,
which leads to the delay at S as (h1cos ic) / V1. Together with the equal delay at Gm, the total delay
time is obtained which is the second term of the equation (3.5).
Being dependent only on the time difference between the slant path of a ray and the projection of
the path on the interface on which the ray strikes, a delay time of ∆ = (2hcos i) / V, using appropriate
values for h and V, can be associated with a ray striking at any angle of incidence i.
For three layers with velocities V1 < V2 < V3, there will be two critically refracted rays: SABGm
along the first interface and SCDEFGn along the second one (Figure 3.7(b)). As before, at short
distances the direct wave SGm will arrive first. As the distance increases the ray SABGm will overtake
SGm and arrive first, while at still greater distances the first arrival will be that of the ray SCDEFGN and

42
the time-distance plot then consists of three intersecting straight segments with slopes 1 / V1, 1 / V2
and 1 / V3.
If the critical angle of incidence at D is denoted by ic2, then this is also the angle of refraction at C
and it is related to the angle of incidence of ray SC (i2), by Snell’s law as.

sin i2 / sin ic2 = V1 / V2 (3.8) ξ

from which is obtained sin i2 = V1 / V3.

T he delay times for the refraction ray SCDEFGn in the two layers through which is passes are then
∆2 = (2h1cos i2)/V1 and ∆c2 = (2h2cos ic2)/V2 respectively and analogous to equation (3.5), the
equation for the third straight-line segment of the time-distance curve is:

t = [x / V3] + [(2h1cos i2) / V1] + [(2h2cos ic2 ) / V2] (3.9) ξ

In this equation, the last two terms together give the intercept of the third segment on the t axis.
This time is known as the intercept time ti3. Similarly, the second term in equation (3.9) can be
denoted ti2, although in that case it happens to be equal to the delay time ∆c. The intercept ti of the
first segment (direct wave of velocity V1) is zero. All the quantities involved in the intercept time ti3
except h2 are known, since h1 is known from equation (3.7). Hence, h2 can be calculated. Explicitly,

h2 = (ti3V2 / 2cos ic2) – (h1V2cos i2 / V1cos ic2) (3.10) ξ

Alternatively, it can be shown that, after solving for h1 in equation (3.6), h2 can be obtained from

xc2(V3 – V2) = 2h1 / V1[V2(V32 – V12)1/2 – V3(V22 – V12)1/2] + 2h2(V32-V22)1/2 (3.11) ξ

where xc2 is the coordinate of the intersection point of the second and third segments.
The advantage of equations (3.6) and (3.11) is that no trigonometric calculations are needed since
only the break points xc and xc2 and the slopes of the segments are involved. However, modern
refraction seismic calculations are done by computer programs and in practice the selection of the
formulas to be used is dictated by which characteristics of the time-distance curve are most distinct.
The previous relationships can be extended to any number of horizontal layers, one below the
other, provided the velocity in each layer is greater than that in the layer immediately above. In
general, there will then be as many distinct segments on the time-distance curve as there are layers.
If, however, a layer is not sufficiently thick or does not have a sufficient velocity contrast with the
adjacent layers, the segment corresponding to it may be missing on the time-distance graph of the first
arrivals. Such a thin layer, called a ‘hidden layer’ or ‘blind zone’, can however be detected sometimes
by recording the later arrivals. The presence of a hidden layer obviously introduces errors in depth
calculations to the lower interfaces. In the three-layer case, if the second layer is a blind zone, the
interpretation will be carried out as if there were a two-layer case and the depth to the V3 layer would
be underestimated.
If the velocity increases continuously with depth, for instance, approximately linearly as is often the
case, the time-distance graph will be a smooth curve concave towards the x axis. If a layer has a
lower velocity that the one on top of it there cannot be any critically refracted ray because the
refraction angle r will always be less than the incident angle i in Figure 3.7(a). No energy can be
transported along such an interface and no segment corresponding to the lower layer will appear in
the time-distance curve. It is worth noting, however, that such layers could be detected if shear waves
were also taken into account.ii

3.11 Non-parallel interfaces

Ithe
n Figure 3.8 is shown a case where the interface between two layers is dipping at an angle θ with
horizontal. The path LA B G is exactly analogous to SABG in Figure 3.7 and the time along it
1 1
can be immediately written down from equation (3.5) after replacing x by x cos θ and h1 by (z1 – ζ ) as

43
(x cos θ / V2) + 2(z1 – ζ )cos ic / V1

S1 x G x S2
θ
ζ x cos θ
L z2
ic
z1 ic
θ
z1 - ζ ic
A2
ic B2
B1
V1 A1

V2

Figure 3.8 Dipping interface

where z1 is the perpendicular distance of the dipping interface from the shot S1. To this, the delay for
the path part S1L which is (ζcos ic) / V1 must be added and this gives the total time t for the path
S1LA1B1G.

t = (x cos θ / V2) – ( ζ cos ic / V1) + (2z1 cos ic / V1) (3.12) ξ

However, ζ = xsin θ and V2 = V1 / sin ic. Substituting and simplifying, equation (3.12) becomes

t = (x / V1) sin(ic – θ) + (2z1 cos ic / V1) (3.13) ξ

which is the total time for the up-dip ray originating at S1. The time for the down-dip ray S2A2B2G is
obtained by replacing - θ with + θ and z1 with z2 so that

t = (x / V1) sin(ic + θ) + 2z2cos ic / V1 (3.14) ξ

It should be noted that these relationships would be interchanged if the surface is sloping and the
interface is horizontal.
The time-distance curve for the direct ray has the slope 1 / V1 whether the ray comes from the
shot S1 or shot S2. However, the segment corresponding to the refracted ray has a slope sin(ic – θ) /
V1 when shooting up-dip but sin(ic + θ) / V1 when shooting down-dip. The reciprocals of these slopes
are the up-dip and down-dip velocities Vu and Vd respectively, and the angles θ and ic are given as

θ = ½ [sin-1(V1 / Vd) – sin-1(V1 / Vu)] (3.15) ξ

and ic = ½ [sin-1(V1 / Vd) + sin-1(V1 / Vu)] (3.16) ξ

(Both Vu and Vd are assumed to be positive.)


The dip is directly determined from equation (3.15) while the depths z1 and z2 which complete the
information about the configuration in Figure 3.8 are obtained from the intercepts of the lines
represented by the equations (3.13) and (3.14) on the t axis after substituting for ic from equation
(3.16).
The up-dip velocity Vu = V1 / sin(ic – θ) is positive if θ < ic and negative if θ > ic. If θ = ic, it becomes
infinite so that the corresponding time-distance segment is horizontal. The infinite velocity is, of
course, only apparent; no energy is transmitted at this velocity. Noting that the down-dip velocity Vd =
V1 / sin(ic + θ) and sinic = V1 / V2, it can be proven that

44
V2 = cos θ [(VuVd) / (Vu + Vd) / 2] (3.17) ξ

In practice, the interface dips encountered in most shallow refraction work are small so that V2 can be
estimated by putting cos θ ≈ 1 so that

V2 ≈ VuVd / (Vu + Vd) / 2 (3.18) ξ

This estimate is correct to better than 5% even if the dip is as much as 20°.
The numerator in equation (3.18) is the product and the denominator the mean of Vu and Vd so
that an easy to remember product divided by mean rule can be made for the estimate of V2. If there
are more than two layers the rule is valid for the true velocity in each layer provided that the
corresponding up-dip and down-dip velocity segments can be identified in the time-distance graphs. It
is, however, very rare that more that three layers can be identified with certainty in shallow refraction
seismics.

3.12 Complex refraction survey

T he topographic and layering conditions in reality are generally quite different from the idealized
ones considered above and the time-distance graphs rarely consist of regular straight-line
segments as in Figure 3.7. An example of a refraction profile that is particularly complex is shown in
Figure 3.9 (Parasnis, 1997). The profile shown was shot for a hydroelectric project in northern
Sweden. Each star or cross represents the travel time to a geophone placed at a point vertically
below it on the ground surface. The depths sounded on this profile are relatively shallow but the
results nevertheless illustrate a number of points discussed above.

0.040
I6
G

0.030 I5

I1

0.020

0.010

S5 S6
S4 -1
Surface S2 S3 350 ms
S1 1000 ms-1
1700 ms-1
350 ms-1 Bog
2100 ms-1 Glacial till 2100 ms-1
-1
4500 ms
Borehole Bedrock interface
0 25 m

Figure 3.9 Time-distance graph for complex near-surface geology

The steep initial segments near each shot point correspond to the upper layers and at large
distances from the shots these are replaced by less steep segments. This is most clearly indicated by
the left-hand geophone set-up from shot 4. The break points between segments or intercept times

45
can only be determined after appropriately correcting for the irregularities (such as I1, I5 and I6) which
may be caused by local inhomogeneities or variations in interface dips, for which case it is useful to
remember that the undulations in a time-distance graph are, qualitatively speaking, mirror images of
an interface undulation. It should also be noted that very small dip variations could cause very
pronounced irregularities of this type in the time-distance curves.
As the refraction method depends primarily on picking the first arrivals, the time of which are
plotted to give a time-distance graph, it is important that the arrivals are correctly picked. When
geological or topographical conditions are complicated, the refraction records, whether photographic
as in analog equipment or on a computer screen as in digital systems, can be unclear and doubt may
remain whether some arrivals have been correctly picked. An internal check that is useful in such
cases employs the principle of reciprocity. This states that the time of any seismic arrival at the
geophone location G when the shot is at point S is the same as that when the geophone is at S and
the shot at G, irrespective of whether interfaces are dipping or not or how complicated the structure
otherwise is.

Example 3.5

Referring to the time-distance graph in Figure 3.9, a careful measurement reveals that the first-arrival
time for a shot taken at S6 to a geophone at S4 is 37 ms. In the reverse graph, when the shot is at S4
and the geophone at S6, the plotted first-arival time is 29.5 ms. Similarily, with the shot at S5, and the
geophone at S3, the time is 33 ms while in the reverse direction, with the shot at S3 and the geophone
at S5, the time is 26 ms. When such discrepancies occur, the refraction record should be rescanned.
In order to more easily observe the variations, timing lines at every one or two milliseconds should be
drawn through the geophone locations.

Often in reverse shooting the geophone has to be displaced slightly due to the explosion crater
from the forward shot so that the geophone locations plotted above are not exactly at the shot.
Nevertheless, the time discrepancies in the shot-geophone pairs in question cannot be attributed to
these small displacements or to inaccuracy in reading the figure because other reciprocal pairs in the
figure give identical first-arrival times. For example, in the pair S2 and S4, the time is exactly 26 ms in
both directions.
The apparent velocity in the bedrock when the geophones occupy positions around G is less when
the shot is fired from S1 (steeper time-distance segment) than from S6. Evidently the ground surface
and the bedrock are not parallel to each other in this region.
In the right-hand part of Figure 3.9 the time-distance curves are more or less smooth indicating a
fairly continuous downward increase in velocity. The interpretation of the profile is also shown in the
figure together with a borehole that checked the depth, but it is clear from this example that some
refinements in the treatment of the data are needed before the formulas in the previous section can be
used in practice.
The basic problems are (1) to determine the various velocities and (2) to determine the intercept
times or break points between segments of different slopes. It is preferable to use intercept times
rather than break points for depth calculations. A method for estimating these parameters is
presented is the following section.

3.13 The ‘mean-minus-T’ method

T his is a rapid method for estimating the velocity and its variations in a refractor in detail. If TA and
TB are the arrival times at one and the same geophone from two sufficiently distant shot points A
and B in opposite directions from the geophones (Figure 3.10) and if ∆T = TB – TA then

∆T / 2 = ((TB + TA) / 2) - TA (3.19) ξ

The first term is the mean of the arrival times while the second term is the time itself. The time
difference ∆T1, etc., between the refractor velocity curves A and B (for the same refractor) are
measured and ∆T1 / 2, etc., are plotted, with due regard to sign, from any arbitrary, conveniently

46
placed horizontal time line (‘α ’ in Figure 3.10). For instance, ∆T1 / 2 in the figure is a negative
difference while ∆Tn / 2 is positive. The points thus obtained are connected by straight-line segments
and form a differential time-distance graph.

C
∆Tn / 2

A B
α

∆T1 / 2 ∆Tn

∆T1 C

A
B

1 Distance n

Figure 3.10 Mean-minus-T method for estimating velocity in refractors

The inverse slope of each segment in this graph gives the velocity Vmn over that portion of the
refractor. For, if ∆Tk and ∆Tk-1 are the plotted time differences at two adjacent geophones separated
by a distance ∆x, the inverse slope of the segment is

1 / Vmn,2 = (1 / ∆x) [(∆Tk / 2) – (∆Tk-1 / 2)]


= (1 / ∆x)[((TB,k – TB,k-1) / 2) + ((TA,k-1 – TA,k) / 2)

= (1/2)((1 / VB) + (1 / VA))

where VB and VA are simply the forward and reverse velocities over the segment (up-dip and down-dip
ones if the segment is dipping). The refractor velocity over the segment ∆x is therefore
ms
Vmn,2 = VBVA / (VB + VA) / 2 (3.20) ξ 5500 ms-1
30

which is the true velocity if there is no dip, and in the 25


case of local dip it estimates the refractor velocity 3400 ms-1
20
over the segment in question to the same accuracy 5100 ms-1
as equation (3.18). 15
An application of this method is shown in Figure
10
3.11. The low-velocity segments in the central
portion of the mean-minus-T curve indicate a fracture 5
zone, the edges of which are given very nearly by the
0
intersections of the segments. Moreover, the 75 100 125 150 175 m
velocities in the fresh bedrock on either side of the 1100 - 1500 ms -1

zone appear to be slightly different from each other. 5100 ms-1 5500 ms-1
The MMT curve should normally be the first step in fractured rock 3400 ms-1
interpretation as it provides the refractor velocities
needed for adjustments to the t-d graph and Figure 3.11 Mapping fracture zone using
subsequent depth calculations. mean-minus-T method

47
4
Prelude to well logging
Radioactivity, electrical and
electromagnetic methods

4.1 Introduction

U p to now, the discussion of geophysical theory has centered around the bird-dog’s understanding
of basic wave theory in order to better perform his duties in reflection and refraction seismic
acquisition. Many non-speculative seismic projects are anchored at some point to oil fields, that is, to
the geologic and geophysical data gathered from drilled oil wells, both producing and non-producing,
and from on-going active wells. This brings us in contact with the personnel of on-going drilling
operations. As we are already in the field, and likely working for the same client as the drillers, it is
not out of the question that some day we might be asked to assist in the well logging program.
Even if this is not the case, it can do us no harm to learn as much about the geophysical
techniques in well logging as possible especially since, as will be seen in Chapter 5, that much of the
information gathered at a well will be utilized later on in seismic data processing and interpretation.
Additionally, reflection seismic acquisition specialists (bird dogs) may be asked to supervise a
well site VSP (vertical seismic profiling) program (section 5.12). This procedure incorporates many of
the surface seismic methods already discussed and is to some extent a liaison between them and
bore hole geophysics.
The previous chapters dealt with the physical properties of elastic waves leading up to the
seismic methods of reflection and refraction. For large-scale investigations covering wide areas,
these time-honored and constantly improving methods are cost effective and very reliable within their
scope. They do however, rely heavily on human interaction in their application and eventual
processing and interpretation. Certain assumptions regarding amplitudes, velocities, frequency
content and depths are made by necessity. Considering the sum of influences on a seismic pulse as
it travels to its reflection destination and returns to the surface, such as noise, convolution, signal
attenuation, diffraction, dip and others, it is obvious that the reliability of such data is greatly
dependent upon the processing and interpretation phases. The fact that these methods have proven
so successful for finding oil over the years is a tribute to the many geophysicists who have dedicated
their professional lives to improving the seismic method.
In contrast to the inherent limitations of the seismic method, the information gathered from well
logs is considered extremely reliable. By geophysical well logging, the physical properties of rocks,
ground water or disposed material are determined. The findings are valid only for the direct vicinity of
the borehole. Ideally every well drilled is logged in order to provide the best key to the proper
interpretation of geophysical field data and helps in the recognition and understanding of the
contamination of the geophysical data by previously discussed means: noise, absorption, spherical
divergence, geometric spreading, multiples etc.

4.2 The science of well logging

T he logging methods discussed in the next chapter and listed in Table C.1 in Appendix C can be
placed generally in five categories according to their respective science; they are: radioactivity,

48
electrical, electromagnetic, sound wave transmission, and mechanical. Falling in-between the
electrical and the electromagnetic methods is induced polarization (IP), which is based on an
electrochemical reaction of a natural liquid solution with interfaces of minerals and measured by
means of the induction of an electrical current. As IP relates to some logging methods, it will be
discussed briefly; however, its application is found more in surface or environmental geophysics such
as mineral prospecting, waste dump surveys, ground water investigations etc.
Mechanical applications include caliper for borehole diameter measurements, temperature for
thermal gradients, flowmeter for locating and measuring water inflow and outflow and deviation for
measuring the borehole axis deviation from vertical and its azimuth relative to magnetic north.
The method of soundwave transmission for seismic velocity studies will be discussed in Chapter
5, sections 5.2 and 5.8.3.
We will find that some of the theories and applications of seismic reflection and refraction have
analogs in the disciplines of radioactivity, electrical and electromagnetic methods. For example, the
signal attenuation discussed in section 1.7 can be plotted on a signal decay curve, which is
analogous to radioactive decay. For another, the Fourier transform (Appendix F) used to convert
time-domain reflection data into frequency and phase components is also used in the geophysical
methods of induced polarization and electromagnetics. Additionally, very low frequency (VLF) and
ground penetrating radar (GPR) signals, in some respects, behave similarly to sonic signals.
A broad and detailed discussion of radioactivity, electricity and electromagnetism is well beyond
the scope of this monograph. However, the basics of each theory will be presented as they relate to
well logging.

4.3 Radioactivity

T he geophysical methods employing radioactivity came into prominence with the demand for
uranium metal in atomic reactors. However, the methods are not restricted in scope only to the
search for the ores of radioactive metals or minerals associated with them. They can often be used
with advantage in geological and structural investigations such as the three well logs: gamma ray,
neutron and density listed above.
The radioactivity of rocks is now chiefly monitored using gamma ray scintillation counters and
spectrometers. Although most radiometric instruments were developed with the search for uranium in
mind, other uses were soon found. Among these were regional geological mapping and correlation,
exploration for some industrial minerals and in situ determinations of phosphates. The same
instruments may also be used to track the movement of ‘artificial radioactive tracers’ deliberately
introduced into ground water and to assess the health risks from natural and artificial radiation
sources.
In recent years, there has been concern about the radiation dose to which the population at large
may be subjected due to indoor concentrations of the radioactive noble gas radon generated in some
rocks and soils. Subsequently, systematic radon surveys of building sites have become common in
many countries. Radon detection is discussed in section 4.3.12.

4.3.1 Types of radiation

C hemical elements that possess more than 84 protons (atomic number Z > 84) are called
radionuclids or radioatoms. They are unstable, i.e. their atomic nuclei can disintegrate
spontaneously under the emittance of radioactivity. During this process, the number of protons of the
nuclei is changed and another chemical element is born. In addition, energy is produced, which is
measured in electron-volts (eV).

1 eV = 1.602 x 10-19 J (joule)

and 1 eV = 4.45 x 10-26 kWh (kilowatt-hours)

The nucleus of an element X with an atomic number Z has a positive charge of Z (atomic) units
and is made up of nucleons (protons and neutrons). The number of nucleons is the mass number A
of the element and the nucleus is symbolically as AZ X. Elements with the same Z but different A are

49
said to be isotopes of each other.
The spontaneous decay or disintegration of certain nuclei produces alpha (α), beta (β) and
gamma (γ) rays. This is the phenomenon of radioactivity. The α rays are particles of the helium
nuclei 2 He and the β rays are particles consisting of electrons and positrons. These emissions alter
4

the nuclear charge as follows: α by –2, β+ (positron) by –1 and β- (electron) by +1. This means that
the disintegrating nucleus is transformed into a nucleus of another element. Very often the resulting
daughter nucleus is also radioactive in it turn. Gamma rays are high-energy electromagnetic waves
that, according to quantum theory, can be treated as if composed of particles.

4.3.2 Alpha and beta particles

A n alpha particle consists of two protons held together by two neutrons to form a stable helium
nucleus. The emission of α particles is the main process in radioactive decay, resulting in a
decrease of four in atomic mass and a decrease of two in atomic number. The particles have large
kinetic energies but are rapidly slowed down by collisions with other atomic nuclei. The α particles
ionize other molecules on their direct way and lose their energy fast. They are not even able to go
through a sheet of paper. As thermal energies they soon gain two orbital electrons and become
indistinguishable from other helium atoms. The average distance traveled in solid rock before this
occurs is measured in fractions of a millimeter.
Beta particles are electrons and positrons ejected from the atomic nuclei. They differ from other
electrons only in having higher kinetic energies and so cease to be identifiable after being slowed
down by multiple collisions. Energy is lost most rapidly in collisions with other electrons.
Because of its lower atomic mass, the β radiation has a higher velocity than do the α particles.
For the same reason, the path that the β ray travels is not straight but zig-zag. Only 3 mm of
aluminum or or 15 cm of clay is enough to shield this radioactivity. In solids or liquids, the average
range of a β particle is measured in centimeters.

4.3.3 Gamma rays

T he nucleus is generally in an excited state after a β emission and returns to its ground state with
the emission of an additional particle, the gamma ray or quantum. In some rare instances an α
emission too is followed by a γ ray, e.g. in radium. The γ ray is a purely electromagnetic radiation
which does not alter the nuclear charge.
According to quantum theory, electromagnetic radiation consists of discrete particles or quanta,
namely photons. The energy of a photon is proportional to the frequency which can be in excess of
about 1019 Hz. The energy is given by hc / λ, where h is Planck’s constant, c = 3 x 108 ms-1, the
speed of light, and λ is the wavelength of the radiation. The energy is commonly expressed in
electron volts (eV). 1 eV, defined in section 4.4.1 as equal to 1.602 x 10-19 J, is the energy acquired
by an electron in falling through a potential difference of one volt. Since h = 6.625 x 10-34 J s which is
equal to 4.14 x 10-15 eV s, the energy of a photon is 1.24 x 10-6 / λeV. The wavelength of γ rays are
of the order of 10-11 to 10-12 m (0.1 to 0.01 Å) and the corresponding energies are of the order of from
0.1 to 1 MeV.
Because they are electrically neutral, photons penetrate much greater thicknesses of rock than
do either α or β particles ( a lead shield of > 30 cm or a clay layer of at least 3 m is required to stop
their penetration) and are consequently the most geophysically useful form of radiation. Even so,
approximately 90% of the radiation detected over bare rock during environmental surveys will come
from within 20 to 30 cm of the surface and even though soil, only 10% will come from below 50 cm.
On the other hand, 100 m of air will only absorb about half of a γ ray flux, so that atmospheric
absorption can generally be ignored.

4.3.4 Law of radioactivity and half-life

Gamma rays are important because they provide information on the presence of unstable atomic
nuclei. Decay is statistical; the average number of decays in a given time will be directly

50
proportional to the number of atoms of the unstable element present. The disintegration of a given
quantity of any radioactive element can be expressed by the formula N = N0-λt where N0 is the number
of nuclei initially present and N the number remaining after a time t. Here λ is known as the decay
constant, and after a time 1/λ the number of nuclei present is reduced to 1/e ≈ 1/3 of the initial
number.
A given quantity of a radioactive element is halved after a time T = (ln 2) / λ = 0.693 / λ. The rate
of decrease in mass of a radioactive material therefore is governed by its half-life during which half of
the original radioactive material is lost. Figure 4.1 graphs the exponential law of radioactive decay.

Exponential Decay:
1.0
m = m0e-λτ Half life = t1/2 = ln 2 / λ
Mass or signal strength

Attenuation:
l = l0e-αd Skin depth = δ1/e = 1 / α

0.5
1/e = 0.368

0.25

0.125
δ1/e

t1/2 2t1/2 3t1/2


Depth or Time

Figure 4.1 The exponential law with parameters for radioactive decay
and radio wave attenuation

R adioactive particle fluxes and seismic and electromagnetic waves are subject to absorption as
well as geometrical attenuation (section 1.7). The energy crossing closed surfaces is then less
than the energy emitted by the enclosed sources. In homogeneous media, the path length and the
attenuation constant determine the percentage loss of signal. The absolute loss is also proportional
to the signal strength. A similar exponential law (Figure 4.1), governed by a decay constant,
determines the rate of loss of mass by a radioactive substance.
Attenuation rates are also described by skin depths, which are the reciprocal of the attenuation
constant. In one skin depth, the signal strength decreases to 1/e of its original value, where e ( equal
to 2.718) is the base of natural logarithms. The radioactivity equivalent is the half-life. This is equal
to ln 2 ( = 0.693) divided by the decay constant λ.

4.3.5 Radioactive decay series

E lements with short half-lives can occur in nature because they are formed in decay series which
originate with very long-lived isotopes, sometimes termed primeval. About 50 natural and 800
artificial radioactive nuclei are known. The principle primeval isotopes are uranium-238, uranium-235,
thorium-232 and potassium-40 (238U, 235U, 232Th and 40K respectively) which account for the principal
radioactive decay series found in nature. Also of interest is neptunium-239 (239Np) an artificial
product with a half-life of 2.355 days that decays and produces 237Np. The latter is the longest-lived
member of the neptunium series with a half-life of 2.14 x 106 years, but this is too short compared to
the age of the earth by a factor of about 2000, so that Np does not occur naturally on the earth. The
radioactivity of Np is natural in the sense that the element has been detected by astrophysical

51
measurements in stars. The neptunium series has been observed in the laboratory in Np produced in
atomic reactors.
Of all the naturally occurring radioactive elements with an atomic number less than lead, the most
9
important is the isotope 4019 K of potassium with a half-life of 1.25 x 10 years. About 89% of 40K is
transformed with the emission of a β particle into calcium-40 and about 11% with the capture of an
electron from the K shell of potassium into argon-40. The transformation is accompanied by high-
energy γ radiation of a wavelength of about 0.08 pm.
40
Table 4.1(a) K decay series

Parent Element Symbol Z Emission Half-life Daughter γ energy in MeV (% yield)


89%−β
9 40
40 1.45 x 10 y Ca
Potassium-40 K 19 10 40 1.46(11)
11%−α 1.17 x 10 y Ar

Others, such as 48Ca, 50V and 58Ni, are either rare or only very weakly radioactive. Radioactive
elements are mainly concentrated in acid igneous rocks and in sediments that were deposited in
reducing environments.
238
Table 4.1(b) Natural radioactive decay, U series

Parent Element Symbol Z Emission Half-life Daughter γ energy in MeV (% yield)

α
238 9 234
Uranium-238 U 92 4.51 x 10 y Th 0.09(15) 0.67(7) 0.3(7)

β, γ
234 234
Thorium-234 Th 90 2454.1 d Pa 1.01(2) 0.77(1) 0.04(3)
234 β, γ 1.14 m 234 0.05(28)
Protactinium-234 Pa 91 U
β 6.7 h
α
234 5 230
Uranium-234 U 92 2.52 x 10 y Th

α
230 226
Thorium-230 Th 90 80,000 y Ra

α
226 222
Radium-226 Ra 88 1,622 y Rn 0.19(4)

α
222 218
Radon-222 Rn 86 3.825 d Po
214
Pb
α, β
218
Polonium-218 Po 84 3.05 m 218
At
0.35(44) 0.24(11) 0.29(24)
β, γ
214 214
Lead-214 Pb 82 26.8 m Bi
0.05(2)
α
218 214
Astatine-218 At 85 2s Bi
214
Po 2.43(2) 2.20(6) 1.76(19)
β, α, γ
214
Bizmuth-214 Bi 83 19.7 m 210
Tl 1.38(7) 1.24(7)
α
214 -4 210
Polonium-214 Po 84 1.6 x 10 s Pb

β, γ
210 210
Thalium-210 Tl 81 1.3 m Pb

β, γ
210 210
Lead-210 Pb 82 19.4 y Bi
210
Po 0.04(4)
β, α
210
Bizmuth-210 Bi 83 5.0 d 206
Tl
α
210 206
Polonium-210 Po 84 138.4 d Pb Stable

β
206 206
Thallium-206 Tl 81 4.2 m Pb Stable

The transformations of the 40K, 238U, and 232Th series are shown in Tables 4.1(a), (b) and(c). The
40
uranium and thorium series end in the stable isotopes of lead: 206 208
82 Pb, and 82 Pb respectively. The K
series, which forms about 0.0118% of naturally occuring potassium, decays in a single stage, either
by β emission to form 40Ca, or by electron capture (K-capture) to form 40Ar. The argon nucleus is left

52
in an excited state, but settles down with the emission of a 1.46 MeV photon. The neptunium series
238
ends in the bismuth isotope 20983 Bi. Figure 4.2 illustrates the U decay to stable lead 206Pb.
He

214
Po He
He -
He
He - 210
Pb 210
Bi
He
218
At
214
Bi -
234 230 He
He U Th
238
U - 210 206
218 214 Po Tl
He Po Pb
234
Pa
- 210

- - Tl
234
Th
222
Rn -
- 206
Pb

226
Ra He

He

238
Figure 4.2 U radioactive decay series

In the 238U chain, 214Bi is notable for the numbers and energies of the γ photons produced. Those
at 1.76 MeV are taken as diagnostic of the presence of uranium, but the gaseous radon isotope
222
Rn, which precedes 214Bi in the chain, has a half-life of nearly four days and so can disperse fairly
widely away from a primary uranium source.
By no means are all decay events accompanied by significant γ emission. The first stage in the
decay of 232Th (Table 4.1 (c)) involves only weak gamma activity. The strongest radiation in the chain
comes from the decay of 208Tl, near the end. This decay is accompanied by the emission of a 2.615
MeV photon, the most energetic radiation to come from a terrestial source
Uranium-235 (235U) makes up only 0.7114% of the naturally occurring element and, although its
rather greater activity allows it to contribute nearly 5% of the overall uranium activity, it may be
ignored in discussions of geophysical radiation detection methods.
232
Table 4.1(c) Th decay series

Parent Element Symbol Z Emission Half-life Daughter γ energy in MeV (% yield)

α
232 10 228
Thorium-232 Th 90 1.4 x 10 y Ra 0.06(24)

β
228 228
Radium-228 Ra 88 6.7 y Ac
1.64(13) 1.59(12) 0.99(25)
β
228 228
Actinium-228 Ac 89 6.1 h Th
0.97(18) 0.34(11)
α
228 224
Thorium-228 Th 90 1.9 y Ra

α
224 220
Radium-224 Ra 88 3.64 d Rn

α
220 216
Radon-220 Rn 86 54.5 s Po

α
216 212
Polonium-216 Po 84 0.16 s Pb
0.30(5) 0.24(82) 0.18(1)
β
212 212
Lead-212 Pb 82 10.6 h Bi
0.12(2)
212
212 66%−β 40 m Po 1.18(1) 0.83(8) 0.73(10)
Bizmuth-212 Bi 83 208
34%−α 97.3 m Tl
α
212 -6 208
Polonium-212 Po 84 0.3 x 10 s Pb 2.62(100) 0.86(14) 0.58(83)

β
208 208
Thallium-208 Tl 81 3.1 m Pb 0.51(25)

53
4.3.6 Radioactive equilibrium

I f a large amount of a primeval isotope is present and if all the daughter products remain where they
are formed, an equilibrium will eventually be established in which the same number of atoms of
each element are created in a given time as decay. In other words, radioactive series, such as those
represented in Table 4.1, are said to be in equilibrium when as many atoms of any unstable element
in it are being formed per second as are disintegrating. Only the concentrations of the two end-
members of the series change.
In equilibrium decay, each member of the chain loses mass at the same rate, equal in each
instance to the mass of the element present multiplied by the appropriate decay constant.
Equilibrium masses are therefore inversely proportional to decay constants. If more (less) of an
element is present than is required for equilibrium, decay will be faster (slower) than the equilibrium
rate until equilibrium is re-established.
Equilibrium will be disrupted if the half-lives of any gaseous or soluble intermediate products are
long enough to allow them to be removed before they decay. The exhalation of radon by uranium
ores notably disrupts equilibrium and the primary source of a ‘uranium’ anomaly (actually due to 214Bi)
may be hard to find. Roll-front uranium ores are notorious for the separation between the uranium
concentrations and the zones of peak radioactivity.

4.3.7 Natural gamma ray spectrum

A radioactive series in equilibrium will emit a definite spectrum of γ rays. It is customary to describe
the spectrum in terms of energy levels, instead of wavelengths as is the case in the spectrum of
ordinary light.
Natural γ rays range from cosmic radiation with energies above 3 MeV down to X-rays. The γ
spectrum of 238U contains lines at various levels from about 0.1 MeV to about 2.4 MeV. The spectrum
of 40K consists of a single line at 1.46 MeV. A spectral line is never perfectly sharp but has a finite,
although usually small, energy width.
A typical measured spectrum is
shown in Figure 4.3. The individual
peaks correspond to specific decay
events. 10,000
The energy of each photon falls
somewhere within a small range
determined by the nuclear kinetic
energies at the time of the decay and by
1000
errors in measurement.
Counts

The background curve on which the


peaks are superimposed is due to
scattered terrestrial and cosmic (mainly K (1.46 MeV)
solar) radiation. Gamma photons can 100
be scattered in three ways. Very
energetic cosmic photons passing close
U (1.76 MeV)
to atomic nuclei may form electromag-
netic pairs. The positrons soon interact 10 Th
with other electrons to produce more γ (2.615 MeV)
rays. At lower energies, a γ ray may
eject a bound electron from an atom,
some of the energy being transferred to
the electron and the remainder 1.0 2.0
proceeding as a lower energy photon Energy (MeV)
(Compton scattering). Alternatively, a
photon may be totally absorbed in
ejecting an electron from an atom
(photoelectric effect). Figure 4.3 A natural gamma-ray spectrum

54
If the whole test sample is in radioactive equilibrium with respect to both parent and daughter and
if energy lines from other nuclides or other radioactive series are not overlapping the measured line,
the content of the parent element of the series (U,Th) can be assesed by comparison with a standard
of known content. Even in less than ideal measuring conditions, by postulating radioactive
equilibrium, the measured intensities can still be converted to U or Th content, which is referred to as
equivalent content (eU, eTh).

4.3.8 Absorption of α, β and γ rays

Tare
he α and β particles lose their energy in passing through matter by collisions, ionization, etc., and
brought to a virtual stop within a certain distance which is called their range. In air at 18º C
the range of α particles is only a few centimeters; in denser substances, e.g. mica or aluminum, it is
still smaller, being of the order 30 µm. Even the β particles, whose range is several hundred times
greater, are completely stopped by a thin sheet of lead or for example, a few centimeters of sand.
The intensity of γ rays in traversing matter decreases exponentially with distance so the term
definite range may not apply in this case. Theoretically, γ rays could be detected across any
thickness of matter but a practical limit is set by the sensitivity of the detecting instruments and the
background effects due to cosmic radiation. Besides the purely electromagnetic effects limiting the
depth penetration of γ rays in matter, there are three principal mechanisms by which they are
absorbed. One is the photoelectric effect in which a γ ray gives up its energy to an electron in an
atomic orbit and raises it to a higher energy level. The second is the Compton effect in which a
knocked-off electron absorbs part of the γ ray energy and the remaining energy is converted into a γ
quantum of longer wavelength than the original one. The third mechanism annihilates the γ quantum
by the producing of an electron-positron pair. For practical geophysical purposes γ radiation may be
taken to be entirely absorbed by 1 to 2 meters of rock.
Evidently, the α and β particles are not of much avail in geophysical field work since they will be
undetectable as soon as a radioactive deposit has the thinnest cover of overburden. Thus, the
search for radioactive minerals is, to a large extent, a search for places with abnormally high gamma
radiation. Uranium, for example, is located indirectly from the powerful γ radiation emitted by two
ii
products of the uranium series: 20983 Pb and
209
83 Bi.

4.3.9 Radioactivity of rocks

M inute traces of radioactive minerals are present in all igneous and sedimentary rocks, in oceans,
rivers and springs, in oil, and in peat and humus. The average amounts of U, Th, and 40K in a
few materials of the earth’s crust are shown in Table 4.2

Table 4.2 Radioactive contents of some natural materials

Material U (gt-1) Th (gt-1) 40K (%)


Basalt 0.9 4.2 0.75
Diabase 0.8 2.0
Granite 3-5 13.0 4.4
Sediments ~ 1.0 5-12
Limestone 1.3 1.1
Oil 100
Ocean water 0.00015-0.0016% 0.0005%
Uraninite 38-80 0-0.3
Carnotite 50-63 (UO2) 0-15
Thorianite 23-26 53-57
Monazite 0.02-0.7 3.5-16.5

55
Table 4.2 indicates that there is a large difference between the radioactivities of basalts and
granites; moreover, the latter have a remarkably high content of 40K. This fact is of consequence
because granites are very common rocks and the γ radiation from their potassium produces a
radioactive background that may make it difficult to locate uranium and thorium ores. Sometimes the
radioactivity of potassic feldspars in pegmatite may be misinterpreted as being due to a concentration
of uranium and thorium. In other cases, it can be of help in classifying granites of different types
within a certain area. Also of note is the high content of uranium but low content of thorium in oil.

4.3.10 Radiation detection and measurement

T he α, β and γ radiations are detected by their ionizing action. In geophysical work primarily the γ
ray detection is considered because the α and β particles are stopped by matter. The earliest
radiometric instruments relied on the ability of radiation to ionize low-pressure gas (usually argon with
a small amount of alcohol or amyl acetate) and initiate electrical discharges between electrodes
maintained at a high potential difference.
When a γ ray passes through the gas it produces ions that are accelerated by the field and
produce further ions. The current passing through the tube can be amplified and registered on a
meter or heard as a click in a pair of headphones. These Geiger-Müller counters responded mainly
to alpha particles and suffered long ‘dead’ periods after each count, during which no new events
could be detected. Only 1% or less of the incident γ rays are responded to by the Geiger counters.
A more efficient type of detector is the scintillation counter. Gamma rays produce flashes of
light when they are photoelectrically absorbed in sodium iodide crystals. Small amounts of thallium
are added to the crystals, which are then termed thallium activated. The produced light can be
detected by photomultiplier tubes (PMT) that convert the energy into electrical currents.

Scattered (lower energy γ ray

dynode dynode anode


To counter

Scintillation
crystal dynode dynode

Incoming γ ray
photo electron electron avalanche
cathode
α and β ray absorber

Figure 4.4 Diagram of an photomultiplier tube (electron multiplier)

As shown in Figure 4.4, the scintillations are detected by a photo-cathode that releases electrons
by photoelectric effect. The electrons are made to strike a number of secondary positive electrodes,
called dynodes, in an evacuated tube and accelerate in a zig-zag manner between the plates,
producing secondary electrons in the process. Although the device is called a photomultiplier, it is
actually an electron multiplier. The initial emission of the electrons at the photo-cathode turns at the
other end into an avalanche that delivers a current to an external circuit. The current is recorded
suitably after amplification and can be correlated after calibration of the meter with the number of
incoming γ photons, or scintillations, since each γ photon produces one scintillation. The number of
scintillations for a pre-set time is counted and the readings are displayed as counts recorded per
second (CPS) or per minute (CPM).
Scintillation counters are almost 100% efficient in detecting γ rays. Their sensitivity to cosmic
rays is about the same as that of Geiger counters so that their relative response to γ rays is much
higher. The sensitivity depends almost entirely on the crystal size; larger crystals record more
events. Count rates are thus not absolute, but depend on the instrument and crystal. Similar
instruments with similar crystals should read roughly the same in the same places.

56
I f a pulse-height analyzer is incorporated into the PMT circuitry, the energy of each γ photon which
produces a scintillation event can be estimated. Events with energies within certain predetermined
energy windows or above preselected energy thresholds can be counted separately.
The differential spectrometer records only radiation falling within predetermined upper and
lower energy limits. If the limits are very close together (e.g. < 30 keV) the spectrometer is said to
respond to a channel or a line, while for wider separation of limits (several hundred keV) it is said to
respond to a window. The entire γ ray flux can be observed at a series of narrow adjoining windows
to obtain a curve similar to that shown in Figure 4.3.
The integral spectrometer is set to exclude radiation below a predetermined energy level, the
threshold, and records all radiation having photon energies greater than this level. In some
instruments the threshold can be varied. There also exist mixed-mode instruments. The choice of
the spectrometer will depend to a large extent on the purpose and the requirements of the survey.
Strictly speaking, the term spectrometer should be reserved for those instruments with 256 or
more channels which can record a complete spectrum, but in practice it is applied to any multichannel
instrument with some degree of energy discrimination. Usually there are only four channels, one for
the total count and one each for the 208Tl peak at 2.62 MeV, the 214Bi peak at 1.76 MeV and the 40K
peak at 1.46 MeV. Typical windows for these peaks might extend from 2.42 to 2.82 MeV, from 1.66
to 1.86 MeV and from 1.36 to 1.56 MeV, respectively. The concentrations of all three parent
elements can thus be estimated, although care is needed in recognizing disequilibria.
The intensities are mostly recorded as counts per minute but many instruments are calibrated in
micro- or milliroentgens per hour. The roentgen is the quantity of γ (or X) radiation that produces
2.083 x 109 pairs of ions per cm3 of air at NTP*. It corresponds to an energy absorption of 8.38 x 10-6
joules per kilogram of air. Several different measurement units are in use in radioactivity work
depending upon the purpose for which the measurement of radioactivity is carried out. Table 4.3 lists
the basic ones. The gray, rad and sievert are essentially for radiological work in monitoring health
hazards. The geophysicist will be primarily dealing in the curie, roentgen or most commonly the
becquerel, named after the French physicist who received the Nobel Prize in 1896 for the detection of
radioactivity. One becquerel corresponds to one radioactive decay per second. Formerly the curie
was used: 1 Ci = 3.7 x 1010 Bq.

Table 4.3 Radiation units of measurement

Unit Sl unit Quantity measured


becquerel (Bq) 1 Bq = 1 decay / s = 1 Hz Activity as number of disintegration
events (decays) per second
curie (Ci) 1 Ci = 3.7 x 1010 Bq Activity of 1 g radium

-6 -1
roentgen (R) 1 R = 8.38 x 10 Jkg (of air) Quantity of X-radiation producing
2.083 x 109 ion pairs per cm3 of air*
-1
gray (Gy) 1 Gy = 1Jkg = 100 rad Dose of ionizing radiation per kg

-2
rad 1 rad = 10 Gy Old unit of energy dose

sievert (Sv) 1 Sv = 1Jkg-1 = 100 rem (old unit) Equivalent dose

4.3.11 Stripping ratios, intensity factors and absorption coefficients

T o estimate thorium, uranium and potassium-40 abundances from spectrometer readings,


corrections must be made for gamma rays scattered from other parts of the spectrum. The
thorium peak must be corrected for cosmic radiation and for the 2.43 MeV radiation from 214Bi in the
uranium decay chain, which overlaps into the commonly used ‘thorium window’. The uranium count
must be corrected for thorium and the potassium-40 count must be corrected for thorium and
uranium. The correction process is known as stripping.

57
Stripping factors vary from detector to detector, primarily with variations in crystal size. They are
listed in instrument handbooks and in some instances can be applied by built-in circuitry so that the
corrected results can be displayed directly. Equilibrium is assumed and interpretational errors will be
made if it does not exist. It is generally better to record actual count rates and make corrections later.

T he intensities recorded in geophysical work are the integrated effects of the radiation from a finite
area on the ground. In foot surveys most of the intensity comes from within a circle of about 3 m
radius, while in car-borne surveys about 90% of it comes (because of the greater detector height)
from within a radial distance of 15-20 m.
Topographic irregularities, absorption and scattering of radiation in the earth materials, the
dispersion or concentration of radioactive materials due to weathering and subsequent transport, etc.,
and the ‘background radiation’ are some of the factors affecting instrument readings. These must be
carefully considered in the interpretation.
The background radiation is due mainly to cosmic rays, potassium-40 and the minute quantities
of uranium and thorium which are almost always present in rocks. It may vary from one area to
another as well as within a single area. A count-rate observation cannot be considered significant in
general unless it is 3-4 times the background rate.
The intensity of light emitted by a point source decreases inversely as the square of the distance
from the source. The γ radiation that is measured in radioactive surveys, being an electromagnetic
radiation, follows the same geometric law of intensity decrease, but as it is absorbed in traversing
through matter by the mechanisms mentioned in section 4.3.8 there is a further decrease by an
exponential factor. Assuming that the part of the ground through which the radiation traverses is
homogeneous, the intensity I at a distance r from a point source can be expressed as

I = (C / r2) e-µr (4.1) ξ

where C is a constant that depends on the amount of radioactive material in the source and the
channel on which the observation is made, and µ is an absorption coefficient. The absorption
coefficient depends upon the energy of the γ radiation as well as on the density ρ (kgm-3) of the
traversed matter. For γ radiation from uranium in equilibrium with its daughter products, µ is found
empirically to be 3.3 x 10-6 ρ m-1.

4.3.12 Radon measurement and α particle monitors

R adium-226, formed in the 238U decay series, has as its disintegration product radon-222, a radio-
active gas that emits α particles and has a half-life of 3.825 days. This gas may seep towards the
surface with currents of water, diffuse through permeable rocks or escape through fissures and
cracks. In practical use the term ‘radon’ refers to the isotope 222Rn. There exist two other natural
isotopes: 220Rn and 219Rn, which are decay products in the 232Th and 235U series respectively, but
compared to 222Rn, have very short half-lives of 3.96 seconds and 55.6 seconds.
Radon transport in the ground was for a long time attributed primarily to diffusion. However,
diffusion of a gas through the pores of soils and rocks is a slow process, and due to its half-life of only
3.825 days, 222Rn cannot be carried by diffusion further than about 5 to 10 meters from the source
before radioactive disintegration reduces its concentration to levels indistinguishable from the
background. No anomalies would be found that could reveal even a strong source. Yet, radon
transport over distances of hundreds of meters from the source is well documented. Also, radon
observations in connection with earthquakes are difficult to reconcile with the diffusion theory.
A plausible dominant process for radon transport over long distances was first suggested by two
Swedish geophysicists, Kristiansson and Malmqvist (1982). According to them, α particles produced
in the decay of radioactive minerals in the earth’s crust initiate bubbles in the groundwater at depth.
The groundwater is supersaturated with dissolved atmospheric gases so that any radon atoms, being
unable to go into solution, attach themselves to the bubbles. The bubbles ascend because of their
buoyancy through cracks and fissures, remaining as free bubbles, and in this way can carry 222Rn to
long distances.
Radon, although its presence may indicate other radioactive material buried below it, does not

58
itself emit γ rays in its disintegration so that portable scintillometers are not of much avail in measuring
radon activity and other procedures have to be followed in systematic radon surveys.
One procedure is to thrust a small tube, about 50 cm long and 6-10 mm in diameter, into the soil
to a depth of about 1 meter. A sample of soil air is drawn by means of a hand-pump into a container
connected to the tube (sometimes after passing a water trap) and is analyzed for it’s α activity in an
ionization chamber. The collection of one air sample may take some 20-30 minutes with an
additional several minutes for a semi-quantitative assay.ii
In another method known as the alpha-track method, small pieces of cellulose nitrate films are
attached to the bottom of small plastic cups and the cups are placed with the open end downward at
the desired depth in the soil at observation points in a network. They are left there for a period of 2-3
weeks, after which the film is retrieved and etched in NaOH solution to enhance the tracks made by
the α particles. The density of tracks (number per cm2) can then be determined by counting the
tracks under a microscope.ii
A similar method utilizes the alphacard monitor, a thin metallic membrane mounted on a frame
about the size of a 35 mm slide. The alphacard is suspended in an inverted flowerpot or similar
container (remembering from section 4.3.2 that a sheet of paper is sufficient to block α rays) in a hole
about 0.5 to 1 m deep. The hole is covered is covered in such a manner as to be able to retrieve the
card after at least 12 hours. This is long enough for equilibrium to be established and longer periods
will not alter the reading. After the card has been removed, it is placed in a special reader that is
sensitive to α radiation.i
Radon as such is not a poisonous gas but its disintegration products, radon daughters, pose a
potential health hazard. They can settle on dust and other particles that are present in the air we
breath and enter the respiratory system, where the α emmissions in the disintegration can damage
lung tissue and lead to lung cancer. Fortunately, good ventilation is effective in reducing radon
concentration in buildings and houses to acceptable levels. The recommended limit in countries of
the European Community for existing buildings is 400 Bq m-3 (yearly average).ii

4.3.13 Radioactive density determinations

T he method of γ ray absorption has been employed for some time in determining soil density in
foundation investigations and hydrological problems. For this purpose an artificial radioactive
isotope such as 60Co (t1/2 = 5.26 y) or 137Cs (t1/2 = 33 y) is used. The former emits γ rays during β
disintegration; the latter changes to excited 137Ba which in turn emits γ rays in returning to a ground
state.

10,000

Sand (ρ = 1.89)
µ = 0.121 cm
· counts / min)

-1

1000
2
r N (m

Clay (ρ = 2.10)
µ = 0.137 cm
-1
2

100
0.5 0.8 r (m)

Figure 4.5 Density determinations by radioactivity. Densities (ρ) are in g cm-3

A metal probe with the γ source at the end is driven vertically to a depth (z) of about 0.5 to 1
meter. The count of a detector placed on the surface at various horizontal distances (x) along a line
is noted for a fixed time. The soil will absorb γ rays so that the count will depend upon its density ρ.

59
If the detector is assumed to be very small it can be shown that

N = (N0 / 4πr2)µ e-µr (4.2) ξ

where N0 is the total number of γ quanta emitted, N is the number reaching the detector per unit
detector volume, r = √ (x2 + z2) is the source-receiver distance and µ is the absorption coefficient.
The plot of ln(r2N) against r will evidently be a straight line with slope µ. It can be proved however
that µ = µ’r where µ’ depends only upon the energy of the radiation. The mean values of 1/µ’ are 18.4
g cm-2 for 60Co and 14.2 g cm-2 for 137Cs but are affected slightly by test conditions. For the examples
shown in Figure 4.5 using 137Cs, the values are 1/µ’ = 15.33 g cm-2 (clay) and 15.65 g cm-2 (sand).
Density values accurate to 1% or better can be obtained with this method whose chief advantage
is that the density can be determined in situ, without having to take samples.

4.4 Electrical methods

A number of geophysical methods use measurements of voltages or magnetic fields associated


with electric currents flowing in the ground. The currents may be natural, but are more often
produced artificially by direct contact or electromagnetic induction. Broadly speaking, the electrical
methods exploit the flow of a steady current in the ground, either direct or sinusoidal, while the
electromagnetic ones rely on the phenomenon of electromagnetic induction and the wave character
of the electromagnetic field. The two electrical methods to be discussed in section 4.4 are (a)
resistivity and (b) self-potential.

4.4.1 Resistivity

4.4.1.1 Earth resistivity

In the earth resistivity method a commutated direct or low-frequency alternating current is


introduced into the ground by means of two electrodes (metal stakes or suitably laid out bare wire)
connected to the terminals of a portable source of e.m.f. The resulting potential distribution on the
ground, mapped by means of two probes (metal stakes or, preferably, non-polarizable electrodes), is
capable of yielding information about the distribution of electrical resistivity below the surface. The
method has been used mainly in the search for water-bearing formations, in stratigraphic correlations
in the oil fields and in prospecting for conductive ore bodies. In recent years it has been used for
detection of fractures and cavities in the subsurface, for delineating archeological features and in
monitoring pollution in the ground.
Graphite and most metallic sulphides conduct electricity fairly efficiently by the flow of electrons,
but most rock-forming minerals are very poor conductors. Ground currents are therefore mainly
carried by ions in pore waters. Pure water is ionized to only a small extent and the electrical
conductivity of pore water depends on the presence of dissolved salts, mainly sodium chloride (Figure
D.1, Appendix D).

4.4.1.2 Ohm’s law

T he current (I) flowing through a conductor of limited diameter and length and of uniform cross-
section is defined by Ohm’s law as equal to the voltage (V) across the length of the conductor
divided by the resistance (R) of the conductor to the flow or

I=V/R ξ

where current is given in amperes or amps (A), voltage in volts (V) and resistance in ohms (Ω).
The symbol for voltage, sometimes written as U, dV or formally as E, is also referred to as voltage
drop, potential or potential difference. Since V is also used for velocity, a measurement used in
electron flow, and E for energy or electrical field, U seems to be less ambiguous. In this section, the
symbol U and the term potential will be used however, it implies a difference of potential or voltage

60
between the two ends of the conductor so that

I=U/R (4.3) ξ

4.4.1.3 Resistivity and conductivity

T he resistance R is directly proportional to the length L (m) of a conductor and inversely


proportional to its cross-section s (m2) so that

R = ρ (L / s) (4.4) ξ

where the constant of proportionality ρ is the resistivity of a conductor with measurable length and
cross-section and is measured in ohm meters (Ωm).
The distinction between ‘resistance’ and ‘resistivity’ then is a geometric one. The resistance is a
characteristic of a particular path of an electrical current where resistivity is the resistance of a unit
cube to current flowing between opposite face. The resistance of rectangular block of conductive
material (Figure 4.6) is therefore proportional to the resistivity and distance the current must flow and
inversely proportional to its cross sectional area.

L
s

Figure 4.6 Resistivity of a parallelpiped

Substituting the value for R in equation (4.4) into equation (4.3) we get

I
=-
1U (4.5) ξ
s ρL

The left-hand side of the equation is the current density j (that is, current per unit area of cross-
section,(A m-2) while – U / L on the right-hand side is the electric field E (V m-1) in the direction of the
current density vector (thus the negative sign which indicates that current flow is from high to low
potential, that is, in a direction opposite to that of the increase of potential or gradient). Therefore,

j=E/ρ or E=ρj (4.6) ξ

Alternatively

j = σE (4.7) ξ

where σ (= 1/ρ) is the conductivity (the reciprocal of resistivity) of the material measured in siemens

61
per meter (S m-1), also called mho per meter (Ω-1m-1), the word ‘mho’ being ‘ohm’ spelled backwards.
If we let the length L in equation (4.5) tend toward zero and consider the linear conductor as an
element of a homogeneous and isotropic medium, e.g. a homogeneous rock mass, then either of the
equations (4.6) or (4.7) expresses Ohm’s law for such a medium. In an isotropic medium σ and ρ
are independent of the direction of flow or restated, isotropic materials have the same resistivity and
conductivity in all directions. Most rocks are reasonably isotropic, but strongly laminated slates and
shales are more resistive and thereby less conductive across the laminations than parallel to them.

4.4.1.4 Point and buried current electrodes

A point electrode on the surface of a homogeneous isotropic earth extending to infinity in the
downward direction and having a resistivity ρ suggests a hemispherical shell of radius r and
thickness dr around the electrode (Figure 4.7 (a)). By symmetry, the current at any point on the shell
will be along the radius. If the total current passing through the electrode into the ground is I the
potential drop dV (=U) across the shell can be written, using equation (4.5)

dV = - Iρ dr / 2πr2

Integrating, we get for the potential at a distance r from a point current source

Iρ 1
U= +C
2π r

where C is an arbitrary constant. If U is chosen to be zero at r = ∞, then C = 0 and we get


Iρ 1
U= (4.8) ξ
2π r
For a current electrode in infinite space, 2π in equation (4.8) is replaced by 4π since the current
then passes across a spherical and not a hemispherical shell.

I A (+) B (-)

r
r r’

P
(a) (b)

Figure 4.7 Point electrodes on the surface of a homogeneous earth

In practice, there are two electrodes on the ground surface, one positive (A), sending current into
the ground, and the other negative (B), collecting the returning current (Figure 4.7 (b)). The total
potential at any point P in the ground will then be
Iρ 1 1 (4.9) ξ
U= -
2π r r’

where r and r’ are the distances of the two electrodes to a point P.

T he case of a buried current electrode in homogeneous isotropic earth is a slightly more general
situation than the one considered above. In finding the potential in this case the boundary

62
condition must be taken into account that there is no current density at the ground surface
perpendicular to the surface, that is, in the vertical or z direction. All the current at the surface in this
case flows in the horizontal direction (Figure 4.8)

Battery

Borehole
Surface

Current
lines
+
Vertical
section

Equipotential
lines

Horizontal
+ plane

Borehole

Figure 4.8 Buried point electrode

From equation (4.5) it is noted that U must satisfy the condition (1 / ρ)(∂U / ∂z) = 0 at the ground
surface. The expression that satisfies this condition and also yields the expression in equation (4.8)
when the electrode is on the ground surface is

Iρ 1 1
U= - (4.10) ξ
4π r1 r1’

where r1 and r1’ are the distances of the point P from the current electrode and its image in the ground
surface respectively. Figure 4.8 also shows the current and equipotential lines on a vertical plane
through the electrode. The three dimensional equipotential surfaces can be visualized by imagining
the figure revolved around an axis through the electrode and perpendicular to the ground surface
(sphere like). On the ground surface the equipotential lines will be circles centered around the
electrode epicenter. With two buried electrodes, a positive and a negative, the potential becomes

Iρ 1 1 1 1
U= + - - (4.11) ξ
4π r1 r1’ r2 r2’

63
where r2 and r2’ are the distances of P from the negative electrode and its image respectively.
The obvious method of measuring ground resistivity by passing a current between a pair of
grounded electrodes and measuring the voltage between them does not work because of contact
resistances which may amount to thousands of ohms. The problem can be bypassed if the voltage
measurement is made between a second pair of electrodes using a high-impedance voltmeter which
draws virtually no current. A geometric factor is needed to convert the readings obtained with such a
four-electrode array to resistivity. Contact resistances at current electrodes limit current flow but do
not affect resistivity calculations.

4.4.1.5 Resistivity of rocks and minerals

O n a non-homogeneous earth the measured potentials when current is introduced into the ground
will differ from those calculated under the assumption that the ground is homogeneous. The
magnitude of the electric anomalies on a non-homogeneous earth depends upon the resistivity
differences between different rocks.
The resistivity of rocks and minerals is an extremely variable property ranging from about 10-6 Ωm
for minerals such as graphite and pyrrhotite to more than 1012 Ωm for dry quartzite rocks. Most rocks
and minerals are insulators in the dry state. In Nature they almost always hold some interstitial water
with dissolved salts and therefore acquire an ionic conductivity which then depends upon the
moisture content, the nature of the electrolytes and the degree to which the open spaces in a rock
(pores, micro-fissures, cracks, fractures, etc.) are saturated with water. The form of the pores in a
rock plays a subordinate role in determining the conductivity but the degree of pore interconnection is
very important. In rocks like basalt or granite, for example, the conductivity can vary enormously
depending upon the degree of pore interconnection or lack of it, from thousands of Ωm in compact,
non-porous masses down to 10-50 Ωm in rocks like porous, water-bearing basalt in most tropical
areas. Water-bearing fracture zones in rocks usually have fairly low resistivities while the resistivity of
clays depends on the properties of clay materials and other material.

Table 4.4 Electrical resistivities of some rocks and minerals

Rocks and sediments Resistivities (Ωm) Ores Resistivities (Ωm)


Topsoil 50 – 100 Graphite (polycrystalline) 0.00001 – 0.0001
Clay 1 – 150 Pyrrhotite 0.00001 – 0.01
Graphite schist 10 – 500 Chalcopyrite 0.005 – 0.1
Gravel 100 – 600 Graphite (shales) 0.001 – 10
Weathered bedrock 100 – 1,000 Pyrite (ores) 0.001 – 100
Moraine 8 – 4,000 Galena 0.001 – 300
Sandstone 150 – 5,000 Cassiterite 0.001 – 10,000
Loose sand 50 – 5,000 Magnetite 0.01 – 1,000
Gneiss 400 – 6,000 Hematite 0.01 – 100,000
Basalt 200 – 100,000 Zincblend >10,000
Granite 500 – 250,000 Quartzite 500 – 800,000
Gabbro 100 – 500,000 Sphalerite 1,000 – 1,000,000
Limestone 200 – 10,000
Greenstone 500 – 200,000
Slates 500 – 500,000
Salt beds & domes 10,000 – 1,000,000
10
Quartz >10
12
Limestone (marble) >10

64
Some minerals, notably graphite, pyrrhotite, chalcopyrite, galena and magnetite, are relatively
good conductors (Table 4.4). A dissemination of such minerals within a rock can also make it a better
conductor but very much depends upon the nature of the dissemination. Pyrite, chalcopyrite,
pyrrhotite, etc., can occur as rolled crystals or flakes with intimate contact between crystal surfaces.
Ores containing these minerals are very often good conductors. Galena and magnetite have a
tendency to form individual, euhedral crystals, more or less isolated from each other, so that galena
and magnetite ores are as a rule poorer conductors than, for example, pyrite ores.
Practically all rocks and minerals are semiconductors, their conductivity increasing with
increasing temperature according to the equation

σ = σ0e-E/κT (4.12) ξ

where T is the absolute temperature (K), κ is Boltzmann's constant and E is the so-called energy gap
or activation energy for the substance in question. The activation energies are, however, poorly
known for most of the naturally occurring minerals. Stoichiometric hematite and zincblend are
insulators but certain impurities in hematite, for example, can reduce the resistivity considerably.
These so-called donor or acceptor impurities also play a great part in determining the magnitude of
the conductivity of naturally occurring phenomena based on quantum mechanics.ii

4.4.1.6 Archie’s law

T he resistivity of porous, water-bearing rocks (free of clay minerals) is roughly equal to the
resistivity of the pore fluids divided by the fractional porosity. A closer approximation in most
instances follows Archie’s empirical equation or law which states that resistivity is inversely
proportional to fractional porosity raised to a certain power. Considerably important in electric well
logging, Archie’s equation is given by

ρ = ρwφ-ms-n (4.13) ξ

where ρw is the resistivity of the water filling the pores, φ is the porosity (volume fraction of pores), s
is the fraction of pore space filled by the water and m and n are certain parameters.
The value of n is usually close to 2.0 if more than about 30% of pore space is water-filled but can
be much greater for lesser water contents. The value of m depends upon the degree of cementation
or upon the geologic age of the rock. It varies from about 1.3 for loose, Tertiary sediments to about
1.95 for well cemented, Paleozoic ones, but can be outside this range for individual formations.
Equation (4.13) can be rewritten ρw / ρ = φmsn where the right side represents the fractional
porosity whose exponents m and n vary according to the shape of the matrix grains. The departures
from linearity are not large for common values of porosity (Figure D.2, Appendix D).
Section D.2 in Appendix D discusses several methods of determining resistivity when the rock
resistivity is high and the samples are inhomogeneous as well as overcoming problems of contact
resistance between the current electrodes and the sample faces.

4.4.1.7 Apparent resistivity

T he result of a single measurement with any array could be interpreted as due to homogeneous
ground with a constant resistivity. The geometric factors used to calculate this apparent
resistivity, ρa, can be derived from equation (4.9). By letting G be a geometric constant representing
[(1/r) – (1/r’)], (4.9) becomes

U = Iρ / 2πG

solving for ρ,
2π U
ρ = (4.14) ξ
G I

65
where U is the electric potential at electrode distances G at the surface of a uniform half-space
(homogeneous ground) of resistivity ρ. The current I may be positive (if into the ground) or negative.
The potential at any point is equal to the sum of the contributions from the individual current
electrodes.
Let A and B be the current electrodes, positive and negative respectively, placed on the ground
surface and M and N the potential probes. If UMN is the voltage difference between M and N it follows
from (4.14) that
UMN
ρ = 2π (4.15a) ξ
IG
where
1 1 1 1
G= - - + (4.15b) ξ
AM BM AN BN

In a four electrode array over homogeneous ground

Iρ 1 1 1 1
UMN = - - + (4.16) ξ
2π AM BM AN BN

where U is the potential difference between electrodes M and N due to a current I flowing between
electrodes A and B.
Equation (4.15a) yields the resistivity of a homogeneous earth. In an actual case, ρ will vary if the
geometric arrangement of the four electrodes is altered or if they are moved on the ground without
altering their geometry. That is, UMN / I will not be directly proportional to G as on a homogeneous
earth. The value of ρ, obtained by substituting the measured UMN / I and the appropriate G in
equation (4.15a), is the apparent resistivity ρa. It can be calculated for given electrode
arrangements for assumed subsurface resistivity distributions by solving the boundary value problem
involved either analytically or numerically.

4.4.1.8 Electrode arrays

F igure 4.9 illustrates several of the most common electrode arrays and their geometric factors.
Through out the following, A and B will denote the current electrodes and M and N the potential
probes. In the case of Figure 4.9 (d), note that the B probe goes to infinity. The distance to a fixed
electrode ‘at infinity’ should be at least 10 and ideally 30 times the distance between any two mobile
electrodes.

(a) Wenner array


This array is very widely used, with a vast amount of interpretational material and computer
packages available. It is considered the ‘standard’ array against which others are assessed. In this
array (Figure 4.9 (a)), the four electrodes are collinear and the separations between adjacent
electrodes are equal, denoted by (a), with M and N in between A and B as shown.
By setting 2π / G = K in equation (4.15a), the equation becomes
UMN
ρa = K (4.17) ξ
I
where K becomes the universal geometric constant or ‘K-factor’ for electrode arrays.
In the Wenner array (Figure 4.9 (a)), AM = 1a, BM = 2a, AN = 2a, and BN = 1a. Therefore,
setting a = unity, the potential UMN = (Iρ /2π)(1-½ -½ +1)a -1 and G = 1a -1. From this it can be seen
that the ‘K factor’ for the Wenner array is 2πa and the apparent resistivity is given by
UMN
ρa = 2πa (4.18) ξ
I

66
Figure 4.9 Some common electrode arrays
I I

V V
A M N B c
A M N B

s s
a a a
L L

(a) Wenner (b) Schlumberger


K = 2πa, ρa = 2πa(UMN / I) K = πL2 / 2s, ρa = πL2 / 2s (UMN / I)

I V V I ∞

A B M N M N A

a na a a na

(c) Dipole-dipole (d) Pole-dipole


ρa = πn(n+1)(n+2)(a)(UMN / I) ρa = 2πn(n+1)(a)(UMN / I)

(b) Schlumberger array


Also using four electrodes, this array rivals the Wenner array in availability of interpretational
material, all of which relates to the ‘ideal’ array with a negligible distance between the inner
electrodes and is favored, along with Wenner, for electrical depth-sounding work.
The M and N electrodes in the Schlumberger array are also between A and B and placed
symmetrically at the center, but MN (=2s) << AB (=2L). At an arbitrary point P on the line AMNB we
get, using equation (4.15a) and measuring x from the center of the line,

A M c N P B
Iρ 1 1
x UMN = -
2π L+ x L-x
L L

from which
dU Iρ 1 1
= - ξ
2-
dx 2π (L+ x) (L – x)2

At the center (x = 0), therefore


dU Iρ
= -
dx πL2

which gives for the apparent resistivity

πL2 dV
ρα = -
I dx (4.19a) ξ

If MN is sufficiently small, we may put (-dV/dx) = UMN / 2s (Figure 4.9 (b)) where UMN is the
measured voltage difference between M and N (potential of M minus potential of N). Then

67
πL2 UMN
ρα = (4.19b) ξ
2s I

where πL2 / 2s is the array constant or ‘K factor’.

(c) Axial dipole-dipole array


The arrangement shown is Figure 4.9 (c) is usually referred to as dipole-dipole. However, a
dipole should consist of two electrodes separated by a distance that is negligible compared with the
distance to any other electrode. Applying the term to dipole-dipole and pole-dipole arrays, where the
distance to the next electrode is usually from one to six times the ‘dipole’ spacing, is thus misleading.
The more accurate terminology would preferably be bipole-bipole.
In this array, (M and N are outside A and B and on the same line, each pair having a constant
mutual separation a. If na is the distance between the two innermost electrodes (B and M) then
equation (4.15a) yields for the apparent resistivity using the dipole-dipole array

UMN
ρα = πn(n+1)(n+2)a (4.20) ξ
I

(d) Pole-dipole (three electrode) and pole-pole (two electrode) arrays


When one of the current electrodes, say B, is very far removed from the measurement area, say
30 times distant, the electrode A is referred to as a current pole. The MN array may be a dipole
(length MN << distance AM) or bipole. We then have a pole-dipole or pole-bipole array.
Conversely, when one of M or N is far removed, the remaining electrode is referred to as a potential
pole. If the current dipole AB is then short compared with is distance from the potential pole we have
a dipole-pole array. A pole-pole arrangement will be obtained when one of A or B and one of M or N
are removed to ‘infinity’. Figures 4.7 (a) and 4.8 really imply and ideal pole-pole arrangement.
The pole-dipole or dipole-pole arrangements are not uniquely defined by these names alone
since the orientation of the dipole (or bipole, as the case may be) with respect to the dipole-pole line
remains to be specified. In their particular forms in which the electrodes are collinear these
arrangements are often referred to as three-electrode arrays while the pole-pole is often referred to
as a two-electrode array. A particular case of pole-dipole is the half-Schlumberger array in which A
(or B) in Figure 4.9 (b) is effectively at infinity.
The pole-pole (or two-electrode) array is very popular in archaeological work and as the normal
array, is one of the standards in electrical well logging. The pole-dipole (or three electrode) array
produces asymmetrical anomalies which are more difficult to interpret than those produced by
symmetrical arrays. Peaks are displaced from the centers of conductive or chargeable bodies and
electrode positions have to be recorded with particular care. Values are usually plotted at the point
mid-way between the moving voltage electrodes, but this is not a universally agreed standard.
Results can be displayed as pseudo-sections, with depth penetration varied by varying n.
The K factors for pole-dipole and pole-pole arrays are 2πn(n+1)a and 2πa respectively. Therefore
the apparent resistivities are
UMN
ρa = 2πn(n+1)a (4.21) ξ
I
for pole-dipole arrays, and
UMN
ρa = 2πa (4.22) ξ
I

for pole-pole or two-electrode arrays.

Other arrays include the gradient array which is widely used for reconnaissance work. Large

68
number of readings can be taken on parallel traverses without moving the current electrodes if
powerful generators are available.
The square array utilizes four electrodes that are positioned at the corners of a square and are
variously combined into voltage and current pairs. Depth soundings are made by expanding the
square; when traversing, the entire array is moved laterally. It is somewhat inconvenient, but can
provide the interpreter with vital information about ground anisotropy and inhomogeneity. Few
published case histories or type curves are available.
The multi-electrode array resembles the Wenner array, but has an additional central electrode.
The voltage differences from the center to the two ‘normal’ voltage electrodes give a measure of
ground inhomogeneity. The two values can be summed for application of the Wenner formula. Other
multi-electrode arrays are designed to focus the current into the ground to give deep penetration
without large expansion.

Wenner offset electric sounding

1000
Place: . South River, Copan

Coordinates: X 268,725 Y 1,690,777


200
Datum: . V84 .
100
Date: . July 7, 2000 .

Operator: Rk .
20
Aparatus: . RMCA4 .
10
Sounding No. 7 .

Profile No. D .

Time: AM .
1
0.1 0.2 1 2 10 20 100

2πa UMN
K= ρa (Ωm) = K
1/AM – 1/AN – 1/BM + 1/BN I

RW =
Sounding Separation (a) m (RD1) (RD2) ρa (Ωm) = 2πaRW
(RD1+RD2)/2
1 0.25 20.6 19.45 20.03 31.46
2 0.5 7.7 9.9 8.8 27.65
3 1 2.7 3.4 3.05 19.2
4 2 1.35 1.65 1.5 18.85
5 4 1.00 1.10 1.05 26.39
6 8 0.80 0.75 0.78 38.96
7 16 0.65 0.5 0.58 57.8
8 32 0.67 0.55 0.61 122.6

Figure 4.10 Operators log for a Wenner electric sounding in Copan, Honduras

69
4.4.1.9 Electrical sounding and mapping

F igure 4.10 shows the results of an electric sounding in Copan, Honduras in July 2000. The
method used was the Wenner offset array stepping out the electrode separation (a) by a factor of
2 from 0.25 m to 32 m. The object of the survey was to investigate the near-surface for the presence
of an ancient riverbed with the ultimate goal of locating Mayan residences that tended to be placed on
the river’s edge. The graph in the upper right-hand corner of the operator’s log indicates first a
decrease in resistivity then increases as the survey extends from clayish topsoil to the riverbed sand.
The above is an example of two procedures used to explore the properties of the subsurface:
electric sounding (or vertical electric sounding, VES) and electric mapping. The object of sounding is
to determine the variation of electrical conductivity with depth by measuring the apparent resistivities
of horizontally bedded strata or deposited material and to determine the thickness and/or depth of
those boundaries where resistivities change.

A B M1 N1 M2 N2

(a)

2
1

A2 B2 A1 B1 M N

(b)

Figure 4.11 Two VES methods: (a) fixed electrodes A and B while
moving electrodes M and N (b) fixed M and N while moving A and B

The essential idea behind VES, assuming conductivity variation with depth only, is the fact that,
as the distance between the current and potential electrodes is increased, the current filament
passing across the potential electrodes carries a current fraction that has returned to the surface after
reaching increasingly deeper levels. This will be appreciated from Figure 4.11 (a). Here, the current
filament [2] flowing across the electrode pair M2, N2 carries current that has returned to the surface
from a deeper level than the filament [1] flowing across the pair M1, N1 nearer to the current source of
A, B. Thus the potential difference UMN(2) and hence the apparent resistivity ρa2 in the position M2, N2
are relatively more influenced by the conductivity at deeper levels than are the quantities UMN(1) and
ρa1.

70
An analogous argument applies when M and N are kept at the same place but A and B are
moved instead (Figure 4.11 (b)). Thus a sounding can be performed by moving either the pair A,B or
the pair M,N or in fact, both. Note that as long as the distance AB is not altered, the current
distribution as such in the earth is not altered, that is, the fraction of the total injected current
penetrating below any given level before returning to the surface is exactly the same irrespective of
the relative positions of the A, B and M, N pairs. This fraction will, however, alter if the distance AB is
changed.
It must be remembered that here, as in all earth resistivity measurements, UMN represents the
measured voltage between M and N minus any self-potential voltage between M and N observed
before the current is passed. Most commercially available modern resistivity instruments
automatically compensate for the SP by first measuring the MN voltage with the current in the positive
direction, that is with A in Figure 4.9 as the positive electrode, in which case the measured voltage
will be U(+) = SP + UMN, and next with an equal current in the opposite direction, which then gives the
reading U(-) = SP – UMN, since only UMN changes sign with the current. Then UMN = (U(+) – U(-)) / 2
while SP = (U(+) + U(-)) / 2.ii

4.4.1.10 VES field procedure

(a) Wenner sounding: In this case the distance a (Figure 4.9 (a)) is increased by steps, keeping the
midpoint of the configuration fixed. The apparent resistivity is obtained from equation (4.17). An
example of this procedure is shown in Figure 4.10. Two readings, RD1 and RD2, are taken from each
direction and then their absolute value, which is read directly as UMN / I, is averaged (RW) then
multiplied by K (=2πa) to give ρa.

(b) Schlumberger sounding: The electrodes M and N (Figure 4.9 (b)) are kept fixed and A and B are
moved outwards symmetrically in steps. At some stage the MN voltage will, in general, fall below the
reading accuracy of the voltmeter in which case the distance MN is increased (e.g. five or ten-fold),
but the condition MN << AB maintained. It is advisable then to have an overlap of two or three
readings with the same AB and the new as well as the old MN distance. The ρa values with the two
MN distances but the same AB may sometimes differ significantly from each other and if, in this case,
the results are plotted as ρa against AB (or, as is the usual practice, against AB/2 as in Figure 4.12),
each set of ρa values obtained in the overlapping region with one and the same MN will be found to lie
on separate curve segments, displaced from each other (Figure 4.13).

ρa
A M
V N B
0

AB/2

Figure 4.12 Illustration of Schlumberger sounding with resistivity graph

The different segments must be suitably merged to obtain a single smooth sounding curve.
Often, it is sufficient to shift a segment obtained with a larger MN towards the adjoining previous one
obtained with the smaller MN, as shown in Figure 4.13, the reasoning being that UMN / MN measured
for a smaller MN value is nearer to the true gradient in equation (4.19a) than the value with the larger
MN. The predominant contribution to the shifts comes, however, in practice from the differing effect
of local inhomogeneities at the old and new positions of M and N electrodes. The effect exists in
Wenner soundings as well but it is not readily identifiable since the steps between successive
positions of M and N are large.

71
ρa

AB/ 2

(MN)2
(MN)3
(MN)1

Figure 4.13 Displacement of MN segments


in Schlumberger sounding

(c) Dipole-dipole sounding: The basic procedure in this case is illustrated in Figure 4.11, which
shows an axial ‘dipole-dipole’ system. The apparent resistivity is given by equation (4.20).
Soundings can be made, in principle, with any of the dipole configurations in Figure 4.14, but in most
other cases the field routine will be inconvenient.
Similarly, the pole-pole, pole-dipole and dipole-pole arrangements could also be used, but the
effect of the far electrode (B or N) is hardly ever negligible in practice and the measurement
becomes, to all intents and purposes, a four electrode one, so that is then just as well to use one of
the three previously discussed principal sounding procedures.

M
N N

r r r
A B M N Azimuthal
Radial
Axial
θ
A B A B

M N
M
M N
N

r r r
Perpendicular Equatorial
Parallel
θ θ
A B A B A B

Figure 4.14 Various dipole-dipole configurations In the axial configuration A, B, M and N are
collinear. In the radial one the MN dipole is orientated along the line joining the dipoles. The axial
array is a special case of the radial. In the azimuthal configuration the MN dipole is perpendicular to
the line joining the dipoles. In the parallel configuration the dipoles are parallel but not collinear. A
special case of it (θ = 90º) is the equatorial configuration in which the MN dipole center is always on
the line joining the dipoles. When the dipoles are perpendicular to each other we have the
perpendicular array. The advantage of dipole-dipole arrays is that the distance between the current
source and the potential dipole can be increased almost indefinitely, being subject only to instrument
sensitivity and noise, whereas the increase of electrode separation in the Wenner and Schlumberger
arrays is limited by cable lengths. The derivation of the formulae for apparent resistivities of the
various dipole-dipole arrays is given in section D.3 in Appendix D.

72
T he object of electrical mapping is to determine the lateral variations in the conductivity of the
ground. Lateral differences in apparent resistivity are mapped for a distinct depth level. Mapping
is primarily useful for detecting local, relatively shallow inhomogeneities and is employed typically in
ore prospecting and in delineating geologic boundaries, fractures, cavities, palaeochannels, rims of
disposal sites and deposits of contaminated materials etc. It has been used in archaeology for finding
ancient buried structures. Very often the problem is to find these various features below a relatively
uniform overburden like soil, deeply weathered rock, glacial drift, etc., in which case it is necessary to
know and control in field work the fraction of injected electric current penetrating below the
overburden.
The data are obtained from a fixed array, which records the potential difference between the
potential electrodes and is moved step-by-step along survey lines until the whole survey area is
covered. The result is presented as a contour map or resistivity section.
The most common arrays utilized in mapping are the Wenner, Schlumberger, gradient, and
dipole-dipole arrays. Arrays are usually chosen at least partly for their depth of penetration, which is
almost impossible to define because the depth to which a given fraction of current penetrates
depends on the layering as well as on the separation between the current electrodes. Voltage
electrode positions determine which part of the current field is sampled so the penetrations of the
Wenner and Schlumberger arrays are thus likely to be very similar for similar total array lengths. For
either array, the expansion at which the effect of a deep interface becomes evident depends on the
resistivity contrast, but is on the order of one-quarter to one-half of the separation between the current
electrodes if the measurements of potential are made between them, or similar fractions of the
distance between the current and potential dipoles if measurements are made by a dipole-dipole
system. (See Appendix E for a discussion on dipole-dipole mapping)
For any array, there is a depth at which the effect of a thin layer of different resistivity in otherwise
homogeneous ground is a maximum. This would be expected to be less than the depth at which the
interface in a two-layer earth first becomes apparent. By this criterion, the Wenner array is the least,
and the dipole-dipole the most, penetrative array. However, the Wenner curve is the most sharply
peaked, suggesting superior resolving power. This is confirmed by the signal contribution contours,
which are slightly flatter at depth for the Wenner than for the Schlumberger array, indicating that the
Wenner array locates flat-lying interfaces more accurately. The contours for the dipole-dipole array
are near vertical in some places at considerable depths, showing that the array is best suited to
mapping lateral changes.
(a) (b)

ρa low

ρa high

(c)
Figure 4.15 Current flow patterns
for (a) uniform half-space, (b) two- ρa high
layer ground with lower resistivity in
the upper layer and (c) two-layer
ground with higher resistivity in the ρa low
upper layer.

Current flow patterns for one and two layers are shown in Figure 4.15. Near-surface
inhomogeneities strongly influence the choice of array. A conductive near-surface layer will in some
places increase and in other places decrease the apparent resistivity. In homogeneous ground these
effects can cancel fairly precisely.
When a Wenner or dipole-dipole array is expanded, all the electrodes are moved and the
contributions from near-surface bodies vary from reading to reading. With Schlumberger, near-

73
surface effects vary much less, provided that only the outer electrodes are moved, and for this reason
the array is often preferred for depth sounding. Near-surface effects may be large when a gradient or
two-electrode array is used for profiling, but are also very local; a smoothing filter can be used.

4.4.2 Self-potential

N atural currents are always present in the earth and are associated with measurable spontaneous
or self-potentials (SPs). The measurement of this naturally occurring self-potential by direct
current methods leads to a better understanding of the electrical properties of the subsurface which in
turn helps in the identification of geological and environmental anomalies. The SP method was used
as early as 1830 for studying sulphide veins in Great Britain but the systematic use of SP and
resistivity methods dates from about 1920.

4.4.2.1 Occurrence and origins of self-potentials

F or the implementation of the resistivity method, discussed in the previous section, an artificially
induced electrical current was required. The SP method, however, as its name implies, is based
on measuring the natural potential differences which generally exist between any two points on the
ground. The potential differences, partly constant and partly fluctuating, are associated with electric
currents in the ground. The constant and unidirectional potential differences between any two points
can arise due to various electrochemical mechanisms (section 4.4.2.3).

Groundwater
flow

Water
table
Sulphide
mass
SP (mV)

-100
-200
-300

Figure 4.16 SP profile across a sulphide body with hydrological influence of ground-
water flow

74
The range of natural potentials is generally from about 0.5 millivolts (mV) to about 50 mV,
although even higher values can be attained. Negative anomalies of a few hundred millivolts are
sometimes produced by sulphide and graphite ore bodies. Large SP values are also found over coal
and manganese deposits, magnetite and other electrically conductive materials. Potentials of as
much as 1.8 V have been measured where alunite weathers to sulphuric acid. Unless superimposed
by other self-potentials these ‘mineral potentials’ are invariably negative with respect to a point far
away from the mineral occurrence when the dips are steep but they may be flanked by positive
maxima. Positive SP maxima on the ground surface over the above-mentioned minerals are found in
cases of flat dip.
Distinct, although considerably smaller in magnitude, are the anomalies often observed on other
features like quartz veins, pegmatite, etc. Self-potentials are also associated with flow of water in the
ground and are therefore of interest in geohydrology.
For an ore body such as sulphide to be a good conductor, it should extend from the zone of
oxidation near the surface to the reducing environment below the water table, thus providing a low-
resistance path for oxidation-reduction currents. Figure 4.16 shows a sulphide mass straddling the
water table and concentrating the flow of oxidation-reduction currents producing a negative anomaly
at the surface. The down-slope flow of groundwater after rain produces a temporary SP, in this case
inversely correlated with topography.
Small potentials, seldom exceeding 100 mV and usually much less, may accompany groundwater
flow. The polarity depends on the rock composition and on the mobilities and chemical properties of
the ions in the pore waters, but most commonly the region towards which groundwater is flowing
becomes more electropositive than the source area. These streaming potentials are sometimes
useful in hydrogeology, but may render mineral exploration inadvisable for up to a week after heavy
rain.
Movements of steam or hot water can explain most of the SPs associated with geothermal
systems, but small (<10mV) voltages, which may be positive or negative, are produced directly by
temperature differences. Geothermal SP anomalies tend to be broad (perhaps several kilometers
across) and have amplitudes of less than 100 mV, so very high accuracies are needed.
As for the fluctuating potential differences, small alternating currents are induced by variations in
the earth’s magnetic field and by thunderstorms. Only the long period components of the associated
voltages, seldom more that 5 mV, will be detected by the direct current (d.c.) voltmeters used in SP
surveys. If, as is very occasionally the case, such voltages are significant, the survey should be
repeated at different times of day so that the results can be averaged.
Self-potentials are surprisingly stable in time as shown in the example of measurements along a
line in northern Sweden (Figure 4.17). Here, several seasonal cycles of snow, rain and drainage
have not affected the pattern of self-potentials to any significant amount.

1960
-mV

200 m
-50 1962
+50
+mV

1967

Figure 4.17 Long-term SP stability over the same 1.5 km track

4.4.2.2 Field procedure

T he measurement of self-potentials is fairly straight-forward. Readings may be taken directly from


any millivoltmeter with a sufficiently high impedance connected to electrodes driven some 10-15
centimeters into the ground. Alternatively, a usual potentiometer circuit of the null type can also be
employed with advantage. The electrodes must be non-polarizable (e.g. copper (Cu) in copper

75
sulphate (CuSO4) solution or platinum (Pt)–calomel in potassium chloride (KCl) solution). Ordinary
metal stakes will generally not work since electrochemical action at their contact with the ground
tends to obscure the natural potentials.
Two alternative procedures are in use for SP surveys. In one, two electrodes separated by a
small constant distance, say 5, 10 or 20 meters, to measure average field gradients. This procedure
is useful if the amount of cable is limited, but errors tend to accumulate and coverage is slow because
the voltmeter and both electrodes must be moved for each reading. More commonly, voltages are
measured in relation to a fixed base. One electrode and the meter remain at this point and only the
second electrode is moved. Sub-bases must be established if the cable is about to run out or if
distances become to great for easy communication. Voltages measured from a base and a sub-base
can be related provided that the potential difference between the two bases is accurately known.

V
V

M M’ N

Figure 4.18 Using sub-bases in an SP survey The value at the new base M’ relative to
M is measured directly and also indirectly by measurements of the voltage at the field point
N relative to both bases.

Figure 4.18 shows how a secondary base can be established. Although the end of the cable has
almost been reached at field point M’, it is still possible to obtain a reading at the next point N, using
the original base at M. After potential differences have been measured between M and both M’ and
N, the field electrode is left at N and the base electrode is moved to M’. The potential difference
between M and M’ is thus estimated both by direct measurement and by subtracting the M’ to N
voltage from the directly-measured M to N voltage. The average difference can be added to values
obtained with the base at M’ to obtain values relative to M.

4.4.2.3 Electrochemical mechanisms

There has been considerable speculation concerning the electrochemical mechanisms producing
self-potentials and even now it cannot be said the question has been resolved. However, the main
mechanisms appear to be as follows.

(a) Electrofiltration: Electrofiltration or streaming potentials (also called zeta potentials) are a
manifestation of a phenomenon, first studied by Helmholtz in the 19th century, that an electric
potential difference is developed between the ends of a capillary tube through which an electrolyte is
flowing. The effect is believed to be due to the attachment of ions to the walls of the capillary,
combined with the unequal flow velocities of positive and negative ions. It is found that the electric
field (V m-1) along the capillary can be expressed as

E = (ερζ / η) p (4.23) ξ

where ε is the dielectric constant of the electrolyte (F m-1), ρ is the resistivity of the electrolyte (Ωm), ζ
is a parameter (streaming potential, V) that depends on the material of the capillary wall and the
electrolyte, p is the pressure gradient (Pa m-1) and η is the dynamic viscosity of the electrolyte (Pa s).
Thus E is in the same direction as the pressure gradient, that is, opposite to the direction of
electrolyte flow. The parameter C = (ερζ / η) is called the streaming potential coefficient and for earth
materials it ranges from a fraction of a mV per kilopascal (0.1 m water head) to about 1.3 mV kPa-1.
In geophysical work, electrofiltration potentials can be found associated with the flow of water
through sand, porous rocks, moraine, etc. All but the potentials associated with the minimum at the
extreme left in Figure 4.17 are streaming potentials. (Note that the y-axis in Figure 4.17 is negative
upwards.) High ground is generally found to be more negative than low ground, indicating that the

76
electric current due to these potentials tends to flow uphill, which is in accordance with equation
(4.23). Thus, a potential of –1.842 V associated with alunite (a basic sulphate of aluminum (Al) and
potassium (K)) has been reported on a mountain top near Hualgayoc, Peru (Gay, 1967b) and –1.940
V over unmineralized quartzite on a hill in Shillong, India (Nayak, 1981).
The flow of water due to a pressure gradient is not a sufficient condition for electrofiltration
anomalies to arise. It is necessary that the flow takes place across a boundary between two rock
media for which C is different. Some flow models of hydrogeological interest and their expected
electrofiltration anomalies are shown schematically in Figure 4.19. Electrofiltration potentials have
been used to detect leakage spots on submerged slopes of earth-dam water reservoirs (Ogilvy et al.,
1969), to detect concealed karstic cavities, springs, etc., and to study the effect of pumping on the
groundwater table.ii

mV

- +
(a) 0

C1 C2

mV

(b) 0

mV

+ + +
(c)

0
C1
C2

Figure 4.19 Schematic of electrofiltration SP anomalies


(profiles and surface maps) on some flow models: (a)
horizontal flow across a vertical boundary; (b) pumping from
a well; (c) vertical flow across a horizontal boundary.

(b) Concentration differences (diffusion): Differences in the concentration of electrolytes in the


ground from place to place may be expected to produce electric potential differences as can be seen
from the following. Imagine an electrolytic cell such as the one in Figure 4.20 with two metal
electrodes in two electrolytic solutions of concentrations C1 and C2 in contact with each other.
Assume the system to be such that when the cell is in action the net transfer of charge from one
electrode to the other within the cell takes place through positive ions. For instance, Cu electrodes in
CuSO4 solutions form exactly this type of cell. It can be shown by thermodynamic reasoning that the
overall electromotive force (the difference (potential of electrode 2) minus (potential of electrode 1) of
such a cell is

U = (2ν / (u +v))(RT / nF)ln(C1 / C2) (4.24) ξ

77
where u and v are respectively are the velocities ( m s-1) of the positive and negative ions in the
electrolyte when the cell is in action, R is the universal gas constant (8.314 J K-1mol-1), T is the
absolute temperature (K), F is Faraday’s constant (96.487 C mol-1) and n is the valence of the ions. If
the charge transfer takes place through negative ions, u and v in (4.24) will be unchanged.

Cu electrode Cu electrode
1 2

Concentration Concentration
C1 C2

Figure 4.20 Electrolytic concentration cell

Now, suppose an electrolyte is so chosen that u = v. The e.m.f. of this cell will be

U0 = (RT / nF)ln(C1 / C2) (4.25) ξ

However, U0 is simply UCu2(C2) – UCu1(C1) if UCu is the potential of the copper electrode involved in
contact with an electrolyte of concentration C. The e.m.f U0 also exists in the cell of equation (4.24)
as well, so that the difference (U - U0) = U1 must be attributed to the fact that u ≠ v so that

U1 = (u – vRT) / (u + vnF)ln(C1 / C2) (4.26) ξ

U1 is called the liquid-junction potential although its cause is not the junction as such but rather the
unequal velocities of the positive and negative ions when the cell is in action. Equation (4.26) is
written with the convention that U1 represents the difference (potential of solution on the right) minus
(potential of solution on the left).
An extensive study of diffusion potentials on peat and bogs was made by Schuch (1963), who
found SP values down to about –120 mV caused by concentration differences.
If the concentration of electrolytes in the ground varies locally, potential differences will be set up
in accordance with equation (4.24) and this may be the cause of the ubiquitous ‘background’ SP
anomalies of the order of fractions of a millivolt to some tens of millivolts. This is plausible but the
difficulty with this theory is that over the passage of time any concentration differences in the ground
will disappear owing to the diffusion of the more concentrated solute to the less concentrated one.
Hence it becomes necessary to postulate a source that will continuously regenerate concentration
differences. No such source has been positively identified in actual natural systems but exchange
with atmospheric oxygen is a strong candidate.ii
One of the most common solutes in natural electrolytes is NaCl. Differences in NaCl
concentrations produce SP anomalies that are of great significance in well logging. This topic is
considered in more detail in section 5.6.1.

(c) Adsorption: SP anomalies of the order of +20 to +40 mV are characteristically obtained over
pegmatite and quartz even when the rocks are completely concealed under an overburden. In the
example of Figure 4.21 the boundaries of gneiss intruded by pegmatite veins are clearly shown by the
sharp gradients. Such self-potentials have been attributed to the adsorption of positive and negative
ions (double layers) on the surface of the veins (although what is actually measured is, the ordinary

78
ohmic drop due to a flow of current) but the details of the electrochemical mechanism are by no
means clear. Self-potentials observed over clays probably also belong to this category.

50mV

200 m
Tertiary Probable boundaries
Pegmatite Gneiss
cover of pegmatitic gneiss
Semenov (1980)
Figure 4.21 SP due to ion adsorption on pegmatite dikes

(d) Mineral potentials: These are often called by the less general name of sulphide potentials since
they are as a rule strongest on ores like pyrite and chalcopyrite but their common feature is really that
they are observed on electronically conducting minerals. Early theories attributed mineral self-
potentials to the oxidation of parts of a mineral deposit above the water table, but such an explanation
cannot fit graphite, which shows high SP anomalies but does not normally undergo significant
oxidation. Moreover, oxidation implies a loss of electrons so that the upper end of a mineral deposit
should become positively charged whereas almost invariably a negative SP anomaly is observed on
the ground surface above ore bodies (Figure 4.16). However, it was pointed out by Sato and Mooney
(1960) that the ore body need not directly take part in the electrochemical reactions. Their theory
amounts briefly to the following.
Since we are concerned with an electronic conductor (the ore) in contact with an ionic conductor
(electrolytes in the bedrock), there must be an electrochemical action at their boundary. It is plausible
to assume that, just as a metal electrode (Me) dipped in an electrolyte liberates ions in the electrolyte
according to the reaction Me → Me+ + e-, an electronically conducting ore sulphide MeS will liberate
ions according to the reaction MeS → Me2+ + S + 2e- where e- is the electronic charge. There may
also be more complex reactions in some cases liberating negative ions such as OH- or SO42-. If the
concentration of ions in the electrolyte is different near the top and bottom parts of the ore body, as
may well happen if the electrolytic solution pressures are different, an ionic current will flow in the
surrounding electrolyte.ii The mechanism may be compared to the concentration referred to in
section 4.4.2.3 (b) when it is in action. The directions of current and ion flow implied by a negative SP
center over an ore body are shown in Figure 4.22.

Negative Evidently, the ore body, being a good


Ground surface center
electronic conductor, serves to transport the
electrons from the lower part to the upper part.
Electric
The conventional electric current within the ore
current is thus from the top to the bottom. In the
surrounding medium, the currents go from the
Positive lower end of the ore to the upper end. The
ions Negative density of this latter current, which may be
ions called ionic current density, can be
Electron
flow estimated from potential measurements in drill
holes in the surrounding rock and a knowledge
of its electrical resistivity, from the simple
bedrock relationship potential gradient = current
density x resistivity.

Figure 4.22 Electron and ion flow in and around an ore body

79
Electrode Electrode

Drop IR2
U3 Drop IR3
Current U4
filament

Drop IR1

Ore body

U1
U2

Figure 4.23 Potential drops in SP observations

4.4.2.4 Static or equilibrium potential

T he theory of Sato and Mooney (1960) explains the prevalence of negative centers over pyrite,
graphite, etc., as well as some converse phenomena, e.g. the absence of SP anomalies on
sphalerite (ZnS) or hematite (Fe2O3), which are very poor conductors. Sato and Mooney also
calculate the maximum potential differences that could be observed on various types of electronically
conducting ores. However, these calculations postulate equilibrium electrochemical conditions in that
the effect of the ohmic potential drop in the electrolyte due to current flow is disregarded. We shall
therefore look somewhat more closely at the phenomenon following Kilty (1984).
The field measurement of SP is essentially a measurement of the ohmic drop IR between two
potential electrodes where I is the current and R the electrical resistance along the current filament
between the electrodes. Figure 4.23 shows such a current filament with a drop IR2. We shall
assume that the electrolytes near the top and bottom of the ore are chemically identical except for the
ionic concentrations. The interface reactions that liberate ions cause a potential difference between
the ore and the solution in contact with it. Thus, the potentials U1 and U4 at the two ends are different
from U2 and U3 just outside the ore. It can be seen from the illustration that V1 and V4 are not directly
accessible to observation. The interface potential drops at the bottom and top are therefore

∆φb = U1 – U2 ∆φt = U4 – U3 (4.27) ξ

In addition, along the current filament there are the ohmic potential drops IR1, IR2 and IR3 where R1,
R2 and R3 denote the electrical resistances along the different sections of the current path. What is
measured as the SP anomaly at a point is the drop IR2.
According to Kirchoff’s law in electric theory the sum of all potential drops along a closed current
path is zero. Hence

(U1 – U2) + IR1 + IR2 + IR3 + (U3 – U4) + (U4 – U1) = 0

that is, the observed SP is

IR2 = -Uore + ∆φt – ∆φb – I(R1 + R3) (4.28) ξ

where Uore is the potential drop (U4 – U1) within the ore. (When referencing Kilty, note that for ∆φt and
∆φb he used ∆φc and ∆φd respectively.) For Uore, equation (4.28) gives

Uore = ∆φt – ∆φb – IR (4.29) ξ

80
where IR is the total ohmic drop in the electrolyte along the filament. The theory of Sato and Mooney
assumes I = 0 (which is not the case in Nature). If I = 0, ∆φt and ∆φb can be found from appropriate
electrochemical tables. However, since I ≠ 0, equation (4.29) shows that Uore is less than the static
or equilibrium value s(0) = ∆φt – ∆φb and that, as the current in the circuit increases, the potential drop
across the ore decreases. However, ∆φt – ∆φb is itself not constant but is a function s(I) of the current
in the system, and since Uore = IRore, equation (4.29) gives

IR + IRore = s(I)

and IR + IRore = s(0) + (ds / dI) I (4.30) ξ

by Taylor’s theorem in differential calculus, provided the current I is very small. This is the equation
that then determines the current involved in an SP anomaly. If I were to become zero, which is
notionally but not actually possible, we would have either R or Rore infinite.ii

4.5 Induced polarization

4.5.1 Introduction

W hen an electric current flows in the ground, certain rock masses become electrically polarized.
The process is similar to charging a capacitor or a battery. If this current is interrupted, the
voltage across the potential electrodes does not drop to zero instantaneously (Figure 4.25 (b)) but
instead discharges over a period of seconds (or minutes) producing secondary currents and
electromagnetic fields. This phenomenon is called induced polarization and easily observed when
electrically conducting minerals or clays are present in the ground.
Disseminated sulphide minerals produce large polarization effects and induced polarization (IP)
techniques are widely used in exploring for base metals. The arrays are similar to those used in
conventional resistivity work, the most common being the gradient and dipole-dipole. The advantage
of these arrays for reconnaissance and detailed surveys is that the current and voltage cables can be
widely separated to minimize induction effects. Current electrodes can be plain metal stakes, but
non-polarizing voltage electrodes must be used to detect the few millivolts of transient signal.
The phenomenon has been known to electrochemists studying the passage of electric currents in
electrode-electrolyte closed systems and has in this connection been called overvoltage. That a
very similar effect exists in pure dielectrics has also been known for a considerable time (Hartshorn,
1925).

4.5.2 Origin of IP

I t takes a finite, although short, time before the voltage U0 is reached when a current is switched on
as shown in Figure 4.25 (a). This implies that for an uninterrupted current flow the induced
polarization should manifest itself as a dependence of the ground impedance on the frequency of the
current. Thus, the IP phenomenon can be observed in the time as well as the frequency domain. It is
an exceedingly complex phenomenon although it superficially resembles the discharge of a capacitor
(time domain) or the variation of the impedance of a circuit consisting of resistance R and a
capacitance C in parallel (frequency domain).
In the frequency domain the apparent resistivity of the ground measured by any of the electrode
configurations in section 4.4.1 decreases as the frequency of the current increases, just as the
impedance of the parallel RC circuit does. This effect is called resistivity dispersion.
An adequate explanation of all aspects of the IP phenomenon has yet to be given but its
mechanisms can be seen generally in two ways: membrane polarization and electrode polarization.

(a) Membrane polarization Induced polarization may be observed in the absence of electrically
conducting minerals. The presence of clay particles appears to be a necessary condition for this
effect for it is not observed on clean quartz sand or similar media devoid of clay minerals. IP probably
takes place due to ionic exchanges and the setting-up of diffusion potentials (section 4.4.2.3 (b)).

81
The surfaces of clay particles, the edges of layered and fibrous materials or the cleavage faces of
crystals normally have unbalanced negative charges that attract a cloud of positive ions from a
surrounding electrolyte causing membrane polarization especially in rocks with small pore spaces.
Positive ions in the formation waters in such rocks congregate near the pore walls. If an electric field
is applied, the positive ion clouds are distorted and negative ions move into them and are trapped,
producing concentration gradients which impede current flow. When an electric current is forced
through a clay-electrolyte system, for example, positive ions can readily pass through this cloud but
negative ions are blocked, forming zones of ion concentration.
The return of the ions to the former equilibrium distribution after the current field is removed
constitutes a residual current and appears as the IP effect.

A B

Diffusion
current
Primary
current

+ -
+ -
+ -
+ -
+ -
+ -
+ -
+ -

Figure 4.24 Diffusion of ions producing IP voltage

(b) Electrode polarization In addition to the static contact potentials between metallic conductors and
electrolytes brought on by ionic conduction paths, additional overvoltages are produced by mineral
particles (e.g. pyrite grains) in which the carriers of current are electrons. An example of electrode
polarization is where grains of electrically conducting minerals are in contact with ground water.
The degree of polarization is determined by the surface area and not by the mass of mineral
involved; thus, polarization methods are well suited to exploration for disseminated porphry
mineralization. Strong anomalies are also usually produced by massive sulphide ores because of the
surrounding disseminated halos.
When a current passes across a metal electrode (electrical conductor) dipped in an electrolyte,
charge can pile up continuously at the interface when all the processes in the electrolytic reaction are
not equally rapid. This produces back - e.m.f. or electrode polarization. The extra piled-up charge
diffuses back into the electrolyte when the current is stopped, re-establishing the original equilibrium
in which a thin layer of negatrive ions is fixed to the metal electrode. The IP observed over sulphide
ores or other electrically conducting minerals like graphite or magnetite is basically a manifestation of
such diffusion process (Figure 4.24). The effect will be enhanced if the mineral grains are dispersed
rather than in a compact mass since it is essentialy a surface phenomenon and the polarization
charge will be large owing to the large total surface of the particles.
Although for equivalent areas of active surface electrode polarization is the stronger mechanism,
clays are much more abundant than sulphides and most IP effects are due to membrane polarization.

4.5.3 Measurement of IP in time domain

4.5.3.1 Apparent polarizability

W hen current is applied to the ground, the measured voltage first rises rapidly then approaches U0
asymptotically (Figure 4.25 (a)). Although in theory U0 is never reached, in practice the
difference is not detectable after a few seconds. Similarly, when a steady current flowing in the
ground is suddenly terminated (Figure 4.25 (b)), the voltage U0 between two grounded electrodes

82
abruptly to a small polarization voltage, Up and then declines to zero.

Up

U0
Potential observed at voltage U0 – Up
electrode initially (U0 – Up) then
U0 due to overvoltage Up Up
(polarization voltage)

Current applied at Current removed


metal electrodes at time t0

(a) (b)

Figure 4.25 Ground response to a square wave signal and to a spike impulse

When measurements are made by passing d.c. pulses of duration T (s) in the ground and ∆U is
the voltage remaining at a definite time t after current cut-off (Figure 4.26), the magnitude of the
observed IP is often expressed ad ∆U / U (millivolt per volt, mV V-1) if ∆U is measured in millivolts and
U, the voltage when the current was on, is measured in volts. Alternatively, the effect is often
expressed as a percentage as 100(∆U / U) if both voltages have been measured in volts or millivolts.
Commonly used values of T are in the range of 1 to 20 seconds while t is a fraction of T. This
measurement is called the apparent polarizability since in practice, the effect comes from the
polarizability of all the regions of an inhomogeneous ground. Physically, it can be written

(PtT)a = ∆Ut / U0T (4.31) ξ

where U0T is the voltage just before the current supplied to the ground for a duration T is cut off, that
is, at the beginning of the IP voltage decay (t=0), and ∆Ut is the voltage measured at time t after
current cut-off.

U0 U0

(a) (b)
∆U M

t =0 t Current on t1 t2
T Current off

Figure 4.26 IP measurement in the time domain

It is usual to send the d.c. pulse first in one direction and then in the reverse direction, after a
current-off time following the measurement of PtT. The cut-off time is generally of the same order of
magnitude as T.

83
4.5.3.2 Apparent chargeability

S imilar to polarizability but distinct in its measurement is chargeability which is formally defined as
the polarization voltage developed across a unit cube energized by a unit current and is thus in
some ways analogous to magnetic susceptibility.
Sometimes a normalized time integral representing the area under the decay curve between two
times t1 and t2 after current cut-off is used to express IP. This measure is known as apparent
chargeability (denoted (MTt1 t2 )a). I Thus,
t2
(MTt1, t2)a = (1 / U0) ∫t ∆UIP dt (4.32) ξ
1

If ∆U is measured in millivolts, U in volts and time in seconds, as is the common practice, the unit of
chargeability becomes millivolt second per volt (mV s V-1) or millivolt millisecond (ms) per volt. The
measure defined in the previous subsection as ‘polarizability’ has also been called ‘chargeability’ in
some literature, but it is desirable to use two distinct terms for the measures in equations (4.31) and
(4.32) since the physical dimensions are different in the two cases, a pure number and time
respectively.
Another definition of apparent chargeability used by some is expressed simply as the ratio
between the polarization voltage to the maximum voltage reached prior to current cut-off or Up / U0
(Figure 4.25). The values of this ratio are generally small and therefore multiplied by an arbitrary
amount such as 100 or 1000 and expressed in millivolts per volt. The direct measurement of this
ratio is made difficult by electromagnetic transients that are dominant in the first tenth of a second.
The practical definition of time domain chargeability, in terms of the decay voltage at some later
time, is only tenuously linked to the theoretical definition. Not only do different instruments use
different delays, but it was originally essential and is still fairly common to measure an area under the
decay curve using integrating circuitry, rather than instantaneous voltage.
A quantity that implicitly contains some information about the shape of the decay curve is the
ration L / M where M (chargeability) is as defined above and L is the corresponding quantity
representing the area above the curve, that is between the dashed lines in Figure 4.26 (b) and the
decay curve between t1 and t2.

25 5000 INDUCED POLARIZATION

20 4000
Chargeability (mVmsV-1)

Chargeability
Resistivity (Ωm)

15 3000

10 2000

5 1000
Resistivity

0 0 0 10N
0 5S 5N

Figure 4.27 Chargeability profile across Pyramide ore body, North West Territory, Canada

A profile of chargeability across a galena-sphalerite ore body in limestone and dolomite (from
Seigel et al., 1968) is shown in Figure 4.27 together with the resistivity profile. The resistivity is fairly
uniform at a low of about 100 Ωm between the coordinates 1S and 3N while the chargeability
anomaly is much broader, extending from 5S to 9N. This is indicative of dissemination around a very
conductive mineral zone.

84
4.5.4 Measurement of IP in frequency domain

4.5.4.1 Frequency effect

F igure 4.25 implies that if the applied current is terminated suddenly, a lower apparent resistivity,
equal to (U0 – Up) / I multiplied by the appropriate array geometric factor, would be calculated. In
the frequency domain, the apparent resistivity of the ground (see Appendix D.2) is measured by any
one of the numerous electrode configurations at two frequencies F and f (F > f). The IP frequency
effect is defined as the difference between the low frequency and high frequency resistivities divided
by the low frequency resistivity or

(FEF,f)a = [ρa(f) – ρa(F)] / ρa(f) (4.33) ξ

The term percent frequency effect (PFE) is also used if this ratio is expressed as a percentage
change in ρa. The lower frequency f (sometimes referred to as the d.c. frequency) is usually in the
range of 0.05 – 0.5 Hz and the higher one in the range of from 1 Hz to usually no more than 10 Hz in
order to minimize electromagnetic induction. To cancel telluric and SP noise ‘d.c.’ measurements are
taken with the current reversed at intervals of the order of a few seconds. There are various minor
variations of the definition of PFE.

a na a
n=2 n=3 n=4
U I I I

n=1 1020 1280 1300 220 410 870 560 685 595

n=2 1230 1130 1250 212 207 380 1130 945 1080

n=3 1200 210 380 680


1020 150 182 1770 1330

n=4 1180 117 188 202 236 478 2120 ρa


164 W 162 W 160 W 158 W 156 W 154 W 152 W 150 W 148 W

200 ft.

n=1 3.4 3.1 3.7 35 16 1.8 0.4 0.3 0.5

n=2 4.6 4.7 4.4 45 44 18 1 0.3 0.2

n=3 6.7 41 20 2.6


6.8 80 58 0.7 0.5

n=4 6.9 141 53 47 35 17 0.4 (MF)a

164 W 162 W 160 W 158 W 156 W 154 W 152 W 150 W 148 W

Concentrated
Sulphide stringers sulphide
mineralization

Figure 4.28 Frequency-


domain IP survey contour plot
in dipole-dipole configuration

85
4.5.4.2 Metal factor

A nother frequency-domain measure of IP, the apparent metal factor (MF)a, is obtained by dividing
the frequency effect (FE) or (PFE) by the apparent resistivity at the higher frequency, so that

(MFF,f)a = [ρa(f) – ρa(F)] / ρa(f) ρa(F) (4.34) ξ

The ratio on the right-hand side is often multiplied by some arbitrary constant A such as 2π x 103
or 2π x 105 to bring the quantity to a convenient of magnitude. Another measure of IP following
Equation (4.34) is

(MFF,f)a = σa(F) – σa(f) siemens per meter (S m-1) (4.35) ξ

where σa(F) and σa(f) are the apparent conductivities at the two frequencies. Since MF has the
dimension S m-1, nor arbitrary factor A is needed to give it a convenient numerical magnitude if it is
expressed as milli- or microsiemens per meter.
Metal factors emphasize rock volumes that are both polarizable and conductive and which may
therefore be assumed to have a significant sulphide content. A time-domain polarizability measure
corresponding to MFF,f would be PtT / ρa but is infrequently used.
An example of IP mapping in the time domain was given in section 4.5.3. Figure 4.28 shows an
example of a survey in the frequency domain (from Hallof, 1967). In addition to the metal factor, the
figure also shows the apparent resistivity at the lower frequency. The survey was made using a
number of dipole-dipole arrays with different values. (See Appendix E and Figure E.1 for an
explanation of dipole-dipole mapping and pseudo-sections)
The correlation between high values of apparent metal factor, low resistivities and the sulphide
mineralization is easily recognized. As mentioned in Appendix E, a plot like that in Figure 4.28 is
qualitatively regarded as representing an electrical vertical section or rather pseudo-section through
the profile, since increasing n values contain relatively more information from increasingly larger
depths, but the plot must not be taken too literally. Vertical tabular bodies of constant thickness, for
example, can be shown to give wedge-shaped anomalies on this plot. Thus, the patterns in Figure
4.28 do not by any means necessarily indicate a broadening of the mineralization towards the deeper
levels.

4.5.4.3 Phase shift

A third measure of the IP effect in the frequency domain is the phase difference between the
current injected into the ground and the voltage at the potential probes between M and N.
Maximum values of φ in practice are usually a few hundredths to a tenth of a radian at a frequency,
for example, of 1 Hz so that φ is expressed in milliradians. In these measurements, the amplitude of
the MN voltage is also measured and this provides apparent resistivity.
A voltage signal such as that measured in Figure 4.25 can be resolved by Fourier analysis into
sinusoidal components of different frequency amplitudes and phase shifts. The asymmetry of the
voltage curve implies that there will be frequency-dependant phase shifts between the components of
the applied current and measured voltage. In spectral IP surveys, these shifts are measured in
milliradians over a range of frequencies. See Appendix F for a discussion of Fourier analysis.
The frequency at which the maximum phase shift occurs depends upon the grain size, being high
for fine-grained mineralization. The sharper the peak, the more uniform the grain size. Most attempts
to distinguish between types of IP source are now based on an analysis of these spectral curves, as
gain size may be correlated with mineral type. However, exploration programs soon reach the point
at which further theoretical analysis of IP curves is less effective than drilling a few holes.
The increase in phase shift at high frequencies may also be caused by electromagnetic coupling.
In some cases simple decoupling can be achieved by taking readings at three different frequencies
and assuming a quadratic relationship (i.e. φ = A + Bf + Cf2) between phase shift and frequency. The
three readings allow this equation to be solved for A, the zero-frequency phase shift value.

86
4.5.5 Electromagnetic coupling

E lectromagnetic induction in the ground creates voltage differences that are superimposed on
those due to the true IP effect. These coupling effects can be troublesome in IP measurements
when the current supply lines are long and the ground is highly conductive. They must be assessed
before the true IP effect can be estimated.
PFE

Conductive overburden Resistive overburden


nL

F
100 L L

PFE0
ρ1 = ρ2 / 10 h=L/2
nL
+◊
L L Homogeneous
ρ2 20
h ρ1 (ρ0 = ρ2) h/L = 0
PFE0F = (ρa0 – ρaF) / ρa0 ρ2
18
ρz (Ωm, d.c.)
L (Meter) ρ2 / ρ1 = 1/10 h/L = 1/5
10 16 +
F (Hz) ◊

14
h/L = 1/2
12
+ h/L = 1

1 n= 6 5 4 3 2 1 10

+
6 ◊
h/L = 2

0.1 4 Homogeneous
+ ∇
◊ (ρ0 = ρ1) h/L = ∞
2 ⊗


∇ ⊗

⊗ ⊗
L F / ρ2
2
0
2 3 4
10 10 10 1 2 3 4 5 6

(a) (b)

Figure 4.29 PFE due to electromagnetic coupling as a function of overburden thickness in


frequency domain IP for a dipole-dipole configuration with (a) conductive overburden and (b)
resistive overburden (From Hohmann, 1973)

A universally applicable method of correcting for electromagnetic coupling does not exist. In
practice, one must assume an electrical structure for an area (e.g. a layered sequence) and calculate,
for the electrode configuration in question, the spurious IP parameter magnitude that can arise due to
EM induction. This is then subtracted from the measured value of the parameter. An approach of
this type is a matter of expediency and although used widely, its theoretical basis is qualitative at
best.
Hohmann (1973) among others has treated in some detail the problem of EM coupling in
frequency-domain methods. Figure 4.29 gives an idea of the spurious PFE effect that will arise for a
dipole-dipole configuration on the two-layered earth shown. Similar calculations for the time-domain
method have been published by Rathor (1977) and Bertin and Loeb (1976).
As a rule, it is advisable to select a measurement technique that minimizes electromagnetic
coupling effects rather than try to estimate them. For example, for a Schlumberger array L for time-
domain work should be such that µL2 / ρ << ~ 1 s, where µ is the magnetic permeability (for most
practical purposes, equal to 4π x 10-7 Ω s m-1) and ρ (Ωm) is the resistivity of the ground. For
frequency-domain work Sumner (1976) suggests that the ratio a /d where a is the electrode spread
and d = (ρ /π fµ)1/2 should be less than 0.1.
Similarly, configurations in which the current-electrodes line is at right angles to the potential-
electrodes line minimize the effect of electromagnetic coupling.ii

87
4.5.6 Negative induced polarization

F igure 4.24 shows that the discharge current, in general, flows in the same direction as the current
supplied to the ground. This means that ∆U and U in equation (4.31) are of the same sign and
the IP effect is positive. There are, however, instances when negative values for IP are observed.
These are usually geometric effects on shallow IP targets or are due to strong electromagnetic
coupling.
The geometric effect will be understood by referring to Figure 4.30. Here the positive ions on the
left side find it easier to diffuse towards the left rather than through the narrow passage that presents
a high resistance above the polarized body and similarly the negative ions on the right diffuse towards
the right. In both cases the electric current in diffusion is opposite in direction to the exciting current.
Negative IP in the frequency domain implies ρa(f) < ρa(F), a reversal that cannot be explained as a
geometric effect. It is probably due, in most cases, to electromagnetic coupling.ii

A B

Direction of + -
exciting current + -
+ -
+ -
+ -
+ -
Diffusion direction + - Diffusion direction
of positive ions + - of negative ions

Figure 4.30 Origin of negative induced polarization

4.6 Electromagnetic methods

4.6.1 Introduction

E lectromagnetic (EM) induction, a source of noise in resistivity and induced polarization (IP)
surveys, is the basis of other geophysical techniques. Wave effects are only important at
frequencies above 3000 to 5000 Hz and the methods can otherwise be understood in terms of
currents in conductors and magnetic fields in space. Because a small, good conductor within a
poorly conducting mass has a disproportionate effect on induction, discussions of EM tend to focus
not on resistivity but on its reciprocal, conductivity (σ), measured in mhos per meter or Siemens per
meter.
Electromagnetic techniques were originally developed for locating conductive ore bodies, but are
increasingly also being used for area conductivity mapping and well logging. They have the
advantage that small, very good conductors can be pinpointed in the course of surveys designed to
monitor bulk ground conductivities such as in landfill site surveys.i
If a time-varying electromagnetic field is produced on the surface of the ground, currents will flow
in subsurface conductors in accordance with the laws of electromagnetic induction. These currents
give rise to secondary electromagnetic fields that modify the total field observed at any point on the
surface. In general, the resultant field, which may be picked up by a suitable search coil, will differ
from the primary field in intensity, phase and direction and reveal the presence of conductors.
If the primary field is transient, the secondary currents and their field will decay gradually when
the primary field has ceased to exist. The decay is faster the higher the resistivity of the medium in
which the currents flow. In this case, unique phase relations are irrelative since the signal contains
an infinite number of frequencies.
A great advantage of the electromagnetic methods is that they can be successfully applied even
when conductive ground connections, indispensable for the methods of sections 4.4 and 4.5, cannot
be made owing to highly resistive (or insulating) surface formations. This is frequently the case in
arid tracts and in polar and sub-polar regions where the ground may be frozen to a considerable

88
depth. On the other hand, one of the troublesome effects in the electromagnetic methods is that the
secondary currents in superficial layers of good conductivity, e.g. clays, graphite shales, etc., may
screen the deeper conductors partially or wholly from the primary field. The latter, which are the real
objects of exploration, will then produce weak or no distortions (anomalies) in the total field and may
therefore be indetectable.ii
It should be noted that the IP method of section 4.5 also exploits time-varying fields but it differs
from EM methods in that it does not involve the phenomenon of electromagnetic induction. The basis
of the IP method is the accumulation and rediffusion of electric charges when there is a surface
separating media with electric and ionic modes of current flow.
Electromagnetic methods are similar to magnetic methods in their principles and units of
measurement. They differ in the fact that magnetic methods rely on the use of magnetometers to
measure natural magnetic fields caused by induction of the earth’s field and/or permanent (or
remanent) magnetization in rocks and other targets, whereas EM methods employ transmitter and
receiver coils to induce and measure a magnetic field. Before entering into electromagnetic theory,
the principles of magnetics are presented here.

4.6.2 Principles of magnetic methods

T he study of the effects of the earth’s magnetic field on the surface and near sub-surface has been
around since the middle-ages making the magnetic method the oldest of all applied geophysical
techniques. It is still one of the most widely used, even though significant magnetic anomalies are
produced by only a very small number of minerals.
In the year 1600, William Gilbert, physician to Queen Elizabeth I, published his book De Magnete
(abbreviated title), giving rise to the concept of a general geomagnetic field with a definite orientation
at each point on the surface of the earth. Later, around the year 1640, compasses and dip-needles,
registering local anomalies in the orientation of the geomagnetic field, were employed in Sweden to
locate iron-ores. This constituted the first systematic utilization of a physical property for locating
specific, small-scale features within the earth’s crust.
The large-scale use of magnetic measurements for investigations of geological structures, other
than those associated with magnetic ore, did not, however, begin seriously until 1915, when Adolph
Schmidt constructed his precision vertical-field balance using a magnetic needle swinging on an
agate knife-edge. Since then, magnetic observations have been successfully employed, not only in
the search for magnetite ore, but also in locating buried hills, geological faults, intrusions of igneous
rocks, salt domes associated with oil fields, concealed meteorites and buried magnetic objects such
as pipelines or archaeological sites.

4.6.2.1 Magnetic field strength, flux density and permeability

J ust as an electric field strength can give rise to an electric charge flux (current), a magnetic field
strength gives rise to a magnetic flux. The magnetic flux density, that is the flux per unit area,
also called magnetic induction, is denoted by B. If H is the field strength then

B = µH (4.36) ξ

where µ is known as the absolute permeability of the medium. In the International System of Units
(SI) H is measured in amperes per meter (A m-1). A convenient way to think of this unit is to note that
1 A m-1 is the magnetic field strength that exists at the center of a one-turn circular coil of 1-meter
diameter, placed in air and carrying a current of 1 ampere (A).
The unit of magnetic flux in the SI system is the volt second (V s), also named weber (Wb), so
that the unit of flux density B is V s m-2 (or WB m-2), which more recently is called tesla (T). For most
geophysical purposes, the tesla is too large a unit and flux densities are more commonly measured in
nanotesla (1 nT = 10-9 T). One nanotesla is exactly numerically equal to the older unit of B, the
gamma (1γ = 10-5 gauss). The distinction between H and B should be carefully noted, particularly
because B is often called magnetic field in common parlance. Associated with a B field in a medium
is the field strength H equal to B / µ.

89
The absolute permeability, being equal to B / H, has the dimension Ωsm-1 (ohm second per
meter). The absolute permeability of a vacuum is an important quantity and is denoted by µ0. Thus, a
field strength H will create in a vacuum a flux density B0 = µ0H. For most practical purposes in
geophysics the absolute permeability of air, and even most rocks, may be taken to be µ0. In this and
the following chapter, µ0 will be taken to have the value 4π x 10-7 Ωsm-1. Thus, if at any point in a
vacuum (or for practical purposes in air) we have a flux density B0, the corresponding field strength is
B0 / µ0.

4.6.2.2 Relative permeability, susceptibility and magnetization

If for a medium other than a vacuum µ = µrµ0 then from (4.36)

B = µH

= µrµ0H

= µ0H + µ0(µr – 1)H

= µ0H + µ0κH (4.37) ξ

where κ has been put equal to µr – 1, that is, µr = 1 + κ. Now, µr is a ratio of two permeabilities and
therefore a pure number. It is called the relative permeability of the medium. Similarly, κ, called the
susceptibility, is a pure number. For a vacuum, µr = 1 and κ = 0.
The susceptibility κ is very small for most natural materials and may be either negative
(diamagnetism) or positive (paramagnetism). The fields produced by such materials scarcely affect
survey magnetometers and most observed magnetic anomalies are due to a small number of
substances that are ferromagnetic or ferrimagnetic. In these, the molecular magnets are held parallel
by intermolecular exchange forces that, below the Curie temperature, are stronger than the effects of
thermal agitation. Magnetite, pyrrhotite and maghemite, all of which have Curie temperatures of
about 600º C, are the only important naturally occurring magnetic minerals and, of the three,
magnetite is by far the most common. Hematite, the most abundant iron mineral, has a very small
susceptibility and many iron ore deposits do not produce magnetic anomalies.
The magnetic properties of highly magnetic rocks tend to be extremely variable and their
magnetization is not strictly proportional to the applied field. Quoted susceptibilities are always for
earth-average field strengths.
From equation (4.37) it is noted that in order to obtain in a vacuum a flux density equal to the
density µH in the medium under consideration, an additional magnetic field strength κH would be
needed. This additional field strength, that may be said to be present at points of space occupied by
a medium subject to a field strength H, is called the intensity of magnetization M induced by H. A
body placed in a magnetic field acquires a magnetization that, if small, is proportional to the field, that
is

M = κH (4.38) ξ

Since κ is a pure number, M is also measured in A m-1. Since B and H are vectors, equation (4.37)
can be written more generally as

B = µ0(H + M) (4.39) ξ

Then, for the x,y,z components of B in an orthogonal coordinate system, we have for example

Bx = µ0(Hx + Mx) (4.39a) ξ

By and Bz would follow similarly.

90
4.6.2.3 Magnetic moment, magnetic dipole, and remanence

A n isolated magnetic ‘pole’ would, if it existed, produce a field obeying the inverse square law
which states for an electric current, for example, that all the electric current flowing in a
continuous homogeneous ground injected by a point electrode must cross any closed surface around
that point. If the surface is hemispherical and centered on the electrode, the same fraction of the total
current will cross each unit area of the hemisphere and current per unit area will be inversely
proportional to the total surface area, i.e. to the square of the radius.
The fundamental magnetic source is the dipole (Figure 4.33), but as a line of dipoles end to end
produces the same effect as positive and negative poles isolated at opposite ends of the line (Figure
4.31), the pole concept is often useful.

+m -m +m -m +m -m +m -m

2L

Figure 4.31 Combination of magnetic dipoles to form a magnet with moment 2Lm

A dipole placed in a magnetic field tends to rotate and so is said to have a magnetic moment.
The moment of the simple magnet in Figure 4.31, which is effectively a positive pole, strength m, at a
distance 2L from a negative pole –m, is equal to 2Lm. The magnetization of a solid body is defined
by its magnetic moment per unit volume and is a vector, having direction as well as magnitude.
If a body of volume V is uniformly magnetized with intensity M then VM = m (A m2) is its
magnetic moment which as stated is a vector. Imagine a particle of infinitesimally small volume but
with a very high intensity of magnetization, so that the product VM is nevertheless a finite quantity,
then you would have a magnetic point dipole. The direction of the magnetic moment of a dipole is the
dipole axis. The importance of the concept of a point dipole is that a body with non-uniform
magnetization can be considered, for the purpose of calculating the field strength as well as the flux
density due to it, to be an aggregate of an infinite number of point dipoles, each of moment
proportional to the local magnetization intensity at the point in question. Since M = κH, for a uniformly
magnetized body,

m = VM = VκH (4.40) ξ

An intensity of magnetization can exist in certain bodies (e.g. ordinary magnets) even in the
absence of an external field strength. Such magnetization is called permanent or remanent
magnetization or simply remanence. If subjected to an external H field, the body will, in addition,
acquire an induced intensity of magnetization.

4.6.2.4 Hysteresis

I f κ in equation (4.37) is independent of H, the medium is said to be a linear medium. The graph of
B against H for such a medium is a straight line as shown in Figure 4.32. In many substances (e.g.
magnetite or pyrrhotite) this is not the case. For example, in an originally non-magnetized sample of,
say, magnetite, the induction (flux density) B increases with the external field strength H according to
the curve (a), which is not a straight line. Now, B does not increase indefinitely but reaches a
saturation value Bs, for a certain field strength Hs. On decreasing H from Hs the flux density follows
the left-hand curve (b), an even when H = 0 the sample retains a flux density and is seen to possess
a remanence. It takes an external field strength Hc in the opposite direction to completely
demagnetize the sample. On increasing H further in this direction, the sample acquires a saturation
flux density –Bs and on reducing [H] again, B follows the right-hand curve (b). So long as Hs is not
altered, the sample will then continue to follow the loop formed by the curves (b) on repeated
increases or decreases of H and will never retrace the curve (a). This loop is called the hysteresis
loop of the sample.

91
B

Bs

b
b a Linear
medium
-Hc
H
Hs

-Bs

Figure 4.32 Hysteresis loop

A sample that follows a non-linear B-H relationship as in Figure 4.32 cannot be assigned a unique
susceptibility value. The κ values quoted for such substances are usually based on the nearly linear
initial part of the curve (a), where the sample is in a virgin state and the external field strength H is
weak.ii

4.6.2.5 Induced magnetism

T he magnetic method of applied geophysics depends upon measuring accurately the anomalies of
the local geomagnetic field produced by the magnetization in rock formations or other objects of
search such as archaeological artifacts. The magnetization is in part due to induction by the
magnetizing force associated with the earth’s field and in part remanent magnetization. The induced
intensity depends primarily upon the magnetic susceptibility as well as the magnetizing force, and the
remanent intensity upon the physical history of the rock or the object, for example heating and
reheating in the earth’s field, chemical changes and so on.
In accordance with the general classification used in modern physics, all substances (including
rocks and soils) fall into three categories: diamagnetic, paramagnetic and ferromagnetic. The last
one is further subdivided into the truly ferromagnetic, the antiferromagnetic and the ferrimagnetic
substances. These terms are briefly explained below.

(a) Diamagnetism: In a diamagnetic substance, κ, in equation (4.37), is negative so that the induced
intensity in the substance is in a direction opposite to the magnetic field strength and the flux density
in such a substance is less than would exist if the corresponding space were a vacuum. The origin of
diamagnetism lies in the motion of an electron around a nucleus. This motion constitutes a miniature
plane current circuit and is characterized by a magnetic moment vector as well as an angular
momentum vector, both perpendicular to the plane of the electron’s motion. In addition, an electron
has an intrinsic magnetic moment, the so-called spin magnetic moment.
An impressed field strength will tend to turn the magnetic moment in its own direction and a
mechanical torque will act on the orbital plane. The orbit reacts in a way similar to the behavior of a
top subject to a torque tending to turn its spinning axis. The result, as is the case of the top, is that
the angular momentum vector and also the magnetic moment vector begin to precess around the
magnetizing force. This is known as Larmor precession. The additional periodic motion of the
electron due to Larmor precession is such as to produce a magnetic moment opposite in direction to
the applied field.
There is a diamagnetic effect in all substances including the ‘typical’ ferromagnetics such as iron,
cobalt and nickel. However, net diamagnetism appears typically only if the magnetic moments of
atoms are zero in the absence of an external magnetizing force, as is the case for atoms or ions
having closed electronic shells, because these contain an even number of electrons, half with the
spin magnetic moment in one direction and half in the opposite direction. However, there are
exceptions to this rule. Examples of diamagnetic elements are copper, gold, lead and sulphur. Many

92
rocks and minerals show net diamagnetism; among them are quartz, marble, rock salt and anhydrite
(gypsum).

(b) Paramagnetism and the Curie-Weiss law: The susceptibility of paramagnetic substances is
positive and decreases inversely as the absolute temperature (Curie-Weiss law). Paramagnetism
makes its appearance typically when the atoms or molecules of a substance have a magnetic
moment in the absence of a field and the magnetic interaction between the atoms is weak. Normally
the moments are distributed randomly, but upon the application of the field, they tend to align
themselves in the direction of the field, the tendency being resisted by thermal agitation. The
paramagnetism of elements is mainly due to the unbalanced spin magnetic moments of the electrons
in unfilled shells, like the 3rd shells of the elements from Sc to Mn. Many rocks are reported to be
paramagnetic, for instance, gneiss, pegmatites, dolomites, syenites, etc. However, it seems certain
that their paramagnetism is not intrinsic but is a manifestation of weak ferrimagnetism due to varying
amounts of magnetite or ilmenite, or antiferromagnetism due to minerals like hematite, manganese
dioxide, etc.

(c) Ferromagnetism: In ferromagnetic materials, the atoms have a magnetic moment and the
interaction between neighboring atoms is so strong that the moments of all atoms within a region,
called a domain, align themselves in the same direction even in the absence of an external field. In
Fe, Co and Ni, this interaction takes place between the uncompensated spins in the unfilled 3rd shells
of the atoms. A state of spontaneous magnetization can therefore exist, consisting of an orderly
arrangement of the magnetic moments of all atoms. Typical of the ferromagnetics are their hysteresis
loops and their large susceptibilities, which depend upon the magnetizing force. Ferromagnetism
disappears above a temperature known as the Curie temperature. There are no known truly
ferromagnetic rocks or rock materials.

(d) Antiferromagnetism: There exist substances in which the susceptibility has an order of
magnitude characteristic of a paramagnetic (10-5) but is not inversely proportional to the temperature.
Instead, it first increases with temperature, reaches a maximum at a certain temperature, also called
the Curie point or the λ point, and decreases thereafter according to the Curie-Weiss law. In these
substances the low magnetic susceptibility below the λ point can be explained by assuming an
ordered state of atoms such that the magnetic moments of neighboring atoms are equal but directed
antiparallel to each other. Thus, the two ordered sublattices, each reminiscent of the state in a
ferromagnetic, cancel each other and their net magnetic moment is zero. This state is called
antiferromagnetism and can be confirmed by neutron diffraction studies. Of the rock-forming
minerals, hematite (Fe2O3) is the most important antiferromagnetic (λ point 675º C).

(e) Ferrimagnetism: Among the antiferromagnetic substances there is a class in which two
sublattices with metallic ions having magnetic moments are ordered antiparallel as above, but in
which the moments of the lattices are unequal, giving rise to a net magnetic moment in the absence
of a field. Such substances are called ferrimagnetic. Practically all the constituents giving a high
magnetization to rocks are ferrimagnetic, chief among them being magnetite (FeO4), titanomagnetite
(FeO(Fe,Ti)2O3) and ilmenite (FeTiO3). Spontaneous magnetization and a relatively high
susceptibility can also exist in an antiferromagnetic if stastically systematic defects are present, as is
believed to be the case for pyrrhotite (FeS). The temperature dependence of ferrimagnetics is
complex, there being theoretically several possibilities.ii

4.6.2.6 Susceptibility of rocks

T he susceptibility of rocks is almost entirely dependent upon the ferrimagnetic mineral content
(mainly magnetite, FeSO4), their grain size, mode of distribution, etc., and is extremely variable.
Sediments and acid igneous rocks have small susceptibilities, whereas basalts, dolerites, gabbros
and serpentinites are usually strongly magnetic. Weathering generally reduces the susceptibility
because magnetite is oxidized to hematite, but some laterites are magnetic because of the presence
of maghemite and remanently magnetized hematite.
The normalized SI values listed in Table 4.5 give a rough idea of the susceptibilities of some

93
common rocks, minerals and ores.

Table 4.5 Magnetic susceptibilities of common rocks and ores (x 106, negative in parenthesis)

Rock/mineral, ore Susceptibility Rock/mineral, ore Susceptibility


Marble (9.4) Rhyolite 250 – 10,000
Salt 0 – (10.3) Dolomite (impure) 20,000
Anhydrite (14.1) Granite (w/ FeSO4) 25 – 50,000
Quartz (15.1) Pyrite (ore) 100 – 50,000
Dolomite (pure) (12.5) – 44 Basalt 1,500 – 75,000
Pyrite (pure) 35 – 60 Pegmatite 3,000 – 90,000
Granite (w/o FeSO4) 10 – 65 Gabbro 3,800 – 90,000
Graphite (100) Pyrrhotite 1,000 – 100,000
Limestone 10 – 100 Dolerite 1,000 – 150,000
Greenstone 500 – 1,000 Chromite 7,500 – 1,500,000
Slate 0 – 2,000 Ilmenite (ore) 300,000 – 4,000,000
Granulite 100 – 5,000 Magnetite (ore) 70,000 – 14,000,000
Hematite (ore) 420 – 10,000 Magnetite (pure) 15,000,000

4.6.2.7 Remanent magnetism

R esearchers all over the world have confirmed that both igneous and sedimentary rocks possess
remanent magnetization in varying degrees and the phenomenon is a widespread one. Well
documented examples of rocks, igneous as well as sedimentary, occur in all parts of the world and of
all geological ages, in which the remanent intensity is not only strong but has a direction completely
different from, at times opposite to, the present direction of the geomagnetic field. Archaeological
objects such as iron or bricks from kilns also show high remanence, the direction of which may differ
from that of the local geomagnetic field at present.
Various types of permanent magnetization are now recognized. One principal type, especially for
igneous rocks, bricks from kilns, slags from ancient iron workings, etc., is thermoremanent
magnetization (TRM) acquired in cooling from high temperatures. Its orientation reflects the
orientation of the geomagnetic field prevalent at the time and place of formation. The predominant
mechanism in the acquisition of TRM is the alignment of the domains in the ferrimagnetic constituents
of the materials. It is in this respect significant that the TRM of rocks disappears when they are
heated above 600º C, which is approximately the Curie point of magnetite (Stacey and Banerjee,
1974). Other principal types of natural remanent magnetization (NRM) are: isothermal remanent
magnetization (IRM) acquired at constant temperature on exposure to a magnetizing force for a short
time; viscous remanent magnetization (VRM) acquired as a cumulative effect after a long exposure in
an ambient field, not necessarily at one and the same temperature; depositional or detrital remanent
magnetization (DRM) acquired by sediments as the constituent magnetic grains settle in water under
the influence of the earth’s field; and chemical remanent magnetization (CRM) acquired during growth
or recrystallization of magnetite grains at temperatures far below Curie temperatures.ii
Appendix G contains a brief discussion of the interpretation of magnetic remanence and a table of
remanence related values for various rock types.

4.6.3 Electromagnetic induction

4.6.3.1 Law of induction

Circular, concentric lines of magnetic field surround a current-carrying wire. Bent into a small loop,
the wire produces a magnetic dipole field (Figure 4.33) which can be varied either by alternating

94
the current (continuous wave or CW methods) or by terminating it (transient or TEM methods).
The basis of electromagnetic methods is the law of induction, the precise formulation of which is
due to F. Neumann, while the phenomenon of electromagnetic induction as such was discovered by
Michael Faraday in 1832. This law, referred to as Neumann’s law, does not describe all the
phenomena of the electromagnetic field but needs to be supplemented by other laws. The complete
system of the laws of electromagnetism was formulated by James Maxwell in 1864.

Figure 4.33 Magnetic dipole field. Dashed lines at angle tan-1 (√ 2) = 54.7º

The law of induction states that, if the magnetic flux normal to a plane circuit changes with time,
an electromotive force (e.m.f.) equal in magnitude to the time rate of change of flux is induced in the
circuit and that, if the circuit is closed, an electric current will flow in it in such a direction that its
magnetic field will tend to oppose the change of flux. A simple way to visualize this is the so-called
corkscrew rule which states that the direction of the magnetic field at a point, due to a linear current
element, is the direction in which the tip of the handle of a right-handed corkscrew tends to move
when the screw tends to advance in the direction of the current.
Instantaneous direction
Thus, if the magnetic flux normal to the of rate of flux increase
circuit in Figure 4.34 is increasing with time with time
upwards, the induced current must flow in the A
direction shown in order that its magnetic field
inside the loop is downwards to oppose the Figure 4.34
increase.
The magnitudes of induced currents are determined by the rates of change of current in the
inducing circuits and by a geometrical parameter known as mutual inductance. Mutual inductances
are large and conductors are said to be well coupled if there are long adjacent conduction paths, if
the inducing magnetic fields are at right angles to the directions of easy current flow and if magnetic
materials are present to enhance field strengths.
A change in the current in a circuit induces an opposing e.m.f. in that circuit. A tightly wound coil
which strongly resists current changes is said to have high impedance and large self-inductance.
The bearing on electromagnetic prospecting of the fact that the time-varying magnetic flux must
have a component normal to the plane of the circuit is that the transmitter in a survey must be suitably
oriented so that its primary magnetic field can be expected to cut a presumed geological conductor or
other target of search.

4.6.3.2 Phase

I nduced currents and the associated secondary magnetic fields differ in phase from the primary
field, and detected signals can therefore be resolved into components which are in phase and 90º
out of phase with the primary field (Figure 4.35). In-phase components are usually termed real, as
opposed to imaginary, quadrature or simply out-of-phase. As electromagnetic waves travel at the

95
speed of light and not instantaneously, the phase will change as the distance from the transmitter
increases or decreases. In most geophysical surveys, these shifts are trivial and usually ignored.

0º 180º or π 360º or 2π

Time

90º or π/2 270º or 3π/2

Figure 4.35 Quadrature relationship (90º or π / 2 radians phase shift)

Consider a sinusoidally varying field of frequency ν Hz with a period T sec. = 1/ν, and an angular
frequency ω = 2πν. Such a field can be represented in a plane diagram as a rotating vector where
one whole period of the field corresponds to a rotation of the vector through 360º (2π radians). If a
vector representing a field B makes an anticlockwise angle ψ in this diagram with the vector
representing a field A, then B is ahead of A in phase and goes through its maximum, minimum or any
other corresponding phase Tψ / 2π = ψ / 2πν = ψ / ω seconds before A. Therefore, if the time
variation of A is A0 cos ωt, then that of B is B0 cos (ωt + ψ).
The secondary field S at an observation point due to currents induced in the subsurface
conductors will, in general, differ in phase from the primary field P. In Figure 4.36, the lengths of the
vectors represent the amplitudes of the respective fields and P is shown in its phase ωt with S ahead
of P in phase by the angle ψ. In general, ψ can take any value between 0 and 360º. The vector sum
of P0 and S0 is the amplitude R0 of the resultant field. The instantaneous value of any field in the
diagram is given by the projection of the amplitude vector on the x axis.

S0sinψ
S0
R0
ψ
S0cosψ
α P0
ωt
P = P0 cos ωt

Figure 4.36 General phase diagram: primary field P0, secondary field S0 and resultant field R0

The projection of S0cosψ of S0 on P is the amplitude of the component of S that is in phase with
P. It is called the real component, and the component of S0sinψ, which is 90º (one-quarter period)
ahead of P, is the imaginary component. These components are often called the in-phase
component and the out-of-phase (or quadrature) component respectively. The terms ‘real’ and
‘imaginary’ which have historical and logical reasons for being so named will be used in this text. A
diagram such as in Figure 4.36 representing the phase relations between various periodic fields is
called a vector (or phasor) diagram. The following relations are obtained from the diagram by basic
geometry and trigonometry:

resultant-field amplitude, R0 = (P02 + S02 + 2P0S0 cos ψ)1/2 (4.41) ξ

resultant-field phase, α = tan-1 [S0 sin ψ / (P0 + S0 cos ψ)] (4.42) ξ

96
secondary-field amplitude, S0 = (P02 + R02 + 2P0R0 cos α)1/2 (4.43) ξ

secondary-field phase, ψ = tan-1 [R0 sin α / (R0 cos α - P0)] (4.44) ξ

real component of resultant field = P0 + S0 cos ψ (4.45) ξ

imaginary component of R and S fields = R0 sin α = S0 sin ψ (4.46) ξ

The last equation shows that the imaginary component of the resultant field is the same as the
imaginary component of the secondary field.
It is important to note that what can actually be measured in practice at an observation point is
the amplitude R0 of the resultant field in a desired direction and its phase α in relation to P. The
amplitude P0 of P at an observation point in any direction can be calculated from a knowledge of the
source-receiver geometry so that equations (4.43) and (4.44) then give the amplitude S0 of the
secondary field in the relevant direction and the phase ψ. Often, the terms ‘real’ or ‘imaginary’
component of a field will be used for brevity when actually referring to the ‘amplitude’ of the
component. In addition, the zero subscripts representing amplitudes will be omitted when the
meaning is obvious from the context.
If the x and y axes in Figure 4.36 are treated as the real and imaginary axes of a complex plane,
the instantaneous value of each field is seen to have a real and imaginary component. The
instantaneous components of R, for instance, will be R0 cos (ωt + α) and R0 sin (ωt + α) and the field
can be written, using i = √(-1), as a complex quantity in the form

R = R0 cos (ωt + α) + iR0 sin (ωt + α)

= R0ei(ωt+α)

This way of representing quantities varying sinusoidally in time is found to be very convenient in
electromagnetic theory. Alternatively, i can be replaced by –i, in which case the coefficient of –i, in
the expansion of the exponential function gives the imaginary component, not the coefficient of i.ii

4.6.3.3 Single-turn coil

T he current systems induced in geological conductors like ores, water-bearing strata, etc., by a
time-varying primary field that acts upon them are, in general, complicated but the phenomena
arising can be qualitatively understood from a study of the response of a single-turn circular coil
(Figure 4.37). This is because in many cases the complicated current systems are predominately
currents flowing along the edges or outer surfaces of the bodies, forming closed loops.

Vertical Horizontal
component component
of S of S

Secondary
field
S Instantaneous direction
of rate of increase of
primary field B with time
Figure 4.37 Single-turn [dB / dt]
coil in an alternating
magnetic field

97
Consider the case of a single-turn circular coil in a homogeneous sinusoidally varying magnetic
field of frequency ν = ω / 2π. The mathematical analysis of this circuit shows that the induced e.m.f
lags 90º in phase behind the primary field while the current in the coil, and consequently the
secondary field in the vicinity of the coil, lags behind the e.m.f. by a further angle φ = tan-1 (ωL / R),
where L is the inductance of the coil and R its resistance. Consequently, ψ = -(π / 2 + φ).
The phase relations between the primary, the secondary and the resultant fields in this case are
shown in Figure 4.38 from which it is seen that the real component of S is –S sin φ and the imaginary
component is –S cos φ.

Primary field, P
α π/2

φ
Resultant field, R
Secondary field, S

Secondary
e.m.f.
Figure 4.38 Phase relations for a single-turn loop

It will be seen that the expression for φ that a very good conductor produces a secondary field
almost opposite in phase to the primary field (R → 0, φ → π / 2) while a bad conductor produces a
field that lags 90º behind the primary one (R → ∞, φ → 0). The same phase effects are obtained on
increasing or decreasing the frequency of the primary field (or the inductance of the coil).
As with φ, the magnitudes of the real and imaginary components of the secondary field of a
single-turn loop depend on the ratio of ω and R (assuming L to be constant), rather than on ω and R
separately. They are plotted in Figure 4.39 against R / ω. This figure is interesting because it
reproduces qualitatively every essential detail of the induction response of many complicated
geological conductors to sinusoidal fields, homogeneous or otherwise. Thus, the real component of S
decreases in magnitude as the frequency is decreased while it increases if the frequency is increased
but reaches a saturation value as ω → ∞ (R / ω → 0 in Figure 4.39). On the other hand, the imaginary
component increases as the frequency is increased from very low values, reaches a maximum and
thereafter decreases with further increase in frequency.
1.0

Real component

0.5
Imaginary component

0
0.1 1 10 100
Good conductor, R/ω Poor conductor,
high frequency low frequency

Figure 4.39 Secondary-field response of a single-turn loop to a sinusoidal field

98
Figure 4.39 also shows that for a poor conductor the imaginary component will increase if the
frequency is increased while for an excellent conductor it will decrease if the frequency of the primary
field is increased. This furnishes a simple rule to judge the quality of a conductor from dual-frequency
measurements. Other simple rules may be formulated from Figure 4.39. Thus, a good conductor
produces a large real but a small imaginary component while a bad conductor produces a relatively
large imaginary but a small real component. If the conductor has a ‘medium’ resistivity, both the
components are moderately large. Quantitatively, the ratio of magnitudes ‘real’ / ‘imaginary’ is often
used, this being greater than unity for good and less than unity for bad conductors. Such rules of
thumb are generally valid only on isolated conductors. If the fields of two or more conductors in the
ground interact strongly with each other, these rules may fail.
In Figure 4.37 is shown a qualitative sketch of the horizontal and vertical secondary fields along a
line above a single-turn loop, perpendicular to the plane of the loop, on which is acting a
homogeneous, horizontal oscillating field. The real and imaginary components of each of the two
fields vary in these fashions along the line. The horizontal component directly above the loop is in the
same direction as the instantaneous direction in which the primary field is increasing and is therefore
reckoned positive. The vertical component, on the other hand, is (in this case) arbitrarily reckoned
positive upwards. The secondary-field components show a variation along the profile that is similar to
that of the respective components of the magnetic field on a plate magnetized perpendicular to its
plane. The equivalent ‘magnetization’ of the loop is in a direction opposite to the primary field
whereas for the magnetic plate it is in the same direction as the inducing field.
A loop carrying a current I has, in fact, a magnetic moment equal to IA (A m2) if A is its area, and
the direction of the moment is that of the advance of a right-handed corkscrew placed at the loop’s
center when the tips of the handle are turned in the direction of the current in the loop. At the epoch
conceived in Figure 4.37 the moment is instantaneously directed towards the left but it is oscillating
with the same frequency as the primary field lagging π / 2 + φ in phase, since this is the phase lag of
the current in the loop.ii

4.6.4 Elliptical polarization

Consider the total fields X and Y in the horizontal (x) and vertical (y) directions at any point in the
vicinity of the loop in Figure 4.37. If the primary field is B cos (ωt) and is horizontal, the
0
components are seen to be

X = B0 cos (ωt) + s cos (ωt – π/2 – φ) = C cos (ωt) + D sin (ωt) (4.47a) ξ

Y = s’ cos (ωt – π/2 – φ) = C’ cos (ωt) + D’ sin (ωt) (4.47b) ξ

where s and s’ are the amplitudes of the horizontal and vertical components of the secondary field at
that point. Eliminating ωt from equations (4.47) results in

(C’2 + D’2)X2 + (C2 + D2)Y2 – 2(CC’ + DD’)XY – (CD’ – C’D) = 0 (4.48) ξ

which is the general equation of an ellipse in an x,y coordinate system. This means that the resultant
field at any point is elliptically polarized, the vector R = √ (X2 + Y2) describing an ellipse ω/2π times
per second. In fact, the resultant field is, in general, always elliptically polarized irrespective of the
nature of the primary field or the number or nature of the secondary circuits.
When dealing with the magnetic field, the plane of the ellipse of polarization is manifest in
practice from the circumstance that no signal is induced in a receiver coil held with its plane parallel,
i.e. with its axis perpendicular, to the plane of the ellipse at the observation point. The receivers in
many modern electromagnetic systems are solenoids rather than coils. A solenoid has to be held
perpendicular to the ellipse of polarization to have no signal induced in it.
The angle that the plane of the ellipse of polarization makes with the horizontal is the dip of the
field while the angle that the major axis makes with the horizontal is the tilt of the field. That these
angles are distinct is better understood by taking the case of a vertical primary field. Then, whatever
the angle made by the secondary field with the horizontal, the plane of the ellipse will be vertical, that

99
is the dip will be 90º. However, the angle between the major axis of the ellipse and the horizontal
need not (in fact, will not) be 90º. Finally, the azimuth of the resultant field is the angle between a
specified direction (for example, North) and the line of intersection of the plane of the ellipse with the
horizontal.ii

4.6.5 Response functions

F igure 4.39 graphs the secondary field response of a single-turn coil to a sinusoidal field. The
electromagnetic response of a body is proportional to its mutual inductances with the
transmitter and receiver coils and inversely proportional to its self-inductance, which limits eddy
current flow. Horizontal-loop anomalies are expressed as percentages of the theoretical primary field
and are therefore also inversely proportional to the mutual inductance between the transmitter and
receiver. These four parameters can be combined in a single coupling factor, MtsMsr / MtrL.
Anomalies also depend on a response parameter which involves frequency, self-inductance
(always closely related to the linear dimensions of the body) and resistance. Response curves such
as in Figure 4.40 illustrate how the responses would vary over targets of different resistivity using
fixed-frequency systems or over a single target as the frequency is varied. The quadrature
(imaginary) field is very small at high frequencies, where distinction between good and merely
moderate conductors tends to disappear.

1.0
In-phase
response
0.8
τ / (l + τ )
2 2

0.6
Quadrature
response
Response

0.4 τ / (l + τ2)

0.2

0.01 0.1 1.0 10 100


Response parameter τ = 2πνl / R

Figure 4.40 Response of a horizontal-loop EM system to a vertical-loop target as a function of


the response parameter (τ). l is the loop self-inductance, R is its resistance and ν is the frequency.

Most single-frequency systems (except those used for conductivity mapping) operate below 1000
Hz and even the multi-frequency systems now in common use normally work entirely below 5000 Hz.
Narrow, poor quality conductors may produce measurable anomalies only at the highest frequency,
or not at all.i

4.6.6 Two-coil fixed source low-frequency systems in free-space magnetic fields

4.6.6.1 Biot-Savart law

T he primary field of an electromagnetic source is the field in free space (vacuum) or, for practical
purposes, in air provided any influence of the ground can be neglected. The free-space field

100
must be removed from the resultant electromagnetic field to obtain the secondary field of subsurface
conductors. It can be calculated by applying the Biot-Savart law to the source circuit. This law
states that a current element of length ds carrying a current I creates at any point P in free space a
magnetic flux density given by

dB = (µ0Ids / 4πr2)sin θ (4.49) ξ

where r is the distance between P and the element, while θ is the angle between the element and the
line joining it to P and µ0 = 4π x 10-7 Ωs m-1. The field is perpendicular to the plane contained by ds
and the joining line, and its direction is given by the corkscrew rule defined in sub-section 4.6.3.1.
The Biot-Savart law is, strictly speaking, only true for steady currents but it may be applied to
time-dependent currents up to some distance from the source depending on how low the frequency
is. This aspect will be considered in the section covering transient-field methods but in the meantime,
assume sinusoidally varying fields of sufficiently low frequency especially in the application of the
following commonly employed EM prospecting sources.

4.6.6.2 Horizontal cable on the ground

I f the cable in Figure 4.41 (a) is carrying a current I0cos(ωt), the field at a point on the same level as
the cable is vertical and equal to

P0cos(ωt) = (µ0I0cos(ωt)/4πr)(cosβ1 + cos β2) Wb m-2

= (100 I0cos(ωt) / r)(cosβ1 + cos β2) nT (4.50) ξ

where β1 and β2 are the angles shown in the figure. The formula is obtained from equation (4.49) by
appropriate integration.
If the cable is infinitely long, β1 = 0 and β2 = 0 so that

P0cos(ωt) = (200 I0cos(ωt) / r) nT (4.51) ξ

at a distance r from the cable. In a vertical plane perpendicular to the cable the field lines are circles
with the ‘cable’ as the center (Figure 4.41 (b)).
In equations (4.50) and (4.51) and all other equations that follow in this section, cos(ωt) may be
replaced by exp(iωt) or exp(-iωt) depending upon which of the conventions mentioned at the end of
section 4.6.3.2 are adopted.ii
(b) Free-space
Field lines magnetic field of a
(a) Long cable β1
long cable seen
as source ~ Ground
end-on
surface
Cable

r P
β2 Cable

Cable Cable
d4
d4 d1
d1 P
Cable

(c) Point inside a P


Cable

rectangular loop (d) Point outside a


source d2 d2 d3 rectangular loop
d3
source

~ ~
Figure 4.41 Magnetic fields for horizontal cable layouts

101
4.6.6.3 Horizontal rectangular loop

T he field at a point on the same level as the loop (Figure 4.41 (c)) is vertical and given by the
following equation obtained by applying equation (4.50) to each side of the rectangle and using
the elementary formula for the area of a rectangle

P0cos(ωt) = 100 I0cos(ωt) [(d1 / A1) + (d2 / A2) + (d3 / A3) + (d4 / A4)] nT (4.52) ξ

where the ‘A’s are the areas of the rectangles having the respective ‘d’s as diagonals. If P is outside
the loop (Figure 4.41 (d)), the last two terms within the large parentheses have a minus sign before
them.

4.6.6.4 Field of a current dipole


Coil

T his is the field of a small current-carrying


circular coil or short solenoid at distances
greater than a few times its dimensions. At any
point P (Figure 4.42), the field is entirely in the θ
plane contained by the direction of the dipole, i.e.
the axis of the coil or the solenoid, and the line Dipole Bθ
joining the dipole center to P. It can be resolved
P
into a component B//, parallel to the instantaneous
dipole direction, and a component B⊥,
perpendicular to the dipole and directed away from
it. Where m0cos(ωt) is the instantaneous magnetic B// Br
moment (A m2) of the dipole (= area of coil x
instantaneous current I0cos(ωt)), the components Figure 4.42 Current-carrying
are given by coil as a dipole

B// = (µ0m0cos(ωt) / 4πr3) (3cos2θ – 1) (4.53a) ξ

B⊥ = (µ0m0cos(ωt) / 4πr3) (3sinθcosθ) (4.53b) ξ

The lines of force for a vertically oriented dipole are shown in Figure 4.33 on page 95. The entire
diagram can be rotated through 90º to obtain the picture for a horizontally oriented dipole or through
any other angle to obtain the picture with the dipole oriented in a desired direction.
It is seen from equation (4.53b) that along the dipole axis, that is for θ = 0º or 180º (sin 0º = 0 and
sin 180º = 0), there is no field perpendicular to the dipole. There is also no such field along a line at
right angles to the dipole and through its center (θ = ± π / 2, (cos 90º = 0)). Along this line the field is
antiparallel to the dipole and is half as strong as the field at the same distance along the dipole axis.
Finally, at points for which cos2 θ = 1/3, that is tan2 θ = 2, we have B// = 0 and the field is exactly
perpendicular to the dipole. This is the case in Figure 4.33 for points lying on the dashed lines
making angles tan-1 (√ 2) = 54.7º with the dipole.
Instead of the fields parallel and perpendicular to a dipole, it is often necessary to know the field
Br at any point in the direction of the radius vector from the dipole to the point and the field Bθ
perpendicular to the radius vector (in the direction of increasing θ). These fields are

Br = (µ0m0cos(ωt) / 2πr3) cos θ Wb m-2 (4.54a) ξ

Bθ = (µ0m0cos(ωt) / 4πr3) sin θ Wb m-2 (4.54b) ξ

Appendix H provides a discussion on dielectric permittivity and calculations for several formulae
which are referred to in the following section on depth penetration of an electromagnetic wave. The
formulae are for quasi-static, near and far magnetic dipole fields.

102
4.6.7 Depth penetration

I nduced currents extract energy from the varying field and so reduce the penetration of
electromagnetic waves. The question of how deep they penetrate is very important in EM
geophysics. If the ground were perfectly insulating, the waves could penetrate to any distance.
However, owing to the finite conductivity of most surface formations and of the underlying rocks, the
incident energy is absorbed and the amplitude of the waves decreases exponentially in traversing the
conductors, due to absorption alone. In addition, there will be a geometric decrease as the wave
spreads. This decrease depends upon the character of the source, e.g. equations (H.1a) and (H.1b)
in which the field decreases as the inverse cube of the distance. It is therefore convenient to discuss
the topic of depth penetration with reference to plane waves, for which there is no such (1/πr3)
decrease.
Attenuation of a penetrating plane wave follows an exponential law (section 4.34, Figure 4.1)
governed by an attenuation constant (α) given by

α = ω{µε[(1 + σ2/ω2ε2)1/2 – 1] / 2}1/2 (4.55) ξ

where µ and ε are the absolute values of magnetic permeability and electrical permittivity respectively
and ω (=2πν) is the angular frequency. This equation can be simplified under certain limiting
conditions; for example, the conductivity, σ, in conventional EM surveys is much greater than ωε and
α becomes equal to √(µσω). The depth of investigation of an electromagnetic investigation is limited
by the skin depth, δ (Figure 4.43), which is inversely proportional to the attenuation constant (α),
unless this is significantly greater than the distance between the source and receiver.

δ = 503.29 (ρ/ν)1/2 meters


500
300
Skin depth, δ (meters, m)

1250 Hz 5 kHz 20 kHz 80 kHz

100

50
30

10

10 100 1000 10000


Resistivity, ρ (Ohm meters, Ωm)

Figure 4.43 Variation of skin depth, δ, with frequency and resistivity

The amplitude of a plane wave is reduced to 1/e of its surface value within a distance δ = 1/b
where b is defined in equation (H.2) and evaluated in section H.2 (a) in Appendix H. Following are
two important extreme cases of depth penetration calculations.
If the field is quasi-static (section H.2 (b)), that is ρεω << 1 (ρν < ~ 108 Ωms-1, assuming that the
permittivity ε = 90 pFm-1), then for ‘non-magnetic’ conductors (µ = µ0)

δ = 503.29 (ρ/ν)1/2 meters (4.56) ξ

which shows that for this case the depth penetration decreases with a decrease in resistivity and an
increase in frequency (Figure 4.43).

103
Example 4.1

If ρ = 2000 Ωm, a value typical of many glacial moraines, sandy clays, wet sandstones and chalk, the
depth penetration will be about 700 m at ν = 1000 Hz and 225 m at ν = 10,000 Hz.

However, owing to the presence of water containing dissolved salts, many surface formations
have resistivities as low as 100 Ωm and δ may be reduced to about 150 m even at ν = 1000 Hz. In
sedimentary formations, clays and shales with resistivities on the order of 1 Ωm are quite common
and in such cases frequencies as low as 10 Hz will be required to obtain a depth penetration of 150
meters. Figure 4.47 contains a nomogram relating resistivity, frequency and skin depth.
The relative response of deep-seated conductors increases and that of shallow ones decreases
as the frequency is lowered. The use of very low frequencies for deep exploration is, however,
limited by the fact that the absolute magnitude of the signals decreases more or less in proportion to
the frequency.
In the other extreme, when displacement currents are appreciable, i.e. ρεω >> 1 (ρν > 1010 Ωms-1)
for non-magnetic media

δ = 1784ε1/2ρ meters (4.57a) ξ

Taking ε = 90 pFm-1 as a representative value for the permittivity of most earth materials

δ = 16.9 x 10-3ρ meters (4.57b) ξ

In this case, δ is independent of frequency.


Note from equation (4.57b) that, if frequencies of about 10 MHz and higher are to penetrate more
than a few tens of meters, the resistivity must be at least a few thousand ohm meters. Experiments
have been reported in which radio and radar frequencies penetrated several hundred meters of rock.
From the above it is clear that these rocks must have been highly resistive. In special circumstances
(e.g. glacier ice), it may be possible to obtain a penetration of several kilometers even at radar
frequencies (Harrison, 1970).
It should be noted that if νρ lies between ~108 and ~1010 Ωms-1, it is generally preferable to
estimate δ from the exact expression obtained from equation (H.2), namely

1/δ = b = (ωµ / 2ρ)1/2 [(ρ2ε2ω2 + 1)1/2 – ρεω]1/2 (4.58) ξ

Apart from the depth penetration defined above, there is another magnitude that is often referred
to as the practical depth penetration. This is the maximum depth at which a conductor may lie and
yet give a recognizable electromagnetic anomaly. This depth depends on the nature and magnitude
of the stray anomalies (‘noise’) caused by near-surface conductivity variations, on the geometry of the
deep conductor and on the instrument noise. Experiments show that in the ideal case where the first
noise type can be neglected, the maximum practical depth penetration of the various low-frequency
portable electromagnetic arrangements is between about one and five times the separation between
the transmitting and receiving systems.

d 2d

d√2 d d√5

Figure 4.44 Spacing and depth penetration. When the two coils are moved apart, the fractional
change in distance between them is greater than between either coil and the conductor at depth. The
increased separation thus increases the anomalous field as a percentage of the primary field. In the
example, doubling the separation increases the coil to target distances by about 60%.

104
At low frequencies and low conductivities, eddy currents are small, phase shifts are close to 90º
and the bulk apparent resistivity of the ground is roughly proportional to the ratio between the primary
(in-phase or real) and secondary (quadrature or imaginary) magnetic fields. The induction number,
equal to the transmitter-receiver spacing divided by the skin depth, can then be much less than unity
for common values of ground conductivity so that the depth of investigation is then determined mainly
by coil spacing. An increase in the inter-coil spacing produces smaller proportional increases in the
distances between the target and the coils (Figure 4.44) and the coupling factor in the response
equation therefore increases. However, the resolution decreases because a larger volume of rock is
being sampled. A common rule of thumb for small targets is that the depth of investigation is equal to
half the coil separation, but this is unduly optimistic in most circumstances.

2.0 100
2 1/2
Rv(z) = 1 – 2 Z + (4z + 1)

φ = Incremental current flow (normalized)


Figure 4.45 Variation of
induced current with depth in

R = Cumulative current flow (%)


homogeneous ground for co- 2 1/2
Rh(z) = 1 – 1 / (4z + 1)
planar coil systems operating
at low induction numbers. 1.0 50
The R(z) curves show the total
current flowing in the region
between the surface and the
φv(z) = 4z / (4z + 1)
2 3/2
plane at depth as a fraction of
the total current flow. The
incremental, φ(z), curves are
φh(z) = 2 – 4z / (4z + 1)
2 1/2
normalized. Subscripts ‘v’ and 0
0
‘h’ refer to horizontal and vertical 0 0.5 1.0 1.5 2.0
dipoles. Milsom Depth / coil separation

The conduction equations for ∞ ρ = 8 Ωm ρ = 17 Ωm


layered media at low induction
100
numbers can be used in particularly
σa = 0.185 / ∞ + (0.6 – 0.185) / 8 + (1.0 – 0.6) / 17
simple ways as the current flow is
= 0 + 0.052 + 0.0235 = 0.0755
entirely horizontal, regardless of the
σa = 75.5 mSm
-1
orientation of the coil, and currents in
ρa = 13.25 Ωm
Cumulative current flow (%)

one layer hardly affect those in any


other. Figure 4.45 shows how the
current flow varies with depth for 0.6
horizontal and vertical inducing coils
over homogeneous ground. One 50
reason for preferring horizontal coils is
that the thin-layer response for
vertical co-planar coils and hence the
apparent conductivity estimate is
dominated by the surface layer. 0.185
Although the curves in Figure
4.45 relate to homogeneous media,
the independence of current flows at 0
different levels means that they can 0 0.5 1.0 1.5
be used to calculate the apparent Depth / coil separation
resistivity of any layered medium
(Figure 4.46). To some extent, Figure 4.46 Calculation of low induction number
layering can be investigated by raising apparent resistivity for a layered earth. The first layer
or lowering the coils within the zero- is dictated by the height of the coils above the ground. It
conductivity air ‘layer’. has infinite resistivity and a thickness of 1 m.

105
10-3 103
107
6
10
10-4 104
10-2 102

106
5
2000 Ωm 10
10-1 10
10-3 103

Specific resistivity, ρ (Ωm)


105
Conductivity, σ (S/m)

Frequency, ν (Hz)
Wavelength, λ (m)
10

Skin depth, δ (m)


1 1

Period, T (s)
10-2 102
104
3
10 10 10-1

10-1 10 225 m
103
2 102 10-2
10

102 103 10-3


1 1
10

104 10-4
10
10 10-1 1

10000 Hz 105 10-5

106 10-6
δ = λ / 2π = (2 / σµ0ω)1/2 = 503.29 (ρ / ν)1/2

ω = 2πν = angular frequency


µ0 = permeability of vacuum 107 10-7

Figure 4.47 Nomogram showing the relationships between specific resistivity, ρ and its
inverse, conductivity, σ with frequency, ν and its inverse, period, T in order to find depth of
penetration or skin depth, δ and its function, wavelength, λ (=2πδ).

Example 4.2 The straight line in Figure 4.47 uses the parameters of Example 4.1, namely, resistivity,
ρ = 2000 Ωm and frequency, ν = 10000 Hz to intersect the depth graph at 225 m. This is proven by
the equation for depth, δ = (2 / σµ0ω)1/2 where σ = 1/ρ = 5 x 10-4 (S/m), µ0 = 4π x 10-7 Ω s m-1 and ω =
2πν = 2π x 10000 (Hz) and the result is 225.079 m. Using the second equation, 503.29 (ρ / ν)1/2 =
503.29 (2000 / 10000)1/2 = 225.078 m.

106
4.6.8 Near-field artificial source continuous-wave methods

T he continuous-wave (CW) electromagnetic methods using artificial, also called controlled,


sources may be divided fundamentally into near-field and far-field methods. The former operate
typically at relatively low frequencies (~100 Hz to ~ 4000 Hz) and with source-receiver separations on
the order of a few tens to a few hundreds of meters, while the latter use frequencies higher than
around 10 kHz and source-receiver separations of several kilometers to thousands of kilometers. It
must be remembered, however, that what is a near-field operation in one area may, on the same
frequency, be a far-field operation in another, more conductive, area and conversely. Since CW
methods use sinusoidally varying fields, they are also called harmonic-field methods.
The principal CW methods will be discussed briefly below. Nearly all of them can be used on the
surface, in boreholes, under water or in the air, although the technical details of instrumentation and
the logistics of operation will differ in each situation.

4.6.8.1 Tilt-angle methods

T hese techniques are also known as dip-angle techniques but as the distinction has been made in
section 4.6.4 between dip and tilt they will be called tilt-angle methods here. It is the tilt of the
magnetic field vector that is measured and generally in the vertical plane through the line of
measurement and not in the vertical plane through the azimuth of the major axis of the polarization
ellipse. Thus, the tilt measured is that of the projection of the major axis of the ellipse on the
measurement plane. Unless the azimuth of the major axis deviates considerably from that of the line
of measurement, the measured tilt differs little from the true tilt; however, this potential source of
deviation should be kept in mind.
Tilt angles can be measured with an accuracy of about 0.5 – 1 % and somewhat better if the
transmitter and receiver are supported on tripods. Although the tilt-angle techniques based on
portable sources are not used now to any significant extent, the geometrical ideas based on a few of
these have a bearing on the modern very-low-frequency (VLF) method that exploits distant radio
sources and is, in one of its versions, essentially a tilt-angle method.

(a) Broadside technique: In this case, the transmitter dipole (T) and the receiver dipole (R), which
may be plane coils or solenoids, are moved simultaneously along separate, parallel lines
perpendicular to the geological strike (Figure 4.48).

T T

R R

Figure 4.48 Broadside tilt-angle technique

T is held with its axis horizontal in the line direction and therefore at right angles to the T – R line.
In the absence of conductors, the field at the receiver, being the primary field only, will be horizontal
and along the line on which R is moved. Normally, there will be no signal in R if it is held with its axis
vertical.
If conductors are present, their secondary field will cause the resultant field vector to tilt out of the

107
horizontal in general, and then R must be turned around the T – R line as axis to make the signal in it
minimum. The tilt of the field for certain geometries of conductors will be in opposite directions on the
two sides of a sheet conductor when the coil system traverses it, so that the tilt profile shows an
inflection (commonly called ‘crossover’) directly above the conductor.ii

(b) Shoot-back technique: Another tilt-angle technique (in-line) in which T and R, instead of being on
separate lines as above, are on the same line, one behind the other. It suffers from the fact that it
gives ‘false’ tilt angles if there is a height difference between T and R in contrast to the broadside
technique in which the normal field at the receiver position is always horizontal. A version of the in-
line method, devised by J.D. Crone, largely eliminates the effect of topography and is used in
reconnaissance surveys. The procedure is as follows.
The coils are constructed so that either can be used as a transmitter or a receiver. The axis of
coil 1 is directed towards 2, but downwards at an angle of 15º to the horizontal. Coil 1 transmits on
the selected frequency and coil 2, acting as a receiver, is turned (for maximum signal) around a
horizontal axis perpendicular to the line 1– 2. The tilt α1 of the field is noted. Thereafter, coil 2
transmits but with the axis pointing 15º upwards. The tilt α2 of the field of the field is measured at coil
1, now acting as a receiver. In the absence of conductors, α1 and α2 are very nearly equal regardless
of the elevation difference between the coils so that any difference α1 – α2 is taken as a measure of
the anomaly due to a subsurface conductor.

(c) Other variations: Variations of the tilt-angle technique using artificial sources include the use of
fixed large loops or grounded cables, instead of dipole transmitters, as sources. In one version, a
large transmitter loop is set up in the vertical plane and the receiver is moved along a line of
measurement that is relatively distant from the transmitter. The axis of the transmitter is directed
towards successive receiver positions while the tilt is measured. When a large loop or a long cable is
laid on the ground, the normal orientation of the receiver dipole axis for no signal will be horizontal as
the normal field at it is vertical. The tilt (or the co-tilt) of the field form the vertical shows a maximum
above a steep conductor, since the secondary horizontal field is strongest here, and not a crossover
as when the normal field is a horizontal one.ii

4.6.8.2 Fixed-source systems exploiting amplitude and phase

T he tilt of the magnetic field vector that is measured in the tilt-angle methods is the result of the
different directions of the primary and secondary fields. The vector does not tilt if the two fields
are in the same direction, whatever their phase difference, but when there is a tilt, its magnitude also
depends on the phase difference besides the angle between the primary and secondary fields.
However, this additional information cannot be obtained from tilt-angle measurements alone.
Following are methods that aim at a more complete measurement of the field than afforded by the tilt-
angle methods. There are two basic variations of these methods: fixed-source systems and moving
source-receiver systems.

(a) The compensator or Sundberg method: The primary layout in this method that was introduced by
Sundberg (1931), consists of a straight cable, some 400 – 4000 m (or more) long, grounded at both
ends, through which an alternating current of low frequency (< 1000 Hz) is passed. A large horizontal
loop, usually rectangular, may also be laid on the ground instead of the cable. The cable (or the long
side of the primary loop) is generally placed approximately parallel to the geological strike in the area
and the electromagnetic field is investigated at regular intervals along lines perpendicular to the
strike. In the case of the loop, observations can be made inside as well as outside the loop.
For many, if not most, purposes it would be sufficient to measure the amplitude and phase of
either the vertical or the horizontal component of the resultant field, but measurements of both may
be called for in detailed work. A search coil consisting of several turns of copper wire on a suitable
frame is held horizontally (to measure the vertical component) or vertically (to measure the horizontal
component) and the voltage induced in it is compared with a reference voltage. The latter is obtained
from an auxiliary (‘feeding’) coil stationed near the primary layout. A method has also been tried in
which the reference voltage is instead transmitted by an ultra-high frequency (UHF) carrier wave

108
modulated at the frequency of the primary current.
The comparison of voltages is made on a compensator, essentially an a.c. potentiometer, in
which the inclusion of a reactive element enables one to determine phase differences. The real
component is balanced by the voltage across the resistor R and the imaginary one by that across the
variometer I, the balance being indicated by a zero signal in the detector. The real and imaginary
signals in the receiver can be expressed after calibration as nanotesla per ampere primary current. In
older papers, the field was expressed as microgauss per ampere (1 µG = 0.1 nT).
The phase of the received field with respect to the primary field is given by tan-1 (imaginary / real).
However, circuits can be devised that give a phase reading directly. The normal primary field of the
layout must be subtracted from the observed real component. The imaginary component needs no
correction. These considerations are valid provided the field is linearly polarized. However, as seen
in section H.2, at appreciable distances from the layout the field may be elliptically polarized and a
simple correction for the normal field may not be sufficient.

(b) The Turam method: The compensator method requires a direct connection between the primary
layout and the observation point and therefore becomes cumbersome when large areas are to be
covered. This operational disadvantage is overcome in the Turam (Swedish: ‘two coil’) method
devised by Hedström (1937).
The primary field is produced as before by a long cable or a large loop and two receiver coils
(dipoles), 10 to 20 meters apart, are carried along the line of measurement. For each position of the
coils, the ratio of the amplitudes and of the phase difference between the voltages induced in them
are measured on a bridge-type compensator, the former with an accuracy of about 0.005, the later
about 0.02º.
The coils are usually held with their axes vertically (so as to compare the vertical components of
the resultant field) but sometimes the axis of one may be oriented vertically and that of the other
horizontally, to measure the horizontal field in the latter. In this section, the coils are considered to be
held with their axes vertical.
The quantities measured in Turam work are V1 / V2, V2 / V3, … and α2 – α1, α3 – α2, …, where the
‘V’s are the amplitudes and the ‘α’s the phases of the vertical electromagnetic field at the stations
1,2,3, …, etc. To correct for the variation of the primary field (p) with the distance from the source,
the measured ratios are divided by the normal amplitude ratios p1 / p2, p2 / p3, …, etc. This is
particularly easy in the case of the long cable since the normal ratio of the vertical fields at any two
points is simply the inverse ratio of the distances of the points from the cable, unless one or both of
the coils are on a level different from the cable, when an easily calculable correction must be applied.
The normalized or reduced ratios V1p2 / V2p1, V2p3 / V3p2, … will all be equal to unity in the
absence of subsurface conductors.
The conclusion that the normal phase differences are zero provided the ground is non-conducting
is not strictly true because the phase of an electromagnetic field varies by 2π over a distance of one
wavelength. However, the wavelength in air of the low-frequency primary field used in geoelectric
work is on the order of several hundred kilometers. This is very long compared with the distance from
the cable of a kilometer or so, up to which the measurements can usually be made, so that the
normal phase changes, although not strictly zero, are quite negligible.
On conductive ground, however, the phase may change appreciably within relatively short
distances. Moreover, the field will be elliptically polarized, the inclination of the ellipse becoming
increasingly horizontal away from the cable. In accurate and detailed work, these effects must be
taken into account in order to correct as far as possible for the conductivity of the bedrock and the
overburden. Deviations of the reduced ratio from unity and the phase difference from zero indicate
anomalous subsurface conditions.

4.6.8.3 Moving source-receiver systems

(a) System description: In both CW and TEM surveys, sources (usually) and receivers (always) are
wire loops or coils. The shapes of anomalies depend on the geometry of the system as well as the
nature of the ground conductor. Coils are described as horizontal or vertical according to the plane in
which the windings lie. The axis of a horizontal coil is therefore vertical and is alternatively described
as vertical dipole.

109
Systems are also characterized by whether the receiver and transmitter coils are coplanar,
coaxial or orthogonal (at right angles to each other), and by whether the coupling between them is a
maximum, a minimum or variable (Figure 4.49). Coplanar and coaxial coils are maximum-coupled as
the primary flux from the transmitter acts along the axis of the receiver coil. These systems are
scarcely affected by slight coil misalignments, but are, because a strong in-phase field is detected
even in the absence of a conductor, very sensitive to changes in coil separation. Accurately
orthogonal coils are minimum-coupled and the primary field is not detected.
Systems in which the receiver coil is rotated to determine the dip of the resultant field were once
very popular, but are now generally limited to very low frequency (VLF) work and to shoot-back
systems in which topographic effects are cancelled by interchanging the receiver and transmitter.

Maximum-coupled

Horizontal coplanar r
(vertical dipole)

Vertical coplanar
r
(horizontal dipole)

Vertical coaxial r

Minimum-coupled
(orthogonal)

Cross-planar r

Figure 4.49 Coil orientations in various moving source-receiver systems

(b) Field procedure: The configuration for moving source-receiver or tandem methods is shown in
Figure 4.50. A battery-operated portable oscillator (1 – 2 W) delivers an alternating current to a
transmitter coil or solenoid. The receiver is spaced at a fixed distance from the transmitter, usually
between 25 and 100 meters, and is connected to a compensator which is also fed by the reference
voltage from the transmitter. The layout is, in principle, the same as that in the Sundberg method
except that the transmitter is now portable and movable.
The most common system in practice utilizes horizontal coplanar coils, with a shielded cable
carrying a phase-reference signal from the transmitter to the receiver. The Swedish term Slingram is

110
often applied to horizontal-loop systems, but without any general agreement as to whether it is the
two mobile coils, their horizontal co-planar orientation or the use of a reference cable which makes
the term applicable.

Transmitter Receiver

25 – 100 m

Feeding coil

Cable

Oscillator Compensator

Figure 4.50 Moving source-receiver system configuration

Figure 4.51 Slingram EM-31 in an archaeological survey in Copan, Honduras

111
15
40
25 37
35
33
20
30
28
26
15
25
40
23

10 21
35 20
29
22 18
5 17
25 16
15
18
0 14
0 5 10 15 20 25 30 35 40 45
(mS/m)
Apparent
0 10 20 30 40
Conductivity

Apparent conductivity map


Northwest platform, Copan
Slingram – EM-31

July 12, 2000

Figure 4.52 Apparent conductivity map of archaeological survey in Copan, Honduras

The Geonics EM-31 (Figure 4.51) is an example of a coplanar coil instrument that can be
used, sometimes with difficulty on steep slopes, by one operator to obtain rapid estimates of apparent
conductivity and inversely, apparent resistivity. Figure 4.52 shows the results of a detailed survey in
an archaeological site in Copan, Honduras. The contour plot, originally in color, indicates an anomaly

112
at coordinates 35.1, 12 with an apparent conductivity of about 40 mS m-1. The ‘color’ bar is coded as
a natural spectrum ranging from ultraviolet - blue at the bottom to red - reddish brown at the top. The
corresponding colors in the contour plot traverse from blues and greens on the left to yellows,
oranges, reds and finally, reddish brown at the anomaly. Eventual excavation of the site uncovered a
previously unknown Mayan structure.
In normal surveys, the coils are held horizontally, giving, at low induction numbers, a penetration
of about 6 m and a radius of investigation of about 3 m with a fixed 3.7 m coil spacing. This
compares favorably with the 20 to 30 meter length of the Wenner array (section 4.4.1.8) with similar
penetration. Measurements can also be made with the coils vertical, halving the penetration depth.
The EM-31 operates at 9.8 kHz. The two-person Geonics EM-34-3 uses frequencies of 6.4, 1.6,
and 0.4 kHz with spacings of 10, 20 and 40 m respectively, i.e. the frequency is quadrupled each time
the coil separation is halved. The in-phase signal is also used to monitor spacing. Penetrations are
15, 30 and 60 m for horizontal coils and 7.5, 15 and 30 m for vertical ones. Both EM-31 and EM-34-3
are calibrated to read terrain conductivity directly in millimhos per meter.
In moving source-receiver systems, the field acting on the receiver is measured in percent of the
primary field present at it when the system is in free space. Where R0 is the amplitude of the field
picked up by the receiver and following the notation of Figure 4.36, it can be written as

R0 cos(ωt + α) = P0 cos(ωt) + S0 cos(ωt + ψ) (4.59) ξ

where P0 cos(ωt) is the appropriate free-space field at the selected nominal distance from the
transmitter. The real and imaginary components of the resultant field are P0 + S0 cosψ and S0 sinψ
respectively. The instrument is calibrated to read 100 [1 + (S0 / P0) cosψ] in percent units on the real-
component display and 100 (S0 / P0) sinψ in percent units on the imaginary-component display. In
some units, the normal primary field at the receiver is ‘cancelled’ and the real-component display
reads the anomaly 100 (S0 / P0) cosψ.
In principle, any mutual orientation of the transmitter and receiver may be used but most surveys
are carried out with both dipoles either horizontal or vertical. The common coil orientations are shown
in Figure 4.49. The transmitter-receiver layout is moved as a whole along their joining line and
readings taken at suitable intervals. For horizontal coplanar coils and vertical coaxial coils, the
direction of movement is perpendicular to the geological strike to ensure that the primary field cuts
any conductors that generally follow the geology. In ground work, vertical coplanar coils should be
moved along lines parallel to the strike as otherwise the reference cable will have to cut across trees,
bushes etc.

(c) Effects of coil separation: Changes in coupling between the transmitter and receiver can produce
spurious in-phase anomalies. The field B at r distance from a coil can be described in terms of radial
and tangential components Br and Bθ respectively (Figure 4.42). The amplitude factor M depends
upon the coil dimensions and current strength. If the coils are coplanar, Br is zero because θ is zero
and the measured field, B is equal to Bθ. The dipole equations, Br = 2M sinθ / r3 and Bθ = M cosθ / r3
which sum to B0 = M (1 – 3 sin2θ) / r3 then imply that, for a fractional change x

B = B0 / (1 + x)3 (4.60a) ξ

where B0 is the field strength at the intended spacing. If x is small, this can be written as

B = B0 (1 – 3x) (4.60b) ξ

Thus, for small errors, the percentage error in the real component is three times the percentage error
in distance. As real anomalies of only a few percent can be important, care must be taken to keep
the separations constant.

(d) Typical conductor responses: The response of a conductor in free space or in a poorly
conducting environment can be visualized in a qualitative manner by elementary arguments.

113
Consider, for instance, a long, thin vertical plate below the plane of the transmitter and the receiver,
both of which are horizontal coils (vertical dipoles), the plate being perpendicular to the transmitter-
receiver line (Figure 4.53 (a)). Suppose that at a particular instant the oscillating transmitting dipole is
directed downwards so that the primary field cuts the plate as shown. The secondary currents in the
plate will be in the plane of the figure at the top edge of the plate and out of it at the bottom. At any
point R1 on the same side of the plate as the transmitter, the secondary field will be directed upwards,
that is, in the same sense as the primary field at it. At a point such as R2, however, the two fields will
be in opposite directions and the resultant field will be less than the primary. Furthermore, if either
the transmitter or the receiver is directly above the plate, the field indicated by the compensator will
be equal to the primary field.

Transmitter Primary
coil field

R1 R2 (a)

Dipole Primary x
field
Secondary
Primary field
field

Conducting
plate
x coordinate is the midpoint
of the T – R line

T R
T R

110 %
(b) (c)
90 %

Figure 4.53 Origin of the electromagnetic anomaly in the moving source-receiver system

Thus, as the horizontal-coils transmitter-receiver system with a fixed spacing is carried across the
plate, the field picked up by the receiver will have the appearance as in Figure 4.53 (b). The
maximum damping in the field (negative anomaly) is obtained when the midpoint of the system is
directly above the plate.
Similarly, if both coils are held vertically and are coaxial, the field will, under certain conditions,
have the appearance as in Figure 4.53 (c), there being a maximum ‘augmentation’ of the primary field
(a positive anomaly) when the plate is midway between the transmitter and the receiver. However,
the response to this coil configuration varies in a rather complicated manner with the depth to the top
edge of the plate and is generally difficult to predict correctly from simple arguments alone.
Figure 4.54 shows the result of taking a horizontal loop system across a thin, steeply dipping
conductor. No anomaly is detected in (A) by a horizontal receiving coil immediately above the plate
because the secondary field there is also horizontal. Similarly, there will be no anomaly detected in
(C) when the transmitter coil is vertically above the conductor because no significant eddy currents
will be induced.

114
The greatest (negative) secondary field values will be observed when the conductor lies mid-way
between the two coils as in Figure 4.54 (B). Coupling depends upon the target orientation and lines
should be laid out across the expected strike. Oblique intersections produce poorly defined
anomalies that may be difficult to interpret. As a rule, readings obtained with moving source-receiver
systems are plotted at points mid-way between the two coils.

Anomaly (%)
A B C Coil positions

Secondary field horizontal


Tx Rx No horizontal-coil anomaly
A

Secondary field
opposes primary
Tx Rx
B

Tx Rx
C
Primary field in plane
of conductor, No
secondary currents

Figure 4.54 Effect of coil position on anomaly detection

Environmental factors can also produce inaccurate results. Examples of these are actual non-
target conductors, roadsides where artificial conductors are often buried and power and telephone
lines which emit 50 to 60 Hz noise that may, if close enough to the survey area, pass through the
built-in rejection (notch) filters. Ground conditions should also be noted, as variations in overburden
conductivity can drastically modify the anomaly shape as well as signal penetration. In hot, dry areas,
salts in the overburden can produce surface conductivities so high that CW methods are ineffective
and have to be superseded by TEM.

4.6.9 Transient-field methods (time-domain electromagnetism)

4.6.9.1 Introduction

E lectromagnetic energy can be supplied to the ground by transient pulses instead of by continuous
waves which have been discussed up to now. When a circuit (e.g. a large circular or rectangular
loop) is carrying a steady current, the constant magnetic flux due to it passes through any electrical
conductor that may be present in the environment but no secondary currents are induced since the
flux is not varying with time. If the steady current is suddenly switched off, the flux falls from its initial

115
steady value to the value zero but during a short interval it is decidedly varying with time and currents
will therefore be induced in a conductor in accordance with the law of induction (section 4.6.3.1).
Their secondary field decays with time as the currents gradually dissipate because of the electrical
resistance in the conductor. This time-dependent field will, in its turn, induce a transient electromotive
force in a receiver circuit on which it may be acting.
In practice, the primary current and flux cannot be switched off instantaneously, but only more or
less suddenly. A commonly used procedure is to bring the flux from its steady value B0 to the value
zero linearly with time within a short interval τ of the order of microseconds. The rate of change of
flux, B0 / τ, and hence any induced electromotive force due to it, will, in this case be proportional to
the initial steady current in the primary circuit. Hence, the receiver signal in the transient-field EM
(TEM) method is expressed as microvolts per ampere (µV/A) primary current.

4.6.9.2 Comparison of transient-field to induced polarization methods

T he electromagnetic methods of continuous-wave (CW) and particularly transient-field (TEM) bear


a superficial resemblance to the induced polarization (IP) which also uses time-varying current
but differ in the fact that they (EM methods) are based on the phenomenon of current induction and
not, as in the IP method, on the diffusion of charges accumulating on certain surfaces by the
electrochemical action of the current. The currents in IP surveys are introduced directly into the
ground and not by magnetic fields generated by electromagnetic coils which in turn induce currents.
However, at least one IP method does use induction and a more fundamental difference lies in the
time-scales.
Time-domain IP systems usually sample after delays of between 100 milliseconds and two
seconds, and so avoid most EM effects. There is a small region of overlap, from about 100 to 200
milliseconds, between TEM and IP systems and some frequency-domain or phase IP units are
designed to work over the whole range of frequencies from d.c. to tens of kHz to obtain conductivity
spectra. However, it is possible to regard the EM and IP phenomena as completely separate and to
avoid working in regions, of either frequency or time delay, where both are significant.

4.6.9.3 Transient field of a single-turn circular loop

Suppose that the primary field in Figure 4.37 is a steady field and it is suddenly switched off. Then
the induced current I in the loop follows the equation

RI + L(dI/dt) = 0 (4.61) ξ

where t is the time, R the resistance and L the self-inductance in the loop. Integrating gives

I = I0 exp [-(R/L) t]

where I0 is the current at time t = 0, that is, immediately after the primary field is switched off. As the
magnetic field of a current is proportional to the current, the signal induced in a receiver due to I will
also decay exponentially with time and may be written as

S = S0 exp [-(R/L) t]

A graph of lnS against t in this case will be a straight line of slope – R/L, so that L/R is the time
constant, that is, the time for the signal to fall to 1/e (Figure 4.1) of its initial value.
If R is large, the slope is steep, the time constant is small and the signal vanishes quickly. If R is
small, the signal persists for a relatively longer time. This is essentially true even when there is a
massive conductor, rather than a single-turn loop, and since it is inappropriate to speak of a
resistance in the conductor, the resistivity (or its inverse, conductivity) of the conductor must be
considered. For such cases, into which category geological conductors also fall, the lnS against t plot
is not a straight line, but the rate of fall of the ‘signal versus time’ curve nevertheless indicates the
quality of the conductor.

116
The inductance of a massive conductor is related to its size. Broadly speaking, the bigger the
conductor, the larger is the inductance. Thus, for the same resistivity, large conductors have longer
time constants than small ones.

4.6.9.4 TEM survey procedures

T he transmitter in a transient-field electromagnetic (TEM) survey is generally a square or


rectangular loop (Figure 4.55 (a)), the square one being the more common of the two. Typical
dimensions are 80 m x 80 m or 200 m x 200 m although smaller and larger loops have also been
used. In principle, a long grounded cable could also be used as a source but there are difficulties in
injecting pulses of sufficient energy in such a source. It is interesting to note that, since transient-field
measurements are made after the primary field has been turned off, the transmitter loop itself can be
used as a receiver (Figure 4.55 (b)).

(a)

Tx

Rx

(b)

Tx
Rx

Figure 4.55 Transient-field (TEM) measurement layouts:


(a) coincident but separate loops, (b) one loop as transmitter
(Tx) and receiver (Rx).

A typical procedure of measurement using a single square loop as transmitter and receiver is as
follows. A known current is sent for, as an example, 10 seconds into the loop and switched off. The
loop is automatically switched to a detecting device as soon as the primary field has ceased to exist.
The signal voltage in the loop is sampled at suitable predetermined intervals such as from 4 ms to
200 ms, divided by the primary current, and the values (µV A-1) for different times are stored in the
memory of a microcomputer.
A second primary pulse is sent after a no-current period of 10 s and the signals received are
added to the corresponding stored values. The current direction in the transmission is usually
reversed for each pulse to avoid effects due to ground polarization. In this case, the sign of the
received signal is reversed for every second pulse before stacking. The procedure of stacking
(section 2.5) is repeated a desired number of times and the final series of values is printed out after
dividing it automatically by the number of stackings. The loop is then moved, parallel to one of its
sides, by a distance equal to that side.

4.6.9.5 Comparison of TEM and CW methods

T he transient-field and the continuous-wave electromagnetic methods are related to each other
through the Fourier transform (Appendix F) according to which a transient pulse can be
decomposed into an infinite number of sine and cosine waves. The transient signal received is a
superposition of the secondary-field response of a conductor to each of these frequency components.
Hence, a transient-field measurement over a sufficiently long time interval implicitly contains roughly
the same information as a multifrequency measurement with the same layout and vice versa.
The transient current system induced by a pulse in a circular transmitter loop laid horizontally on

117
a homogeneous, conductive ground may be approximated, immediately after the pulse is over, to a
single circular transient current at the surface, coincident with the loop but flowing in a direction
opposite to the primary current before it was switched off. Physically, this ring is the path in the
ground along which the induced current density is maximum at that instant. Along rings inside and
outside this ring, the current density is smaller. This shallowest transient current system will induce
another system below it and so on. In each of these systems, the ring of maximum current density is
larger than the previous one. It should be clearly realized, however, that there is no current flow
‘outward’, that is in the radial direction from the axis of the transmitter loop. All the induced electrical
field and the current density is purely azimuthal in this case.
The response of the ground may now be considered qualitatively to be the diffusion of the ring of
maximum current density into the ground with the velocity of electromagnetic waves, accompanied by
a broadening at the same time, somewhat like a smoke ring puffed into the air from a cigarette. From
this it can be inferred that the signals measured at later times are governed by the part of the ground
lying at depths greater than the depths that govern the earlier signals. From section 4.6.7 it is evident
that a high frequency is damped out at a shallower depth than a low frequency. Thus, qualitatively
speaking, the early-time response in the transient-field method is analogous to the high-frequency
response of the continuous-wave method while the late-time response corresponds to the response
at low frequencies. Theoretically it is not possible, however, to attach a difinite frequency to a definite
time and vice versa. The equivalency should only be looked upon in broad terms.
Transient methods possess certain advantages. It is evident from section 4.6.7 that either very
great primary field strengths at high frequencies or, if the field strengths are moderate, very low
frequencies must be used in the continuous-wave methods, if subsurface conductors below highly
conductive overburden are to be excited. In either case, the instrumentation becomes very difficult,
cumbersome and costly. The production of very high transient field strengths, on the other hand, is
easier although the difficulties involved should by no means be underrated. Experiments have been
reported in which a power of 700 kW was dissipated in 100 ms through a square single-turn loop of
300 m sides (Dolan, 1970).
Several instances have been reported of sulphide ore bodies being detected by transient-field
methods below highly conducting overburden. Another advantage of transient-field methods is that
the exact geometry of the source with respect to the receiver is largely immaterial since the
measurements, carried out in the absence of a primary field, are purely those of the secondary field
and no knowledge of the primary field strength is required to obtain the secondary field. In this
respect, transient-field measurements are exactly comparable to those of the imaginary component or
the phase shift in continuoous-wave methods. It follows that the effect of topographic irregularities is
often negligible in transient-field measurements.
The main disadvantage of TEM methods are a relatively slow speed of operation and the
proneness of the receiving circuit to electrical noise since the circuit has to respond to a broad band
of frequencies and cannot be tuned to a particular frequency in order to enhance the received signal,
as can be done in continuous-wave equipment.

4.6.10 VLF (far-field) methods

G eophysical instruments have been developed which take advantage of high-power military
communications transmissions in the 15 – 25 kHz band. Termed very low frequency (VLF) by
radio engineers, these waves have frequencies higher than those used in conventional geophysical
work, but allow EM surveys to be carried out without local transmitters.
An EM wave consists of coupled alternating electrical and magnetic fields, directed at right angles
to each other and to the power vector which defines the direction of propagation. Electric field
vectors align themselves at right angles to perfect conductive surfaces and a wave can therefore be
guided by enclosing conductors. The extent to which this is possible is governed by the relationship
between the wavelength of the radiation and the dimensions of the guide. Waves at VLF frequencies
propagate efficiently over long distances in the waveguide formed by the ground surface and the
ionosphere. Neither the earth nor the ionosphere is a perfect conductor and some VLF energy is lost
into space or penetrates the surface. Without this penetration, there would be neither military nor
geophysical applications of this method. As it is, the waves can be detected tens of meters below the
sea surface and are ideal for communicating with submerged submarines. Amplitudes decrease

118
exponentially with depth and the secondary fields produced in subsurface conductors are similarly
attenuated on their way to the surface, i.e. VLF surveys are skin depth limited. The relationship for a
20 kHz wave was shown in Figure 4.43.
The study of VLF waves dates back to 1933 (cited by Pedersen et al., 1994) when observations
were made of the ground effect on radio waves. The U.S. Navy used VLF radio signals to
communicate with and among their vessels during World War II. The application of distant radio
transmitters to geophysical prospecting dates back to about 1965 (Paál).

4.6.10.1 VLF radio transmitters

T he electromagnetic methods discussed up to now operate essentially in the fear-field region. The
effects of the intermediate region may at times manifest themselves, as has already been hinted
at in connection with fixed-source systems, but the far-field region (section H.2 (d)) is almost never
mentioned in these methods.
When an oscillating source radiates electromagnetic waves, the far field dominates beyond a few
wavelengths from the source. In the previously mentioned methods, the source is not generally
powerful enough for measurements at large distances. Such measurements, however, are possible
with the use of radio transmitters.
Radio transmitters throughout the world transmit modulated or unmodulated carrier waves for a
variety of purposes such as ordinary radio and television broadcasting, fax, telex, time
synchronization signals, aircraft and marine communications, etc. By far, the world’s military is the
most prolific user of transmitters of VLF radio waves that contain (usually coded) messages
superimposed by frequency modulation on a sinusoidal carrier wave or chopped into dots and dashes
for Morse and other codes. It is possible to measure the electric and magnetic fields of these
transmissions at distant points by tuning in to their frequencies. Table 4.6 lists seventeen of over
twenty military VLF stations around the world with their respective frequencies in kHz and power
outputs in kW.

Table 4.6 World-wide VLF radio transmitters

Freq. Power Freq. Power


Station Location Station Location
(kHz) (kW) (kHz) (kW)
15.1 300 FVO France 18.9 16 PKX New Guinea
15.5 1000 NWC North West Cape, Australia 20.7 60 UFT France
16.0 750 GBR Rugby, England 21.4 400 NSS Virginia, USA
16.4 350 JXZ Norway 23.4 600 NPN Hawaii, USA
17.0 310 RUR Russia 23.6 70 LPZ Argentina
17.1 200 UMS Russia 24.0 1000 NAA Cutler, USA
17.4 50 NDT Japan 24.0 150 NBA Panama
18.1 315 UPDS Murmansk, Russia 24.8 230 NLK British Colombia, Canada
18.3 undscl 3SB Datang, China

4.6.10.2 VLF transmitters and the distant field in air

T he antenna of a VLF transmitter constitutes an oscillating electric dipole that sends out electric
and magnetic fields. The magnetic field of a vertical magnetic dipole has been considered in
previous pages but not its electric field which is also emitted. In this section, the magnetic field
strength H (A m-1) will be referred to as opposed to the flux density B = µH (Wb m-2) that was used in
the discussion of near-field methods.
In free space of permittivity ε and magnetic permeability µ, the electric field of an electric dipole of
moment p0exp(-iωt) is obtained from the same two equations (H.1a) and (H.1b) for the magnetic field
of a magnetic dipole by replacing B on the left-hand side by E, and µm0 on the right-hand side by p0/ε
and the remainder of Appendix H applies, duly noting these changes.
However, in contrast to free space, the propagation of VLF waves from an electric dipole on the

119
actual surface of the earth is a very complex phenomenon involving a ground wave, a space wave
and a wave guided by the ionosphere and the earth’s surface. It will be sufficient, however, to note
that at large distances from the transmitter, the field of a dipole may, for practical purposes, be
considered to be uniform within a small area. The structure of the field at such a point is shown in
Figure 4.56 where the ‘transmitter bearing’ (the propagation direction of the ground wave) is
designated as the x direction. It will be seen that the incident magnetic field strength has only the
horizontal component Hiy perpendicular to the transmitter bearing and that there is no vertical
magnetic field. The incident electric field has a horizontal component Eix in the direction of the
transmitter bearing and a vertical component Eiz but no component in the y direction.

Eiz

VLF transmitter
antenna
x Eix

Hiy
y

Figure 4.56 VLF field at a distant point

It was shown several decades ago (Norton, 1937) that on homogeneous, conductive ground, Eix
leads Eiz in phase by 45º, so that the vector E is elliptically polarized. Furthermore, along the ground,
at distances beyond about one wavelength form the antenna, the angle λ that the major axis of the
ellipse makes with the vertical, and the ratio of the ellipse axes on such an earth, remain virtually
constant at all points and Ez >> Ex. The exact expression for λ is complicated. Fortunately, since ρεω
<< 1 (section H.2 (b)) in geophysical applications, it can be simplified and then gives for the field in air
at the surface

tanλ = (πρε0ν)1/2 = 0.527 x 10-5 (ρν)1/2 (4.62) ξ

where ε0 = 8.852 pF m-1 (section H.1). It can be shown that even if ρ = 10 kΩm and ν = 20 kHz, λ is
barely 4.5º, so that the major axis is practically vertical.

4.6.10.3 VLF field effects

(a) Detecting VLF fields: Geophysical users of a VLF signal have control over neither the amplitude
of the signal nor its phase. Readings of a single field component at a single point are therefore
meaningless; one component must be selected as a reference with which the strengths and phases
of the other components can be compared. The obvious choices are the horizontal magnetic and
vertical electric fields, as these most closely approximate the primary signals.
VLF magnetic fields are detected by coils in which currents flow in proportion to the core
permeability, the number turns in the coil and the magnetic field component along its axis. No signal
will be detected if the magnetic field is at right angles to the axis.
A VLF electric field will induce an alternating current in an aerial consisting of a straight
conducting rod or wire. The signal strength is roughly proportional to the amplitude of the electric
field component parallel to the aerial, and to the aerial length.

(b) Magnetic field effects: Eddy currents induced by a VLF magnetic field produce secondary
magnetic fields with the same frequency as the primary, but generally with a different phase. Any

120
vertical magnetic component is by definition anomalous, and most VLF instruments compare vertical
with horizontal magnetic fields either directly or by measuring tilt angles. Eddy currents produce
secondary magnetic fields which oppose the primary field change. Directly above a steeply dipping
sheet-like conductor, the secondary field may be strong, but will be horizontal and will not be detected
by most systems. On either side, there will be detectable vertical fields, in opposite directions,
defining an asymmetrical anomaly.
Steeply dipping contacts also produce VLF anomalies, which are positive or negative depending
upon the sign convention. The classical asymmetrical ‘thin conductor’ anomaly can be looked upon
as produced by two contacts very close together. Two steeply dipping conductors close to each other
produce a resultant anomaly generally similar to the sum of anomalies that would have been
produced by each body individually. However, where one of the bodies is steeply dipping and the
other flat-lying, the results are more difficult to anticipate. Conductive overburden affects, and can
actually reverse, the phase of the secondary field.

(c) Electric field effects: Because the earth is not a perfect conductor, VLF electric vectors near its
surface are tilted, not vertical, and have horizontal components. Above homogeneous ground, the
horizontal field would differ in phase from the primary (vertical) field by 45º, would lie in the direction
of propagation and would be proportional to the square root of the ground resistivity. Over a layered
earth, the magnitude of the horizontal electric field (or tilt of the total field) records the average
(apparent) resistivity, strongly biased towards the resistivity of the ground within about half a skin
depth of the surface.
The phase angle will be greater than 45º in a layered earth if the resistivity increases with depth
and less than 45º if it decreases. Sharp lateral resistivity changes distort this simple picture and very
good (usually artificial) conductors produce secondary fields that invalidate the assumptions on which
the resistivity calculations are based.

(d) Elliptical polarization: Combining horizontal primary and secondary fields with different phases
produces a field which is also horizontal, but which differs from its components in both phase and
magnitude. A secondary field that is vertical and in phase with the primary field produces a resultant
which has the same phase but is tilted and stronger. A vertical secondary field in phase quadrature
with the primary field results in elliptical polarization (section 4.6.4). Figure 4.57 shows the results of
combining alternating vertical and horizontal field vectors. In the in-phase case of example (a), the
vertical vector has its maximum value OA when the horizontal vector has its maximum value OP and
the resultant field vector has its maximum value OT. All three are zero at the same time and at any
other time, the resultant field vector is directed along OT, but with lesser amplitude. In the case of
phase quadrature in example (b), the vertical vector is zero when the horizontal vector has its
maximum value OP, and has its maximum value OA when the horizontal vector is zero. At other
times, represented by OB, OQ and OS, the tip of the resultant lies on an ellipse.
A T A

B B S
S

O Q P O Q P

(a) In-phase (b) Phase quadrature

Figure 4.57 Combination of alternating vertical and horizontal magnetic field vectors

These are special cases. In the general case of an inclined secondary field which is neither in-
phase nor in phase quadrature with the primary field, a tilted, elliptically polarized wave is produced.
Because the secondary field has a horizontal component, the tangent of the tilt angle is not identical

121
to the ratio of the secondary field to the primary field. Because of the tile, the quadrature component
of the vertical secondary field does not define the length of the minor axis of the ellipse. However,
dip-angle data are usually interpreted qualitatively, when such distinctions, which are only significant
for strong anomalies, are ignored. Quantitative interpretations are based on physical or computer
model studies, the results of which can be expressed in terms of any quantities measured in the field.

(e) Coupling: The magnetic component response of a good conductor depends critically on its
orientation. This is also true in conventional EM surveys, but EM traverses are usually laid out at
right angles to the probable geological strike, automatically insuring good coupling. In VLF work, the
traverse direction is almost irrelevant, the critical parameter being the relationship between the strike
of the conductor and the bearing of the transmitting station. A body which strikes towards the
transmitter is well coupled, as the magnetic vector is at right angles to it and current can flow freely
(Figure 4.58). Otherwise the current flow will be restricted, reducing the strength of the secondary
field. If the probable strike of the conducors is either variable or unknown, two transmitters, bearing
roughly at right angles, should be used to produe separate VLF maps.

4.6.10.4 VLF wave in the ground

O wing to the almost infinite conductivity of the ground when compared with air, the VLF wave
arriving at any point is refracted vertically downwards into the ground irrespective of its angle of
incidence (section 1.4). In the wave transmitted into the ground, the horizontal field E lags 45º in
tx
phase behind the vertical field Etz and, moreover, Etx >> Etz. The electric vector Et of the transmitted
field is therefore also elliptically polarized but the major axis of the ellipse is now nearly horizontal. If
its inclination with the horizontal is denoted by λ, the equation for it is again (4.62). The transmitted
magnetic field has only the component Hty in the y direction.
The electric and magnetic field strengths in the transmitted wave can be written as

Etx = (µωρ)1/2 Hty(0)exp(-bz)cos(ωt – bz + π/4) (4.63) ξ

Hty = Hty(0)exp(-bz - cos(ωt-bz) (4.64) ξ

where z is the depth, b = (µω / 2ρ)1/2 and Hty(0) is the amplitude of the magnetic field strength at the
surface. These equations show that the amplitudes of Etx and Hty decrease exponentially and that Etx
is ahead of Hty in phase by π/4 and either field lags in phase behind the corresponding surface field
by the amount bz. Also the ratio of the amplitudes of Etx and Hty is seen from (4.63) and (4.64) to be

|Etx / Hty| = (µωρ)1/2 (4.65) ξ

This relation holds at all depths in homogeneous ground.ii

4.6.10.5 Measurements with VLF station in strike direction

(a) Conductor detection by an H-field anomaly: This case, in which a station is situated so that the
VLF ground wave arrives in the geological strike direction in the area of measurement, is illustrated in
Figure 4.58. It is often referred to as ‘E polarization’ since the E vector lies in the vertical plane
through the strike direction and the H field is transverse to the geological strike and therefore
optimally directed to induce currents in conductors that lie conformably with the geological strata.
The transmitted field Hty induces currents in the sheet conductor shown in the figure but their
secondary magnetic field Hs is not, in general horizontal. Moreover, as the fields Hiy and Hs are not in
phase, their resultant is polarized and describes an ellipse, changing in magnitude cyclically as it
rotates.
At distant points on a profile like the one drawn in Figure 4.58, a solenoid held vertically will show
no signal. In the vicinity of the conductor a signal is induced in it due to the secondary magnetic field.
In most VLF measurements, in practice, the solenoid is now tilted in the vertical plane through the
profile (the yz plane) to make the signal a minimum. In this position the solenoid is perpendicular to

122
the major axis of the ellipse and the residual signal in it is due to the field along the minor axis. The
tilt of the solenoid axis from the vertical is equal to the tilt of the major axis of the ellipse of
polarization from the horizontal. The procedure assumes the ellipse to be in the vertical plane
through the profile which, though not always justified, is very nearly the case for long conductors.
The tilt is above the horizontal on one side of the conductor and below it on the other with a crossover
directly above the conductor where the tilt is zero.

Eiz

Distant transmitter

Direction of wave
propagation Eix Ground surface
x
Hiy
+
j +
Etx
+
A +B
y Hty +
+

Direction of wave
propagation in
the ground

Figure 4.58 Conductor striking parallel to the propagation


direction of the VLF wave (E polarization)

A second solenoid at right angles to the first and with its axis in the same vertical plane will pick
up the field corresponding to the major axis. A comparison of the signals in the two solenoids gives
the ratio b/a of the minor-axis field (b) to the major-axis field (a). This ratio is sometimes referred to in
geophysical VLF literature as the ‘ellipticity’ even though by definition of the ratio, the greatest
ellipticity, 1, is a circle (b = a) and the smallest, 0, is a straight line (b = 0). In other literature it has
been referred to as ‘eccentricity’ which, by mathematical convention, is (a2 – b2)1/2 / a and whose
range is from zero (a circle) to < 1 (since a straight line would be division by zero).
If ∆V is the amplitude of the vertical component of Hs, θ is the tilt and ψ is the phase difference
between Hiy and Hs, it can be shown that tan(2θ) ≈ (2∆Vcosψ) / Hiy. Since usually ∆V << Hiy(0) we
may put tan(2θ) ≈ 2θ and hence θ ≈ (∆Vcosψ) / Hiy(0). Now, ∆Vcosψ is the real (or in-phase)
component of ∆V. Hence, 100 x θ (in radians), which is the quantity read on VLF instruments in
percent units, measures the real component of the vertical secondary field in percent of the primary
VLF magnetic field strength.
The axes ratio b/a can be shown to be approximately (∆Vsinψ) / Hiy(0), that is, the imaginary (or
quadrature) component of the secondary vertical field as a fraction of the primary field strength.
It has been assumed above that the incident field Hiy is exactly at right angles to the strike.
Generally, this is not the case in practice. If α is the angle between the strike and Hty, the effective
inducing field transmitted in the ground is H’ty = Htysinα < Hty. However, the strength of the induced
currents and their secondary field is reduced in the same ratio, namely sinα, and, to a very good first
approximation, neither the magnitude of θ nor the ratio b/a along a profile perpendicular to the strike
is affected in practice.

(b) E-field anomaly and apparent resistivity: The electric field in the ground is effectively only the
horizontal field Etx. Due to this field, an (oscillating) electrical charge density is developed on the
faces of the conductor marked A and B in Figure 4.58 in contact with the host rock. Its secondary
electric field Es is opposite to the incident field Eix at all points along the profile and is especially

123
strong directly above the conductor. If the electric field is measured at different points along the
profile by means of an electric antenna (a wire) laid perpendicular to the profile, it will be minimum
above the conductor. Therefore, the horizontal fields Ex and Hy (but not the vertical field Ez) are
continuous at the ground surface, that is, Eix = Etx and Hiy = Hty at the surface. Hence, the subscripts i
and t can be suppressed for fields measured by a wire antenna laid on the ground and by a solenoid
held near the ground surface with its axis horizontal and denote the fields Ex and Hy.
In practice, modern VLF instruments for prospecting do not measure the Ex field as such but the
ratio Ex / Hy and the phase difference between Ex and Hy. The ratio has the unit of ‘ohm’ and is called
the wave impedance. From equation (4.65), putting µ = µ0

ρ = (1 / 2πµ0ν)|Ex / Hy|2 (4.66) ξ

Thus, the measurement of the amplitude ratio Ex / Hy will yield the resistivity of a homogeneous
ground.
On non-homogeneous ground, an apparent resistivity ρa may be defined by equation (4.66). The
dials of VLF instruments are calibrated to read ρa and the phase difference directly. The latter is, as
equations (4.63) and (4.64) show, 45º on homogeneous ground and differs from this value on non-
homogeneous ground.
If ρa, given by equation (4.66), is measured along a profile across the conductor in Figure 4.58, it
will show a minimum above the conductor because the net Ex is reduced, as explained above.
Thus, when a VLF wave arrives in the strike direction, a geological sheet conductor gives an
electric as well as a magnetic field anomaly. The electric measurements are, however, relatively slow
since the wire antenna has to be laid across the profile and cannot be trailed behind the operator in
moving along the profile. This disadvantage does not exist in the alternative described in the next
section.ii

4.6.10.6 Measurements with station bearing perpendicular to strike

T his case, illustrated in Figure 4.59, is sometimes referred to as ‘H polarization’ since the incident
magnetic field vector lies in the vertical plane through the strike direction. The transmitted
magnetic field will not induce any currents in the plane of the conductor and there will be no
secondary magnetic field along the measurement profile. The transmitted electric field Etx (at the
instant assumed) induces a positive charge density on the conductors face abcd and a negative one
on the efgh face.
Eiz

Distant transmitter

Eix Ground surface


Measurement d
h
line
Hiy +
e a +
+
Etx
+c
Hty
+
+
f b
Direction of wave
propagation in
the ground

Figure 4.59 Conductor striking perpendicular to the


propagation direction of the VLF ground wave (H polarization)

124
The secondary electric field due to these charges is opposite to Eix above the conductor so that
the conductor will be indicated by a minimum in the apparent resistivity calculated from equation
(4.66). In contrast to the earlier case of E polarization, the electrical antenna in the present case is
trailed along the profile of measurement while the solenoid for picking up the magnetic field is
perpendicular to the profile and horizontal. When a VLF wave arrives at right angles to the geological
strike, there is only an electrical but no magnetic anomaly over a sheet conductor.ii

4.6.10.7 Comparison of VLF and conventional EM methods

B ecause of the high frequencies at which VLF systems operate, most conductors appear strong
and many more anomalies are usually located than by EM surveys over the same ground. VLF
methods are best suited to mapping near-vertical contacts and fractures. Conductive mineralization
may be detected, but the magnitudes of anomalies associated with very good conductors may be no
greater than those produced by unmineralized but water-filled fractures which are likely to occupy
larger volumes.
Because of the increasing importance of VLF surveys in hydrogeology, portable transmitters are
now marketed which allow the method to be used in areas where the military signals are weak or
poorly coupled to the dominant conductors.
VLF measurements can be made quickly and conveniently by a single operator. They are
therefore sometimes used to assess the EM characteristics of an area before the expense of a
conventional EM survey is incurred. This is especially useful in populated areas where artificial
sources are to be expected.

4.6.11 Ground penetrating radar

R adar, an acronym for ‘radio detection and ranging’, is a system that observes the reflection of
short-duration electromagnetic pulses, sometimes called ‘chirps’ spanning a range of frequencies
from about 1 MHz to several gigahertz. It was brought to a practical development in Britain by
English and American military scientists for defense against enemy planes during the World War II,
although parallel development was underway in France and Germany. The ultra-high frequency
content of radar waves with their propensity for ‘bouncing’ off almost any material, has led to the
development of many civil as well as military applications such as microwave ovens for cooking,
warning devices for storm system detection in aviation and proximity devices in maritime use for
approaching vessels or icebergs.
Ground penetrating radar (GPR) originated in the use of radio echo-sounding to determine the
thickness of ice, from which it was only a short step to studies of permafrost. It was soon realized
that some penetration was being achieved into unfrozen ground and that the depth of investigation,
although unlikely to ever amount to more than a few tens of meters, could be increased by processing
techniques virtually identical to seismic data. GPR is a relatively new application of radar to
geophysical prospecting and although it has not yet found its way into well logging or seismic
acquisition, it is an interesting method to study as it is correlates electromagnetic wave behavior to
various laws of reflection seismics.

4.6.11.1 GPR fundamentals

(a) Similarity with seismics: A radar pulse emitted by an antenna is partly reflected and partly
transmitted when it meets with an electrical discontinuity, that is, an interface at which there is a
change in electromagnetic wave impedance. If the time for the pulse to go to the reflector in the
ground and return to the receiving antenna (the two-way time) is measured, the position of the
reflector can be determined when the velocity of the pulse is known.
In this and some other respects like the sampling, processing and display of data, GPR has
similarities with reflection seismics, and much of its present ‘software methodology’ of data
acquisition, processing, display and interpretation has been adapted from that originally developed for
reflection seismics. Since the basics of elastic wave theory and reflection seismics have been
covered in Chapters 1 and 2 respectively, only the parts of GPR theory necessary to understand its
behavior will be taken up in this section.

125
(b) Antennas: Most GPR equipment uses dipole antennae for transmission as well as reception.
In one of its simplest forms, an electric dipole antenna consists of a metal tube, a few millimeters to
some 20 mm in diameter, usually of aluminum plated with nickel. The antenna length depends upon
the character of the transmitted pulse, above all its duration, which is known as the pulse width.
Antenna lengths on the order of 0.9 – 1.2 m are required for the efficient transmission of pulses of
widths between about 8 and 12 nanoseconds duration, while narrower pulses of 1 – 2 ns widths are
transmitted by shorter, 0.15 – 0.40 m long, dipole antennae. With optimum design, the antenna
lengths in the pulse-width range of 1 – 100 ns tend to be almost directly proportional to the pulse
width. Thus, the broader the pulse, the longer is the antenna required.

(c) Pulse width and center frequency: In practice, a GPR antenna does not emit a single pulse
but a succession of pulses, each of the same form and duration, at definite intervals, typically from 50
to 2 µs. The inverse of the interval between two successive pulses is the repetition frequency, νr.
Typical repetition frequencies are thus in the range from 20 to 500 kHz. The duration of a single
pulse is usually between about 1 and 100 ns.

1 1

0.5 0.5
Amplitude

Amplitude
0 0

-0.5 -0.5

-1 -1
0 1 2 3 4 5 0 1 2 3 4 5
Pulse duration (ns) Pulse duration (ns)
(a) (b)

1.00
Unit amplitude

0.50

0.00
10 100 1000
Frequency (MHz)
(c)

Figure 4.60 (a) GPR pulse, (b) Idealized asymmetric ‘square’ GPR pulse,
(c) Power spectrum of pulse in (b)

Figure 4.60 (a) shows a commonly used form of a GPR pulse, the width here being 5 ns. Such a
pulse can be shown to be made up of a number of sine and cosine waves of different frequencies,
amplitudes and phases. The square of the amplitude of any frequency is proportional to the power
emitted at that frequency. The frequency around which most of the pulse energy is concentrated is
called the center frequency. The pulse in Figure 4.60 (a) is difficult to analyze algebraically in this
way and therefore we shall take the idealized form shown in Figure 4.60 (b) in which the field rises

126
sharply from zero to a value of one unit in one direction, at which it remains for 2.5 ns, then reverses
sharply to take a value of one unit in the opposite direction for a further 2.5 ns and then drops sharply
to zero.
The function A(f), giving the amplitude at each frequency present in this pulse (the amplitude
spectrum of the pulse) is calculated following a procedure identical to that described in Appendix F for
the pulse in Figure 4.61 below.

A square-wave pulse of duration τ and


τ strength 1 is defined by the equations

1 |t| < τ / 2
f(t) = ½ |t| = τ / 2
0 |t| > τ / 2

It can be shown following the procedure


given in Appendix F that the corresponding
amplitude spectrum is
f = -2/τ -1/τ 0 1/τ 2/τ
A(f) = τ [(sin(πfτ) / (πfτ)] ξ

Figure 4.61 Fourier amplitude spectrum of a where f is the component frequency using
square-wave pulse the notation in Appendix F.

The power at each frequency is where τ is the duration of the pulse. The power is plotted as a
function of frequency in Figure 4.60 (c) for τ = 5 ns from which it can be seen that it is maximum at a
frequency of 150 MHz. This is the center frequency in the present case and it is observed to be
approximately the reciprocal of 5 ns. In general, the center frequency of a radar pulse like that in
Figure 4.60 (a) can be taken to be approximately the reciprocal of the pulse width. It is also possible
to shape a pulse in which this relation is more or less exact.

(d) Time window and sampling: In the interval between two successive pulses, the receiver antenna
measures the electric field in the incoming reflected pulses as an analog signal. This is amplified and
converted in most modern equipment to digital form for storing in the memory of an on-line computer
for subsequent processing. The length of time for which the reflections are recorded between two
transmitted pulses is called the time window. It is selectable in practically all commercially available
GRP systems. Recalling from Chapter 2 that the travel time of a reflection is the two-way time for the
pulse to go to the reflector and return, then a time window τw is equivalent to a maximum sounded
depth of (Vτw) / 2, if V is the average velocity of the pulse over the path.
The interval at which the incoming signals are read (sampled) is the sampling interval. It is
often stated in equipment specifications by its inverse, the sampling frequency. Each reflection
sends back a pulse that, under ideal conditions, has the same form and duration as the transmitted
pulse but different amplitude and, like it, contains a number of different frequencies. If the pulse is to
be faithfully represented by the samples, the criterion for an adequate sampling interval τs, is that it
must be less than 1 / (2νmax) if νmax is the highest frequency in the signal. This means that the
sampling frequency νs = 1/τs must be greater that 2νmax. If τ is the pulse width then, as has been
seen, the maximum frequency in the pulse is about 1/τ which is also approximately the center
frequency of a GPR, so that, as a rule of thumb, the sampling frequency must be greater than twice
the GPR center frequency.
Sampling can be carried out in two different ways. One way, sometimes called ‘flash A/D
conversion’, is to sample the incoming signal at the selected sampling interval for the entire time
window after one transmission pulse. Another (and more commonly used) way, because of its
flexibility, is the sampling-oscilloscope technique, in which only one sample of the electrical signal
received in the time window is taken at a specific instant. After the next transmission pulse, another
sample is taken, but displaced in time by the sampling interval. Thus, after a number of pulses the
whole signal will have been recorded. The number of transmission pulses needed for one scan like

127
this is νsτw, which is the number of samples into which the time window is divided by the sampling
interval, and the time required to assemble the whole signal will be νsτw/νr if νr is the repetition
frequency. To take a numerical example, if νs = 300 MHz, τw = 2 ms and νr = 100 kHz, the time for a
scan will be 6 ms.

(e) Display: In displaying GPR data on paper the scanned, digitally recorded values are converted
back to analog signals and plotted as ‘signal voltage versus two-way time’ with the downward vertical
axis at the measurement point representing the two-way time and the horizontal direction
representing the signal intensity. The plot is referred to as a normal-incidence time section and, as it
applies to seismic reflections, is described in section 2.10.
In one type of display, the intensity is plotted as a wiggle curve with the positive area in each
wiggle filled-in. Typical processing practice is to sample the signal, say, every 2 ns and plot the
running average of five such samples against the two-way time of the central sample of the group.
Successive scans at points along a profile are plotted side-by-side producing a two-dimensional
variable-area display (VAR) that is analogous to a seismic time section such as that in Figure 2.2.
Another display technique is the variable-intensity display in which the field intensity values in
each part of the record covering, say, three consecutive time samples over three consecutive traces
are averaged. The corresponding rectangular picture element in a plot is centered at the central time
sample and the central trace of the group, and then filled-in with a color (or a gray tone)
corresponding to the calculated average intensity.

(f) Power ratios expressed in decibels (dB): Radar systems are frequently described in terms of
processes involving power amplifications and attenuations (gains and losses) which are measured in
decibels (dB). If the input power to a system is P1 and the output power is P2, the gain or loss in
decibels, is equal to ten times the logarithm of the power ratio or 10 log10 (P2 / P1). A 10 dB gain thus
corresponds to a ten-fold, and a 20 dB gain to a hundred-fold increase in signal power. A negative
decibel value indicates a loss or attenuation of the input signal. The logarithmic unit allows the effect
of passing a signal through a number of stages to be obtained by straightforward addition of decibel
gains of each stage.
A logarithm base ten (log10) is the exponent of 10 required to raise (or lower) it to a specific value
(usually a ratio of two numbers). For example, 10 raised to the power of 0.30103 (100.30103) is equal to
2. Thus, log102 is approximately equal to 0.301 and a doubling in power (10 log102) is thus equivalent
to a gain of almost exactly 3 dB. This convenient approximation for doubling is widely used and has
almost become the arbitrary definition of the decibel.
The decibels of power ratios should not be confused with those of amplitude ratios that are
commonly used in seismics. If A1 is the reference amplitude and A2 the amplitude of comparison, the
attenuation or gain is equal to twenty times the log of the ratio of the amplitudes (20 log10 (A2 / A1).
For example, if an input signal of amplitude of 200µV is attenuated to 100µv the attenuation in
decibels is 20 log10 (100 / 200) = 20 log10 (0.5) = - 6.0206. The minus sign indicates a loss or
attenuation of the signal. From this, it is evident why 6 dB is universally considered to be a doubling
of amplitude (or voltage). Appendix I contains tables of common logarithms and decibels.

4.6.11.2 Radar parameters

(a) Q and loss tangents: In considering the theoretical aspects of the propagation of radar pulses in
the ground, it is generally more convenient to consider an electromagnetic wave of a single frequency
ν = ω / 2π instead of a pulse which contains a number of different frequencies. However, the salient
features of the propagation of a GPR pulse are well described if ‘frequency’ is taken to mean the
center frequency of the pulse.
The parameter that determines how efficiently an electromagnetic wave is propagated in a
medium is one referred to in section H.2, namely ρεω, often denoted by Q, a quality indicator. It can
also be written as ωε / σ where σ is the conductivity. The inverse, 1 / Q = (σ / ωε) = tan δ, is known
as the loss tangent. Large loss tangents imply high signal attenuation. Generally, if Q >> 1 (tan δ is
much less than 1), the wave will propagate efficiently without appreciable loss of energy over large
distances. If Q << 1 (tan δ >> 1), the wave is attenuated within a short distance. The ground cannot

128
in general be considered as a homogeneous medium but has different values for Q in different parts
of it and the above considerations will apply to each of these parts.

Table 4.7 Typical values of radar parameters for some common materials

Material ε σ (mS m-1) V (m ns-1) α (dB m-1)


Air 1 0 0.30 0
Ice 3–4 0.01 0.16 0.01
Fresh water 80 0.5 0.033 0.1
Salt water 80 3000 0.01 1000
Dry sand 3–5 0.01 0.15 0.01
Wet sand 20 – 30 0.01 – 1 0.06 0.03 – 0.3
Shales & clays 5 – 20 1 – 1000 0.08 1 – 100
Silts 5 – 30 1 – 100 0.07 1 – 100
Limestones 4–8 0.5 – 2.0 0.12 0.4 – 1
Granite 4–6 0.01 – 1 0.13 0.01 – 1
Salt rock 5–6 0.01 – 1 0.13 0.01 – 1

(b) Permittivity measurements: The attenuation equation of section 4.6.7 also applies to radar
methods, however, the low frequency approximation cannot be used and the parameters permittivity
and conductivity are generally more important to consider. Permittivity (ε) was originally known as the
‘dielectric constant’ and defined in terms of the ratio between the capacities of otherwise identical
parallel plate capacitors with vacuum or material filling the space between the plates. Changes in
units of measurement later made it necessary to assign a value ε0 as the permittivity of empty (free)
space.
In section H.1, dielectric permittivity is defined as the magnitude of the force in newtons between
two massless electrical point charges, Q1 and Q2 coulombs, situated in the medium. If the charges
are separated by a distance r, then the force is Q1Q2 / (4πεr2). The magnitude of permittivity of free
space, ε0, is defined as 107 (m Ω-1 s-1) divided by 4πc2 (m2 s-2), where c is the velocity of light in
vacuum. The dimensions of permittivity are therefore Ω-1s m-1 (or S s m-1, siemen seconds per meter,
or F m-1, farads per meter) and for vacuum ε0 = 8.852 x 10-12 F m-1. However, it is convenient to use
the symbol ε and the old values, renamed relative permittivities, as multipliers instead of assigning an
absolute value to every material. The relative permittivities of rocks and soils are dominated by εwater,
which is about 80. For most other common substances, ε is close to unity.

(c) Permeability: Relative magnetic permeability (µr) also affects radar propagation, but most
materials encountered in GPR surveys have relative permeabilities close to unity and absolute
permeabilities close to the free space value (µ0 = 4π x 10-7).

(d) Velocity of an electromagnetic wave: The velocity of an EM wave in an insulator is equal to the
free space velocity of light divided by the square root of the product of the relative permittivity and the
relative permeability i.e., v = c / (εrµr)1/2. The relative permittivity of most dry and non-conducting
rocks and soils varies within relatively narrow limits of 5 – 10 so that the velocity in the ground is
normally between about 0.05 and 0.15 m ns-1, well below the 0.30 m ns-1 (300,000 km s-1) free space
velocity of light. Table 4.7 lists typical values of radar parameters for some common materials.
Electrical conductivities at radar frequencies differ, sometimes very considerably, from d.c. values
and increase with frequency at roughly log-linear rates (Figure 4.62).
The radar wavelength in any material is equal to the radar wave velocity divided by the frequency,
or c / (ν √ εr) if µr can be taken as unity. The calculations should be straightforward, but because GPR
velocities are usually quoted in m ns-1 and frequencies in MHz, it is easy to lose a few powers of ten
unless orders of magnitude are appreciated. The wavelength of a 100 MHz signal in air is 3 m, in
rock with a velocity of (0.1 m ns-1 it is 10 cm and in salt water (V = 0.01 m ns-1) only 1 cm.

129
100

Conductivity (mS m-1 10


Gabbro
Andesite
Coal

Sandstone

Dolerite

1
10 100 1000
Frequency (MHz)

Figure 4.62 Approximate rates of change in


conductivity with increasing radar frequencies for
some common rock types

4.6.11.3 Reflection and transmission coefficients

W hen, during propagation, an electromagnetic wave meets with an interface between two media
with different wave impedances, it is partly reflected and partly transmitted by mechanisms
analogous to those of a seismic wave when it encounters a velocityxdensity (ρV) boundary as was
discussed in section 2.9. The transmitted energy, as in reflection seismics, may be further reflected
and transmitted by another interface below and so on. The impedance Z (ohms) presented by any
medium to the passage of electromagnetic waves is expressed by the equation

Z = ωµ / κ (4.67) ξ

where κ is given by equation (H.2). In general, Z is a complex quantity because κ is complex.


Let Ei, Er and Et be the amplitudes of the incident, reflected and transmitted electric fields. If the
incidence is normal to the interface so that the electric field is parallel to it, the case now under
consideration, the reflection and transmission coefficients for the amplitude of the electric field, Er/Ei
and Et/Ei, respectively, are given by

Z2 – Z1 2Z2
Rc = and Tc = (4.68) ξ
Z2 + Z1 Z2 + Z1

where the subscripts 1 and 2 refer to the two media on either side of the interface and where the
wave is supposed to be incident from medium 1. These two together satisfy the condition of
continuity of the tangential component of the electric field across the interface, that is, Ei + Er = Et and
1 + Rc = Tc. Normal incidence implies that the same antenna is used for transmission and reception
of the signals. However, if the offset between the antennae is small compared to the distance to the
interface, the incidence may be assumed to be normal and equations (4.68) are adequate.
If we consider only non-magnetic media, that is if µ2 = µ1 = µ0, the reflection coefficient can be

130
expressed using equations (4.67) and (H.2) as

(ωε1 + iσ1)1/2 – (ωε2 + iσ2)1/2


Rc = (4.69) ξ
(ωε1 + iσ1)1/2 + (ωε2 + iσ2)1/2

where ε stands for absolute permittivity.


If both media are non-conducting (σ = 0) and we put ε1 = ε1rε0 and ε2 = ε2rε0, where ε1r and ε2r are
the relative permittivities (section H.1), the reflection coefficient can be expressed as

Rc =
√ ε1r – √ ε2r (4.70) ξ
√ ε1r + √ ε2r
Dry rocks usually have a relative permittivity of about 5 – 7. Moisture increases the permittivity
and if water is present in appreciable amounts in pores and cracks, the permittivity can reach quite
high values. Below the water table a permittivity of about 25 in sedimentary rocks is not uncommon,
but the conductivity is often not affected greatly so that in equation (4.69) we may still put σ2 << ωε2
and neglect it. The reflection coefficient at the groundwater table can be seen from equation (4.70) to
be about – 0.10, which is a fairly high value, and a significant part of the energy of an incident radar
pulse will be reflected from the water table.
The power reflection coefficient, equal to (Rc)2, is sometimes used, but conceals the fact that an
increase in permittivity produces a 180º phase change upon reflection. Radar reflections from
geological materials are determined almost entirely by variations in water content, but the conductive
term dominates in metallic materials.
Reflection power is also governed by the area of the reflector and the nature of its surface. The
strongest responses come from smooth surfaces at which specular reflection occurs. Rough
surfaces scatter incident energy, which reduces reflection amplitudes, and small targets produce
weak reflections. Success with GPR generally requires at least 1% of the incident energy to be
reflected (i.e. Rc > 0.01) and the smallest lateral dimension of the target should be not less than one-
tenth of its depth.
Except when metal objects, unusually high conductive sulphide or graphite ores, waterlogged
soils, saline-water interfaces, etc., are the reflectors, equation (4.70) adequately expresses the
reflection coefficients in the ground. Hence, in many practical applications like soil surveys and
geotechnical work, GPR may be considered to be primarily a tool for mapping interfaces between
layers of different relative permittivities in the subsurface.

4.6.11.4 Application of the GPR method

(a) Monostatic and bistatic arrangements: A GPR system has four basic components: transmitter
and receiver units, and control and display units. The control unit generates pulses that are fed to the
transmitting antenna and also processes the incoming signals for storing on tape. In some
equipment, there is provision for an on-site display as well as for storage of data for playback and
printout in the office. The display and control units in this case are stationary and connected to the
antennae by cables while the antennae are moved from point to point. Very light, fully portable
systems constructed with a view to covering large areas quickly and in variable terrain mostly have
provision only for storage of data.
Although the transmitter and receiver units are separate entities, the two antennae need not be
so physically. The same antenna used for transmission can also be used for reception, an
arrangement called monostatic, in contrast to the bistatic arrangement (Figure 4.63) in which there is
a constant, small offset between the two. In surface GPR, the antennae are nominally horizontal
which in practice means parallel to the ground surface.
The bistatic configuration may be ‘cross-line’ in which the two dipole antennae are parallel and
the line joining their centers is perpendicular to the antennae as in Figure 4.63 (a) or it may be an ‘in-
line’ one as in Figure 4.63 (b). In either case, the profiling direction is along the line joining the

131
antenna centers. In low-frequency bistatic arrangements, the antennae are generally mounted in
separate cases, about 1 m apart, which are fitted with wheels and handle so that they can be towed
along the measurement line. In high-frequency bistatic radars, both antennae are usually mounted,
about 0.30 m apart, in the same case. In other modes, only the transmitter may be stationary and the
receiver mobile, or else both may be moved symmetrically outwards from a fixed point, a procedure
corresponding to the common mid-point (CMP) method in seismics, and used primarily for sounding
large depths, e.g. on glacier ice.

Profiling direction

x x
Tx Rx Tx Rx

Er Ei Er
Ei
h h

(a) (b)
Cross-line antennae: In-line antennae:

Figure 4.63 Cross-line and in-line bistatic GPR configurations for


transmitter (Tx,) and receiver (Rx,)

(b) Profiling and stacking: Most surface investigations with GPR are made as either continuous
profiling or stationary point collection. In continuous profiling, the antennae are pulled along the
profile at what may best be described as walking pace, but sometimes towed behind a vehicle at a
faster speed. If u is the antenna speed, the spacing between scans will be

s = uνsτw / νr ξ

Taking the example at the end of section 4.6.11.1 (d), if the towing speed is 1 m s-1, a scan will be
obtained every 6 mm. However, in practice, a number of scans are stacked to enhance weak
reflections and the lateral separation of the output traces depends upon the scan rate and the number
of scans rather that on the towing speed. At a rate of 128 scans per second, a stack of 32 scans for
example will take 250 ms. During this time, the instrument will have moved 0.25 m which will be the
actual separation of the output traces. The trace obtained will then be a sort of average of the
conditions in the ground over this horizontal distance.
In stationary point collection, the antennae are stationary at the measurement point while the
scans are being stacked. When the operator is satisfied with the data quality, the antennae are
moved to the next station.

(c) Reflections and diffractions: The principal reflections and diffractions are generally more or less
immediately identifiable on a GPR profile as can be seen in Figures 4.64 and 4.65. Here, it should be
noted that the two uppermost signals are not reflections but two direct waves from the transmitter to
the receiver, one in the air and the other in the ground.
Radar waves undergo diffractions from small inhomogeneities and objects comparable to, or
smaller than, the dominant wavelength in the radar pulse as well as from sharp edges. Diffraction
patterns in GPR applications arise, in principle, in the same way as for seismic waves (section 2.12).
It is sufficient to say at this point that on a time section with the monostatic radar, or the bistatic one
with constant antenna offset, diffractions are identifiable as hyperbolae. An example is seen in Figure
4.64 which is a profile across a culvert. The apex of the hyperbola corresponds to the top of the
culvert and on the sides is seen the reflection from the interface between the filling material and the

132
underlying rock.
Assuming a velocity of 1.20 x 108 m s-1, the dominant wavelength in the ground for 10 MHz and
1000 MHz radars will be 12 m and 0.12 m respectively, so that especially in high-frequency GPR
measurements, diffraction patterns in the time section are frequent.

6 8 10 12 14

0 0

10

Depth (m), v = 0.120 m ns-1


1

20
Time (ns)

30
2

40

50 3

Figure 4.64 Diffraction of a radar wave by a buried culvert. Pulse EKKO system, 225 MHz, 64
stacks. Antenna separation, 0.125 m. Horizontal coordinates in meters. (From Parasnis, 1997)

4.6.11.5 Distance determination

I n the monostatic arrangement, the depth of a reflector, or restated, the distance of a reflector from
the transmitter, is obtained simply as vt / 2 where v is the velocity of electromagnetic waves in the
medium and t is the two-way time for the pulse to go to the reflector and return to the antenna. In the
bistatic arrangement with an offset x between the transmitter and receiver (Figure 4.63), the two-way
time is

2 2
h2 + x
t= (4.71a) ξ
v
√ 2

If (4.71a) looks familiar, it is, being exactly the same as equation (2.1a) that gives the two-way time
for a single-layer seismic reflection. From this is obtained

(v2 t2 – x2)1/2
h= (4.71b) ξ
2

where the velocity v of electromagnetic waves in non-conducting and non-magnetic media was given
in sub-section 4.6.11.2 (d), as equal to c / √ εr where µ is taken as unity. It was also stated in this

133
sub-section that the narrow limits of relative permittivities result in ground velocities between about
0.05 to 0.15 m ns-1. In view of the small depths, less than about 20 m, sounded in much of the GPR
surveys, a ‘flat’ value, for example 0.120 m ns-1 (120 m ms-1 or 1.20 x 108 m s-1), is often used for
normal ground in preliminary interpretation. In special situations such as measurements on
waterlogged soils/sands, lakes or glacier ice, this flat value will not be applicable. In such cases, or
when large depths are involved as well as in final work, an attempt should be made to estimate the
relative permittivities of the various materials involved by laboratory measurements on samples or
from calibration measurements in areas of known characteristics.

Tx to Rx separation (m)
5 10 15 20 25 30 35

50

100 Air wave


-1
150 v = 0.3 m ns

200
Two-way travel time, t (ns)

250
300

350 Ground wave


-1
400 v = 0.120 m ns

450

500
550

600

650
700

750

Figure 4.65 Determination of radar pulse velocity in ground. (From Parasnis, 1997)

Very often, the simplest approach to determining the velocity of a radar pulse within a small part
of the ground is to locate some buried reflecting object such as a utility line, conduit in concrete etc.,
whose depth is known and measure the two-way time for the reflection from it. Measurements on
surface outcrops of rock can also be used for the purposes as shown in Figure 4.65. Here, the
distance between transmitter and receiver antennae was varied and the near-surface velocities in air
and ground were calculated from the wiggle move-out of the corresponding direct waves. This pair of
events is seen to be repeated about 50 ns down in the figure. The repetition is not a multiple
reflection, but the result of the processing procedure in which the sampling was carried out every 2 ns
and a running average of 5 – 10 samples was taken. The transmitted pulse, being of 40 ns duration,
the repeated pair is the expression of the complete transmitted waveform. It will be noticed that the
velocity in the ground is 0.120 m ns-1. Several reflections are also seen.
Antennae may be separate or integrated into a single module. Separability is desirable even if
the spacing is to be kept constant because the optimum value depends on the environment as well as
on the frequency and transmitter size.

134
4.6.11.6 Migration

I n the concept of a normal-incidence time section, the two-way time is plotted vertically below the
transmitter position even if the actual ray path is slanted. Thus, in Figure 4.66, the reflection when
the transmitter is in the position T0 will be plotted at A0 (assuming a suitable time scale for the time
axis T0 – A0), but the reflections from points A1 and A2 in transmitter positions T1 and T2 will be plotted
at B1 and B2 on the respective time axes T1 – B1 and T2 – B2.

T0 T1 T2

B2
A2
B1
A1
A0

Figure 4.66 Migration required in GPR measurements

The position of a dipping reflector as it manifests itself on a normal-incidence time section is thus
not the true position. If the reflector has a more complicated form than in Figure 4.66, a raw time-
section display can be misleading. The procedure whereby the recorded reflector position is moved
to the true position is known as migration, the computer software for which is now provided with most
commercially available GPR equipment.
Inherent in the stack-trace method of display, in both GPR and seismic methods, is the concept of
normal incidence, discussed briefly in section 2.10, which, for cases of small dips and almost
horizontal layering, represents fairly faithfully the geological structure. Figure 2.15 shows an example
of how a normal-incidence time section can be quite different from that of the geological structure
when dips are appreciable owing to the fact that the reflector point is not plotted it its true horizontal
position.
An unmigrated time display of an inhomogeneous earth section with curved reflectors contains
both apparent reflections and diffractions. The latter are actually a series of point diffractions that
during the process of normal-incidence stacking form diffraction hyperbolae (Figure 2.17). These in
turn, will, when plotted, give false appearances of reflections. The object of migration is to remedy
this situation and move all the apparent reflector and diffractor events to their correct positions along
the seismic profile, usually leaving the vertical axis in time. A migration program is usually at the end
of the data processing sequence. In addition to the diffraction-hyperbola migration method, there are
many others used in processing centers. Coffeen (1978) and Sheriff and Geldart (1982) provide
excellent discussions of all of the migration methods.

135
5
Well logging

5.1 Introduction

W ell logging is, and has been almost since the beginning, an important and integral part of the
drilling program. Whether the well is successful in producing oil or gas or not, the information
gained from the suite of well logs, leading to a better understanding of the areal geology, is vital to the
promulgation of additional wells in the prospect area. Although the costs incurred by a proposed
logging program are invariably included in the predrilling budget, the actual implementation of the
program depends on several, sometimes unforeseen, factors. The decision to log rests ultimately
with the drilling manager; however, on-site geologists and geophysicists should press to make the
case for the benefits obtained from a logging program. Often, the impact of the monetary losses
accrued in drilling a dry hole can be somewhat diminished by the gain in geophysical knowledge
obtained by logging. Sometimes adverse factors in the drilling operation such as loss of circulation,
stuck tools, bore wall collapse, potential H2S presence may require the drilling manager to cancel part
or all of the logging program in order to save the well or maintain the safety of the drilling personnel.
There are many different types of logs with different objectives that provide a wide array of
results. Which logs to use in a particular case depends on the formation column being drilled, the
scientific needs, the drilling history of the area, the economy of the drilling budget and other factors.
The various logging methods include the use of electrodes, electromagnetic devices, gamma
rays, neutron radiation, sonic waves and mechanical devices for measuring temperature, borehole
diameter, deviation and dip. The methods presented in the subsequent sections are sometimes
referred to as wire-line well logging as their operation requires lowering a specific logging tool into the
borehole suspended by a cable which in turn is connected at the surface to a logging truck. In
addition to wire-line logging, there is mud logging, a continuous program in which the recirculated
drilling mud is examined constantly by the geologists to identify the rock and mineral content of the
drilling cuttings at successive depths, to record changes in formation temperatures, and to identify the
presence of harmful gasses such as H2S. Based upon the identification of the cuttings, a lithologic
column is constructed and any hydrocarbon presence is noted. (See Figure 5.1)
In laying the groundwork for well logging in Chapter 4, the different logs were categorized
according to their respective geophysical method; gamma-ray, density and neutron logs were given
as examples of the radioactivity method; electric, monoelectrode, microlog, focused electric and SP
logs utilize the electrical methods of resistivity and spontaneous potential. The induction log, or
focused induction log, is an example of the electromagnetic method. Additionally, there exists an IP
log that employs the method of induced polarization. Aside from this manner of classifying well logs,
Parasnis (1997) groups the logs, according to their geophysical objective, into three areas namely, (i)
permeability-zone logs, (ii) resistivity logs and (iii) porosity logs. Section 5.2 gives a brief description
of common logging tools. Subsequent sections discuss their methodologies.

5.2 Logging tools

Gamma Ray (GR) The gamma-ray probe (Figure 5.3 (a)) measures the natural gamma
radiation, which originates from the potassium isotope 40K and from the uranium and thorium decay
series (see section 4.4, Table 4.1). This log enables the discrimination between layers of clay and
sand; even the clay content of clayish sediments can be estimated. Barrages by impenetrable
minerals may also be controlled by gamma-ray logs.

136
The probes contain scintillometers and can be operated in dry or cased boreholes. The logging
speed should not exceed 5 meters per minute. Beds of only 0.3-meter thickness may still be
resolved.

Density Log (D) This is also a radiation log, but here variations of the radiation of an artificial
source are registered. This source, 137Cs (Cäsium 137) is placed at the low end of the probe. The
gamma detector is housed above, shielded by a lead column against radiation from the source
(Figure 5.3 (b)).
The gamma radiation that is emanated into the surrounding rock is absorbed by the latter,
according to their density, by the Compton effect. The part of radiation that still reaches the detector
is recorded and is a measure of the rock density.
The source-to-detector distance is conventionally 40 cm. Then, a horizontal expansion of roughly
15 – 20 cm is reached. The density log is applied to differentiate among several dumped materials,
to determine their boundaries, and to locate fractured seepage paths in consolidated rocks.

ground surface SP ES FEL IEL GR D


sand, silty

silt

coarse sand

silt

silt
layers

shale (clay)

sand

Figure 5.1 Comparison of different logs with their relation to core lithology From left to right:
Lithologic column; SP, self-potential survey; ES, electrical survey measures resistivity in 16” and 64”
point array; FEL, focused electrical log for thin layers; IEL, induction electric log measures electric
conductivity; GR, gamma-ray measures natural radiation; D, density log by artificial gamma source
and detector

Neutron Log (N) An artificial neutron source (mostly americum-beryllium) radiates in the
borehole fast neutrons. They collide with atoms of the drilled rock and thereby lose energy. When
they reach a certain level of energy, some are caught by other nuclei. These are excited to higher
energy radiation.
Detectors that are arranged at the probe at some distance from the neutron source measure the
captured gamma radiation and/or the thermic neutrons. Needless to say, the detectors must be
shielded against natural radiation. Neutron logs depend on the hydrogen content. Therefore, water
and hydrocarbons are encountered as pore fluids. After borehole effects of diameter, mud, etc., are
corrected, the rock porosity and permeability can be estimated.

137
Electric Log (EL,ES) This log determines the apparent resistivity of rocks in multiple point
arrays. Most probes allow also the simultaneous registration of the self-potential between a surface
current electrode B and a potential electrode M in the hole. The measured apparent resistivities, also
called mixed resistivities, are a combination of the electric parameters of the borehole mud, the mud
cake, and the flushed and invaded zones of the rock. Figure 5.2 displays these conditions.
Therefore, the apparent resistivity must be converted into true rock resistivity via correction programs
or departure curves.

Vogelsang

Figure 5.2 Distribution of resistivities around a borehole

The most common ES electrode ray is the combination of the 16” normal and the 64” normal
(Figure 5.3 (c)). The shorter has a small penetration and is strongly influenced by mud and its
infiltration into the surrounding rock. However, it resolves even small layers down to 0.5-meter
thickness. The longer spacing reproduces the rock resistivity better, but can resolve only layers > 2
meters.
ES logs should always be run in environmental drilling programs since the measured resistivities
provide knowledge about the sequence of deposits, the depth and salinity of leachates and the clay
contents of basal or top sealings, if perforated by drillholes.

Monoelectrode Log (PR) Other than the multipoint ES log, a single-pointer electrode array may
be used. It is the simple determination of the resistivity between an electrode on the surface and
another at the probe. It has a fair resolution of thin beds but is strongly influenced by the mud and the
invaded zone, and therefore unsuitable to detect the rock resistivity. Wheras ES logs are non-
focused and of simple technology, better resolution and deeper penetration is reached by focused
methods; these, however, require more complicated electronic guidance.

138
(a) Gamma Ray (GR)
Na J crystal stabilized high-voltage supply

photomultiplier signal processor and driver

(b) Density (D)


source (137Cs) lead column Na J crystal electronic section

multiplier

bow spring
(c) Electric (ES)

64”

16”
M2

M1

B
surface

N
A–B measuring current
M1/M2 – N measuring signals

(d) Induced Polarization (IP)


M A1 M* A2 ∅ 63 mm ∅ 70 mm
∅ 25 mm

∅ 61 mm
0.4 m 1.25 m 0.66 m

(e) Salinity/Temperature (Sal/Temp)


temperature sensor electrode carrier

open BN MA B
cross section
lateral slots electronic section

sensing arm
(f) Caliper (CAL)
coil spring and threaded rod with ball-bearing nut

DC motor electronic section

(g) Sonic/Acoustic (SV,BHC)


transmitter acoustic isolator receivers

centralizer centralizer centralizer

microlog pad
(h) Microlog (ML)

electronic section

spring-loaded
arms

back-up pad

Figure 5.3 Probes for geophysical well logging

Microlog, Microlaterolog (ML, MLL) The resistivity distribution in the immediate vicinity of a drill
hole is measured by an array with electrode spacing of 1 – 2 inches. The electrodes are pressed
hard against the borehole wall while moving; the mud is pushed away and the mud corrections can
be omitted.
The small electrodes may be arranged in the unfocused microlog or in the focused microlaterolog
or similar arrays. The purpose of this fine-logging is to resolve very thin layers, to locate single
fractures, joints and fissures, and to determine the resistivity of the mud-flushed volume.

139
Induction Log (IES, IEL) This log, which is also known as the focused induction log, determines
the reciprocal of resistivity: the conductivity 1/Ωm. Its unit is the mho/m or S/m (siemens/meter). A
coil on a probe transmits electromagnetic waves of the frequency 20 kHz. This causes eddy currents
in materials or rocks of different conductivity. They are received by another coil, which is located
approximately 1 meter away. Rock conductivity is calculated from amplitudes and phases of the
received secondary field.
The advantage of IES is that it can work out true rock conductivities of very low resistivitites. It is
suitable for the exploration of rocks that are infiltrated by saline fluids or leachates.

Focused Electric Log, Laterolog (FEL, LL) the electric current of one borehole electrode is
focused to resolve beds, which may be as thin as 0.2 m. The measuring electrode length of 4” is
carried by a probe of the total length of 2 m with the small diameter of about 14”. In spite of this
minute array, its lateral penetration reaches as far as the 64” normal. This is achieved by placing
additional electrodes above and beneath the current electrode.
However, these advantages of FEL are accompanied by drawbacks. Higher resistivities cannot
be recorded and the resolution is reduced by larger hole diameters and by mud invasion.

Self-Potential (SP) Such electric potentials are composed of electrochemical and kinetic
potentials. They may originate from dumped metals, or sulfidic ores, which undergo oxidation or
reduction processes, or from fast-moving gases or liquids. Self-potential can also be caused by the
metal casing.
In principle, it is valid that positive SP hints at increasing salinity of the pore fluid and that
negative SP may indicate fresh water, provided the clay content remains stable within the logged
depth.
In combination with ES, GR and IP logs, estimations of the clay percentage and the permeability
of strata can be attempted. A precondition for a quantitative interpretation is a clearly marked sand-
clay interbedding and different salinities of mud filtrate and pore water.

Salinometer (SAL) The salinometer probe (Figure 5.3 (e)) measures the specific resistivity
(salinity) of the borehole fluid. The electrodes are closely spaced and housed in an insulated metallic
tube, through which the borehole fluid passes. This arrangement is made to prevent the influence of
the resistivity of the drilled rock.
Knowledge of the salt content of the borehole fluid is necessary to calculate the true resistivities
of rock from various electric logs. In gauge wells or boreholes sunk through hard rock, the salinity
can indicate where water inflows or outlets occur. The salinometer probe is often combined with a
temperature tool and is used on the down-hole run.

Temperature (TEMP) The temperature of the borehole fluid is continuously monitored by an


electric resistance thermometer in relation to depth. The accuracy of the measurement may be within
0.01º C. To obtain undisturbed temperatures, one has to wait until the disturbances caused by
drilling have subsided.
Because of the annual temperature variation, the natural increase with depth can first be
observed from 20m downwards. The geothermal gradient normally averages 3º C /100 meters.
Deviations from this can indicate ground-water movements. This is very pronounced at water inflows;
there, the temperature log shows sharp kinks. Especially at hazardous waste dumps, temperature
anomalies in boreholes may indicate exothermal chemical or biological reactions that occur between
the deposited materials near the borehole.

Sonic Velocity (SV) Also known as acoustic or sonic log, the SV continuously records the travel
time of longitudinal waves between two points on the rocks of the borehole wall. The sound
transmitter is housed in the probe (Figure 5.3 (g)), as well as one or more detectors. The measured
travel time is related to the lithology and the porosity.
Sonic logging is difficult and expensive; the raw data have to be corrected for many effects such
as cycle skips, borehole parameter variations, etc. Sonic probes contain a lot of electronic gear and
are prone to disturbing influences. Most of the modern sonic surveys are borehole compensated
(BHC) to increase the reliability of the velocity data obtained.

140
Caliper (CAL) Caliper tools (Figure 5.3 (f)) measure the borehole diameter continuously. The
important result is the deviation from the diameter of the drilling bit, which tells about cavings caused
by loose sediments, frayed material or fractured hard rock. A narrowing of a borehole by swelling of
clay or mud cake is also located.
Customary are caliper tools with three or four arms, which are pressed against the borehole wall
while moving upward. Their spreading is recorded and is linearly connected to the depth. Caliper
logs provide important data for the corrections of other logging methods and can locate cemented
zones, casing and filters.

Flowmeter (FLOW) This probe measures the vertical fluid flow in boreholes or wells. Its main
purpose is to locate the depth at which ground water flows into a well. During a pumping test, the
probe is lowered at a constant speed and continuously records the velocity of vertical flow.
Whenever it passes a water-producing depth, the revolutions will decrease. Thereby, the depth
of the inflow is well marked. Flowmeter data, which are reduced by the cable speed, allow the
computation of the share of the total production each producing layer has. This production can even
be negative, if, for example, water flows out of the hole into a zone of fractured hard rock.
It is necessary to run an additional caliper log in hard rock to aid the interpretation of the
flowmeter results. Preconditions for flowmeter logging are that the velocity of water flow be sufficient.
The vertical flow velocity has to be converted into pumping rates by multiplication with the cross
section of the hole, found by the caliper.

Deviation (DV) The purpose of this tool is to ascertain the deviation of the borehole axis from the
vertical and its azimuth towards north. The dipmeter probe is often a multishot instrument, which
registers the dip and the orientation by taking photographs at every sequence of still measurements.
More advanced are continuously surveying dipmeter probes, which record the spatial geometry of the
borehole axis.
Deviation tools with a magnetic compass are restricted to open holes and do not work in steel
casing. In this case, the expensive gyrocompass has to be employed. The deviation of a borehole
should be measured as often as possible during the drilling operation to allow for early recognition of
an unwanted course of the drill.

Other tools In addition to the above listed logging tools, there are others used by loggers for
borehole and casing inspections. They include video camaras and televiewers to view the borehole
walls to learn about fissures and strata. Also, x-ray tools are used to check casing for cracks.

5.3 Reservoir estimates

U nlike the classic images of the Oklahoma ‘gushers’ in the early 1900s, the rising to the surface of
hydrocarbons in modern wells is greatly subdued by the weight of the drilling mud, limiting or
denying information on the production potential of the well. Although cuttings in the recirculated mud
reveal the general lithology in the well and may perhaps show traces of hydrocarbons, they furnish no
estimates of oil and gas in place. It is through the data of measurements of various kinds in the wells
that such estimates can be made.
If A (m2) and h (m) are the area and thickness of an oil reservoir, then

oil in place = Ah(1-Sw)φ m3 (5.1) ξ

where φ is the porosity of the oil-bearing stratum and Sw is the fraction of total pores that contains
water, so that the fraction 1 – Sw contains oil. An exactly similar equation can be written down for gas
in place but the gas volume exists at the temperature t (ºC) and pressure P (Pa) at the depth in
question. Using Boyle’s law, however, it can be shown that at standard pressure (P0) and
temperature (t0), usually atmospheric pressure and surface temperature,

P (273 + t0)
gas volume = (1 – Sw)φ Ah m3 (5.2) ξ
P0 (273 + t)

141
where t and t0 are in degrees Celsius. If P at the depth d (m) is not known, it may be taken as
10,000d (Pa), this being the hydrostatic pressure assuming the water throughout the geologic column
to be freely communicating.ii

5.4 Permeable zones

W hile porosity is effective in determining the quantity of oil or gas in place, it is not in itself
sufficient to ensure recovery. A high porosity with little or no connection between pores (and
between fractures and other similar spaces) will not lead to any recovery. For recovery, it is also
necessary that the formation be permeable, that is, allow the flow of hydrocarbons under a pressure
gradient.
Permeability is expressed in the oil industry in darcy (or the subunit millidarcy), after H. Darcy
who first studied the flow of fluids through porous media in the 19th century. Physically, it has the
dimensions (length)2 and, in particular, 1 darcy = 0.987 x 10-12 m2. The quantitative aspects of rock
permeability will not concern us since they belong to the province of reservoir engineering. Suffice it
to say that permeabilities are usually such that the recoverable quantity of oil is about 20% of the in
situ estimate, and that for gas, it may be as much as 70%.
From the point of view of estimating hydrocarbon quantities in place, the primary aim of logging is
to locate permeable zones and determine φ and Sw in equations (5.1) and (5.2). The logs easily
yield h for the bed in question but not A. If no other information is available, A can be taken as the
square of the spacing between two wells in which the bed has been observed.
The most important hydrocarbon-bearing rocks are sandstones, limestones and dolomites, all of
which can have high porosities and permeabilities. Sandstone is largely of primary porosity due to
the spaces between sand grains. It is rather uniform and averages about 20%. Limestone and
dolomite possess, in addition, secondary porosity due to fractures and also to cavities formed by
the solution of some rock matrix through the action of water. The porosity of limestone and dolomite
is much less uniform than that of sandstone.
Besides the above three rock types, shales (a category that also includes clays in the present
context) are an important constituent of the sedimentary rocks of oil and gas fields. Shales may have
high porosity but their permeability is practically zero since the clay fraction in them effectively binds
practically all water and renders it immobile. Basically, shales account for the impermeable sections
on a log and non-shales the permeable ones. The latter may contain shaly beds to a lesser or
greater extent, complicating estimations of hydrocarbons in place.ii

5.5 Archies law

U sing the notation common in the logging industry, this law that was discussed in somewhat
greater detail in section 4.4.1.6, can be rewritten as

Rt = Rw / φm Swn (5.3) ξ

where Rt is the true resistivity of a geologic formation in place and Rw that of the water in the pores.
The resistivity of oil or gas is virtually infinite and does not influence Rt. It should then be sufficient to
put m = 2 and n = 2 (see section 4.4.1.6). Equation (5.3) can be rewritten then as

Sw = (Rw / Rt)1/2 φ-1 (5.4) ξ

This is sometimes stated with the right-hand side multiplied by an empirical ’correction factor’
depending on rock type. The factors are usually close to unity, their values and the practice of using
them varying from one logging company to another. Sw can be calculated from (5.4) if Rt, Rw and φ
are known. It is fair to say that much of the logging industry is based on equation (5.4).
As mentioned above in section 5.1, the logs are basically categorized, according to Parasnis, into
three groups according to their geophysical objectives i.e., permeability zone, resistivity and porosity.
The following sections will briefly discuss the common logs in use today as they apply to each of
these three categories. Tables C.1 and C.2 list the logs according to type and objective respectively.

142
5.6 Permeability-zone logs
SP Gamma ray
mV
5.6.1 Self-potential
20 0 API units 100
- +

T he self-potential (SP) log is obtained by lowering an 5500’


electrode (usually a lead piece) into the borehole
filled with conductive mud and registering the electric
potential at different points in the hole with respect to a
fixed point at the surface. The recording is continuous,
the logging rate being about 1500 m per hour. A typical
SP log is shown in Figure 5.4. The SP is constant as
long as the electrode is opposite one and the same
formation but jumps more or less suddenly to another

Shale line
level on crossing the boundary between two formations.
The maximum SP gradient is usually observed almost
exactly opposite the boundary. This enables h in
equation (5.1) or (5.2) to be estimated fairly accurately
in most cases. 5600’
The SP against permeable zones is strongly
negative in general with respect to that opposite shales.
The observation can be explained by a system of two
principal electrochemical potentials as shown in Figure
5.5, where Esh is the e.m.f. across the shale/nonshale
interface and Emw that across the interface between the
water in the permeable zone and the mud filtrate (dotted
area in the figure) that has invaded the zone. There
may also be electrokinetic potentials at the well wall but
these can be neglected if the flow of water across the
wall is small. Assuming that the current in the circuit
ABCD is vanishingly small, the application of Kirchoff’s
law to the circuit yields the result
5700’
Emw + Esh + UB – UA = 0 (5.5) ξ
Figure 5.4 SP and γ-ray logs

if UA and UB are the potentials in the borehole mud at points just on either side of the shale/non-shale
interface. Equation (5.5) leads to the SP jump:

SP = UB – UA = - (Emw + Esh) (5.6) ξ


- 0 +

A
Shale D
UA Esh

UB B C
Non-shale

Emw

Figure 5.5 Simplified electrical circuit to explain SP jump across shale/non-shale boundary

143
The geologic formations and the mud contain many different species of ions but it is now
assumed that the e.m.f., Emw + Esh, arises essentially due to an equivalent concentration cell as in
Figure 4.20 with only Na+ and Cl- ions participating and with C2 and C1 representing respectively the
ionic concentrations in the mud filtrate and formation water of the permeable zone. Equation (4.24) in
section 4.4.2.3 (b) can then be applied. However, as the resistivities of (dilute) electrolytes are
inversely proportional to ionic concentrations, the ratio C1/C2 can be replaced by Rmf/Rw where Rmf is
the resistivity of the mud filtrate.
Inserting the experimentally determined values of the ionic velocities in equation (4.24) and
changing the base of logarithms to 10, equation (5.6) can be written as

SP = - (65 + 0.24t)log10(Rmf/Rw) mV (5.7) ξ

if t is the temperature (ºC) at the observation point.


In practice, therefore, a shale line representing UA is identified on the log and the SP jump
measured with respect to it. Since Rmf can be measured directly on a sample of the mud filtrate,
equation (5.7) can be solved for Rw.
It will be realized that a number of assumptions, some of them already stated above, are implicit
in this method so that the water resistivities obtained are called equivalent resistivities Rwe. To obtain
the true values Rw, it is necessary to use empirical correction charts.ii For a more detailed discussion
of self-potential (SP), refer to section 4.4.2.

5.6.2 Gamma-ray logging

T his technique utilizes the natural radioactivity of rocks and is used chiefly for correlating
sedimentary strata in petroleum prospecting. The apparatus consists, in principle, of a γ-ray
detector, usually a photoelectric scintillometer that ‘counts’ scintillations produced by gamma-ray
photons releasing energy in the form of light, and its preamplifier suspended in a borehole by means
of a waterproof electrical cable. The counts are quantified to the energy produced by photons
measured in electron volts (eV). This relationship of scintillation counts to energy is seen in the
gamma-ray spectrum in Figure 4.3. The output of the detector is further amplified on the surface and
recorded continuously as the detector is lowered (or raised) in the hole. The logging rate averages
about 33 m h-1.
Generally speaking, shales and shaly sandstones show a very high radioactivity (the highest in
sedimentary rocks) while salt, coal, anhydrite, limestone, quartz sands, etc., are weakly radioactive.
Thus, in a γ-ray log, the peaks in the intensity will correspond, in general, to shales while the lows will
indicate the presence of limestones, salt, etc., and other non-shale formations (see Figure 5.4).
Owing to the statistical nature of radioactive decay (section 4.3 and Figure 4.1), the γ-ray log is
not quantitatively reproducible in an exact way, in contrast to SP which reproduces itself remarkably
well in successive runs. The standard gamma ray log is not put to any quantitative use in log
interpretation but in a semi-quantitative way it is an important indicator of the shaliness of a possible
non-shale formation.ii Section 4.3 covers in greater detail radiation in general and specifically the
radioactivity of rocks.

5.7 Resistivity and conductivity logs

T he object of these logs is to estimate the true formation resistivity Rt of permeable zones for
insertion in equation (5.4). A major problem in estimating Rt is that the mud filtrate invades the
permeable zone up to a longer or shorter distance form the well bore and the apparent resistivity or
conductivity obtained with any device will be influenced to a greater or smaller extent by the ‘invasion
profile’ and the mud resistivity. Moreover, the resistivity of all strata above and below the zone in
question will also affect the measurements.
Analogously to the electrical soundings in Chapter 4, an attempt is made in logging to determine
Rt by means of ‘deep’ logs, e.g. by varying electrode separations. However, a deep log is necessarily
more influenced, owing to its large electrode separation, by beds adjacent to the one being
investigated so that the apparent resistivity obtained usually needs large and uncertain corrections
before Rt can be estimated.ii

144
5.7.1 Resistivity logs

I n one of the earliest resistivity log devices (‘normal array’), one current electrode A and one
potential electrode M are lowered into the hole while B and N are kept at the surface. The standard
short normal log uses a spacing AM = 16 inches, the long or deep one a spacing of 64 inches. The
apparent resistivity is calculated in the appropriate manner using equation (4.14) in section 4.4.1.7.
Although a number of theoretical, empirical and semi-empirical corrections to the measured apparent
resistivities to obtain Rt have been proposed over the years, the present consensus of logging experts
seems to be that neither the short normal nor the long normal array is useful for estimating Rt with
any degree of reliability (Dewan, 1983), although the short normal, in particular, is of considerable
help in geologic correlation.
Another early device is the lateral array in which, besides A, both M and N are in the borehole at
a relatively short distance from each other. The arrangement is basically a pole-dipole array (section
4.4.1.8 (d)). While efficient in locating interfaces, the array is no better for obtaining Rt than the
normal arrays.

Uninvaded zone

Invaded zone

Figure 5.6 Permeable bed invaded uniformly

The present tendency in resistivity logging is to use the ‘Laterolog’ device in which a system of
guard electrodes besides the main A and B electrodes is intended to prevent the spreading of the
current over the ground at large and to confine it instead between two horizontal planes at a distance
roughly equal to the length of the electrode spread (Jackson, 1981). The concept works in
homogeneous media (in which case it is not needed) but in inhomogeneous media like a stratified
earth, a ‘focusing’ of the current cannot be achieved without a prior knowledge of the resistivity
distribution in the ground. Notwithstanding this fundamental objection, the Laterolog is now used
routinely.
An empirical method of estimating Rt from the Laterolog apparent resistivity Ra, provided the
permeable bed is thick and invaded uniformly up to a constant distance (Figure 5.6), is as follows.
The invaded and uninvaded zones are assumed to contribute to Ra in proportion to their
resistivities, Ri and Rt respectively, and in proportion to their ‘weights’. If w is the weight of the
invaded zone, that of the deep zone is 1 – w, so that

Ra = Riw + Rt(1-w) (5.8) ξ

The weight w depends on the diameter of the invaded zone, on the ratio Ri/Rt and on the array
parameters. The problem of estimating the weights, or the pseudo-weights as they might better be
called, for the situation shown in Figure 5.6 is somewhat similar to finding the effect of the superficial
layer on ρa in electric sounding on a two-layer earth (Chapter 4). An estimate of Rt can be obtained
from (5.8) if Ri and w are given reasonable values. There are also various other empirical methods
for estimating Rt form Ra (Pirson, 1963).

145
5.7.2 Induction logs

T he induction log instrument is basically the


electromagnetic coaxial coil system (see Figure
4.49) with the axis along the well axis. Frequencies
Resistivity
Ohm m
Conductivity
-1
Siemen m
Induction
16“ Normal conductivity
used are in the range 20 – 30 kHz. In earlier 0 200 1000
40“ Spacing
0
instruments, only one transmitting and one receiving
coil were used. More recent constructions use an
array of auxiliary coils spaced above, below or
between the main coils (spacing c. 1 m) to focus the
primary field. Much the same objection can be raised
to the concept as in the case of the Laterolog
instrument.
Inhomogeneous media, the secondary voltage in
the receiver coil is approximately proportional to the
conductivity. In inhomogeneous media, the readings
give an apparent conductivity. An empirical formula
such as equation (5.8), with the resistivities replaced
by conductivities, is used to estimate the true
conductivity Ct = 1/Rt of the permeable zone. For other

100 ft
empirical methods of determining Ct, reference should
be made to the specialized literature already cited and
to Lynch (1962).
In contrast to the other logs discussed so far, the
induction log can be run in empty holes or in holes
filled with non-conductive mud. Logging speeds for
resistivity and conductivity logs are on the order of
1500 – 2000 m per hour. The resistivity logs are
generally preferred in relatively high resistivity
formations (Rt > 200 Ωm) and the induction logs in low-
resistivity formations (Rt < 200 Ωm). An example of
resistivity and conductivity logs is shown in Figure 5.7.
Note that the horizontal scale for resistivity is
ascending left to right, while for conductivity it is
descending in the same direction. This is done for
ease of direct comparison. Figure 5.7 Resistivity and induction log

5.8 Porosity logs

P orosity (φ) is the last remaining parameter needed to calculate Sw from equation (5.4). For this,
three logging methods are in common use to measure φ, namely, density, neutron and sonic logs.

5.8.1 Density log

I f ρma is the matrix density of a ‘clean’ formation, that is, a non-shale permeable section, devoid of
any shaly beds, ρf the density of the fluid filling the pores and ρb the bulk density then

ρb = φρf + (1 – φ)ρma ξ

from which

φ = (ρma – ρb) / (ρma – ρf) (5.9) ξ

The bulk density ρb is measured by lowering a γ-ray source into the hole and counting the number of
high-energy γ rays arriving at a fixed distance when the source radiation is scattered by collisions with

146
the electrons in the rock material. This type of scattering is called Compton scattering (section 4.3.7).
The method is essentially an adaptation to borehole work of the procedure described in section
4.3.13, although the method was initially developed for well logging.
In modern density logging, there are two detectors; one at a distance of about 1 m from the
source, the other nearer, and the sonde on which these and the source are mounted is pressed
against the borehole wall. Owing to its short distance from the source, the near detector gives
greater weight to any mud cake on the wall so that corrections can be estimated for the scattering in
the cake. The dual-detector density log is therefore called a compensated density log (CDL).
The depth of investigation of the CDL is about 10 cm from the borehole wall, which means that
the log senses mud filtrate in the invaded zone, the density of which (ρf in equation (5.9)) is very
nearly 1000 kg m-3. Standard values used for ρma are typically 2650 kg m-3 for sandstones and
sands, 2680 kg m-3 for limey sands or sandy limes, 2710 kg m-3 for limestones, and 2870 kg m-3 for
dolomites.
A further development of the CDL is the litho-density tool (LDT) based on registering very low-
energy gamma rays within a narrow band in addition to the high-energy rays in the CDL. The
absorption of low-energy γ rays is a photoelectric phenomenon and not a loss of energy due to
collisions with electrons as in the Compton scattering of high-energy rays. The long-spacing detector
is calibrated to measure Pe, the photoelectric absorption coefficient per electron present in the
medium through which the rays have passed. Pe is strongly dependent on the average atomic
number of the formation and is therefore a good indicator of the type of rock matrix, making possible
a more accurate estimate of ρma in equation (5.9) than would otherwise be possible in cases of
complex lithology.ii

5.8.2 Neutron log

T here are two borehole methods known as neutron-gamma and neutron-neutron logging which
employ neutrons. A convenient source of neutrons often used is a mixture of radium and
powdered beryllium. The beryllium is bombarded by α particles from the radium and fast neutrons
are produced according to the reaction
9
4 Be + 42He → 12
6 C + 10n + energy

In collisions with nuclei, neutrons are gradually slowed down until they reach thermal velocities.
Hydrogen nuclei are particularly efficient in producing such ‘thermal’ neutrons and then capturing
them according to the reaction
1
0 n + 11H → 12 H + γ

with the production of γ radiation.


Hydrogen nuclei are present in rocks containing oil, water, natural gas, etc., and the γ radiation to
which they give rise can be detected by lowering a neutron source just ahead of a γ-ray counter.
In the neutron-neutron method, the intensity of the neutrons scattered by the hydrogen nuclei,
rather than the intensity of the γ radiation due to their capture, is detected. Neutrons do not produce
appreciable ionization so that a special device is needed for their detection. One such device is a
Geiger tube filled with boron triflouride gas. The neutrons react with the boron transforming it into
lithium and releasing an α particle, which in turn ionizes the gas and reveals the presence of
neutrons.
In passing through matter with high hydrogen content, the neutrons are slowed down and
captured by the hydrogen nuclei at a very small distance from the source. On the other hand, if the
hydrogen content is low, they travel a relatively large distance before reaching thermal velocities.
Since virtually all hydrogen in rocks is in the pore fluids, the tool can be calibrated to record the
porosity directly by measuring its response in reference models of known porosity in which the pores
are filled with water. The matrix in the reference model also has some effect on neutron velocities.
Most neutron tools have dual calibrations made by using two different reference models having
limestone or sandstone respectively as the matrix.

147
The logging engineer has the option on the well site to record the porosity using one of the two
calibrations depending on his or her judgement of the local geology. If a certain rock formation in the
well does not conform to the selected option, only an ‘apparent limestone (or sandstone) porosity’ will
be obtained. This can be converted to the true porosity by means of experimentally obtained charts
such as that in Figure 5.8, which is valid for readings, recorded using the option ‘limestone matrix’.
Thus, for example, if the recorded porosity value is 14% at some depth where the rock happens to be
a dolomite, the true porosity will be 11%.

35

30

25
Apparent limestone porosity – (%)

20

15

10

0
0 5 10 15 20 25 30 35
Corrected porosity – (%)

Figure 5.8 Example of chart for matrix correction. After Dresser Atlas (1975)

The neutron-gamma and the single-detector neutron-neutron logs have now been largely superseded
by the compensated neutron log (CNL) in which there are two detectors about 25 and 40 cm
respectively from the neutron source and the ratio Nn/Nf of the neutron intensities in the near and far
detectors is measured. The ratio increases with porosity and as with the older instruments the
porosity is recorded directly. The depth sensed is about 30 cm.
The neutron log is generally run together with the density log (and the natural γ log) and seldom
by itself. Normally, the neutron log indicates a somewhat higher porosity than the density log since
the neutron tool also senses the bound water in the clay particles that are nearly present. In addition,
dolomitization, which, as it makes the rock denser, leads to higher neutron absorption than in pure
limestone or sandstone. Empirically, the true porosity is found to be very nearly the mean of the two
values.
However, if the pore space in a formation is filled with gas, the bulk density as well as the
hydrogen content will decrease and the neutron log will indicate a lower porosity than the density log.
This so-called ‘crossover’ in a combined record of neutron and density logs is therefore a clear

148
indicator of gas (Figure 5.9). False crossovers can, however, be obtained if the porosity is recorded
assuming a limestone matrix where the actual matrix is sandstone, while crossovers may be
suppressed if the matrix in this case is dolomite instead. The logging rate for neutron and density
logs is about 550 m per hour, that is, the same as for the natural γ-ray log. In fact, all three tools are
usually mounted in tandem on the same sonde, about eight meters long.

Porosity index (%) LS matrix


Compensated formation density porosity
30 20 10 0 -10

Compensated neutron porosity


30 20 10 0 -10

3500´

3600´

3700´

Figure 5.9 Density and neutron logs with crossover

5.8.3 Sonic or acoustic log

5.8.3.1 Porosity from sonic travel times

T he knowledge of sonic travel times and their derived sonic velocities is important for the
recognition of seismically active horizons, for the calibration of seismic refraction and reflection
surveys, and for estimating the porosity and the amount of fracturing in a formation.
The sonic log is based on the observation that the velocity of sound in a porous rock depends
upon the velocity in the matrix as well as that in the fluid occupying the pore space. The effect is
extremely complex in reality and no simple theoretical description of it can actually be given even for

149
highly idealized systems (Tolstoy, 1973). However, as an empirical aid in estimating the porosity φ, it
is assumed that the travel time through a block of porous rock is the same as if the matrix material
were all pressed as a uniform medium at one end and all the fluid filled the pores into a second
contiguous medium.
Considering a block of unit thickness, the thickness of the fluid and the matrix parts will be φ and
1 – φ. If tf and tma are the travel times for a wave to go through unit thickness of fluid and matrix
respectively, the so-called specific times, the two times in the block will be φtf and (1 – φ)tma so that
the total travel time t for a wave to go across the block of unit thickness will be given by the following
formula (the Wylie relation):

t = φtf + (1 – φ)tma (5.10) ξ

This relationship then gives porosity as

φ = (t – tma) / (tf – tma) (5.11) ξ

Sonic logging, or continuous velocity logging (CVL), measures travel times in rocks adjacent to
the borehole, over intervals of about one foot. Contrary to the implication of the name ‘continuous
velocity logging’, the sonic log does not measure the velocity as such but the time taken by the
compressional wave to go through the distance (typically about 0.6 – 0.7 m) between two receivers,
the near one of which is about 1 m from the source.
Sonic logs are used for several geological and
geophysical applications, including the
determination of interval velocities over one-foot Upper
intervals throughout the depth range of the log. transmitter
These values are approximately equal to
instantaneous velocities. Modern versions use
a system of two sound transmitters and four
receivers between them to compensate for R1
borehole effects (Figure 5.10). Logging speeds R2
are about 1500 – 1600 meters per hour.
R3
The observed time is converted to a specific
time (s m-1), that is, the time needed for the wave R4
to go 1 m. The tf in equation (5.11) is assumed to
be that for the mud filtrate (typically 620 µs m-1)
whereas tma depends on the rock type concerned,
Lower
which must be known if accurate φ values are to transmitter
be obtained. Standard recommended values for
tma are 144, 161 and 177 µs m-1 for dolomite,
limestone and sandstone respectively. Figure 5.10 Sonic log sonde, 2 Tx, 4 Rx

5.8.3.2 Synthetic seismograms and low-frequency models

T he sonic log in conjunction with the density log is instrumental in providing data necessary in the
processing of seismic data. The reflection coefficient, as discussed in section 2.9, is a measure
of geologic changes at velocity boundaries and is derived from the acoustic impedances of the two
layers. Acoustic impedance is the product of a medium’s density ρ and its phase velocity V.
The instantaneous velocities obtained from an ideal sonic log, considered as ‘true velocities’, are
used in the calibration of interval velocities, which are needed for RMS or stacking velocity
corrections.
Additionally, the sonic and density data can be combined to form a synthetic seismogram or
simply synthetic. Since the well-logged data are not subjected to adverse effects such as multiples,
noise or signal attenuation, a series of reflections can be synthesized to give a more accurate
representation of a vertical time section. The synthesized trace is repeated several times and filtered
to resemble the actual seismic data. The synthetic program can also insert multiples to help predict

150
the location of multiples in the seismic section. The correlation of a synthetic to a seismic section is,
of course, only as good as the latter’s proximity to the actual well location.
In Figure 5.11, the well was drilled on the seismic line at SP 495. The resulting synthetic
therefore correlates fairly well especially in the “C” section from the top of C-4 to the base of C-9. The
apparent seismic event at 1.2 sec on the seismic section may possibly be a multiple, as it does not
appear on the synthetic. Conversely, the apparent geologic events between 0.7 to 0.9 sec on the
synthetic do not correlate to the seismic section indicating possible signal attenuation due to noise,
improper statics solution or poor velocity control.
Another geophysical use of the sonic log is in the creation of a low-frequency model for inversion
processing. Normal seismic data is typically filtered on the low end from between 0 to 8 Hz to 0 to 15
Hz to suppress low frequency noise associated with instrument limitations and ambient noise during
acquisition. When time-domain amplitude data is transformed into frequency-phase, an assumption
is made that the input data contains the full spectrum of frequencies from 0 to Nyquist. In order to
replace the low-frequency component, a low-frequency model is constructed from the lithologic
column of the sonic log.
Figure 5.12 represents a lithologic
column in which the horizontal
extreme is given by the velocity of the
layer medium. The curve at the right
is the low-frequency representation of
the lithologic model. If the column
were converted to time and covered
one second, the low-frequency curve
would contain frequencies from 0 to
about 10 Hz with a dominant
frequency of 4 Hz.

VELOCITY
5000 10000 15000 20000 fps

T/ C-4 T/ C-4
Possible
multiple
T/ C-6 T/ C-6
DEPTH

B/ C-9 B/ C-9

(b)

Lithologic column Low-frequency (a)


velocity model
Figure 5.11 (a) Actual seismic section;
Figure 5.12 Low-frequency model (b) Synthetic seismogram

5.8.4 Electromagnetic propagation travel-time log

T he EPT log, as this is called, measures primarily the specific travel time of very-high-frequency
electromagnetic waves (50 MHz – 1 GHz) in the zone invaded by the mud filtrate. The velocity,
the specific travel time and the penetration of such waves depend only on the dielectric constant
(sections 4.6.7 and 4.6.11) and not on the frequency. By an assumption similar to that underlying
equation (5.11), the porosity can be calculated by the same algebraic formula, using the times for EM
propagation instead.

151
The dielectric constant of a porous medium depends strongly on the water content.
Consequently, the EPT-derived porosity φEP indicates essentially that part of the pore space which is
water-filled whereas the porosity φ, obtained from a neutron/density log combination, provides a
measure of all the pore space that is fluid-filled. If the pore space contains water as the only fluid,
then φEP = φ. If not, the difference between φEP and φ can be attributed to the presence of oil. For
more on the EPT log, refer to Dewan (1983).ii

5.9 Auxiliary logs and measurements

B esides the above-described primary logs that measure physical properties of the formations, a
few other types of measurements are also routinely carried out in well logging. These are
needed principally for applying various corrections to the data acquired by the primary logs. The tools
used for these auxiliary functions are described in section 5.2 and shown in Figure 5.3.
An estimate of the temperature is required to correct for the dependence of resistivity
(conductivity) on it. Usually it is only measured at a few points in the well (often only at the top and
bottom) and the temperature at intermediate points is estimated by simple linear interpolation.
The well diameter and its variations are recorded by the continuous caliper log. This information
is needed to estimate the effect of borehole mud on resistivity measurements and the effect of
rugosity of the hole on density and neutron logs. In addition, for certain logging methods that require
the logging tool to make good coupling of the sonde to the borehole wall, such as the microlog tool or
the SAT tool for vertical seismic profiling, the caliper can indicate e.g. washout zones that may
negatively affect the coupling. Once these zones are located, the logging engineer can better
interpret the results of the micrologging or in the case of VSP surveys, relocate the tool up or down a
few meters, to encounter better coupling.
Another auxiliary log that is included in many logging programs is the dipmeter log that records
the dip of a formation. This is required for determining where in a geologic structure a well is located,
the kind of faulting encountered, etc. It is not primarily a log for calculating hydrocarbon in place.
Not included in the category of logging methods but nonetheless important to the well-site
geologists is the extraction of core samples in the borehole. A long open-ended sleeve with a cutting
bit is lowered into the hole ahead of the drill pipe and cuts its way into virgin formation, not unlike the
action of a round cookie cutter. Once retrieved and the sleeve opened, the cores reveal rock type,
hydrocarbon content and provide information about permeability and porosity through microscopic
examination of the pore spaces. These are run at nonspecific intervals and are dependant upon the
intuition of the geologists and the discretion of the drilling manager.

5.10 Basic log interpretation procedure

T he physical background of well-log interpretation should be evident from the previous pages but it
will be convenient to recall and summarize the principal steps. Even if logs may be recorded
digitally, the final displays are analog ones on paper or PC screens with two or more logs alongside
each other. Much of the modern processing of logging data is either almost fully automated or done
partly interactively on PC displays. However, the basic steps involved are directly or implicitly as
follows:

1) Scan the SP and/or gamma-ray log and pick out non-shales, that is, permeable zones (high
negative SP, low γ-ray count) after determining the shale line.
2) Estimate Rw from the SP jump using equation (5.7) and read bed thickness h from the log,
which is the distance between the inflection points of the relevant indication.
3) Scan resistivity (conductivity) curves looking for high apparent resistivities (low
conductivities). These may be due to the presence of hydrocarbons.
4) Estimate Rt from equation (5.8) or its counterpart if conductivity is scanned.
5) Read porosities φ in the zones of interest. Locate any ‘crossovers’ in density and neutron
logs to identify gas occurrence.
6) Calculate Sw in the zone of interest from equation (5.4).
7) Finally, calculate hydrocarbons in place from equations (5.1) and (5.2) assuming a
reasonable estimate of area (A).

152
The preceding steps show only the fundamental logic involved. A considerable amount of
detailed log interpretation is based on empirical and semi-empirical observations requiring special
charts, correction tables, etc. For a description of these, reference should be made to specialized
logging literature, bearing in mind that many of the correction procedures vary in detail from one
logging company to another.

5.11 Other geophysical methods

A s the requirements of modern oil and mining exploration, hydrological investigations, civil
engineering and environmental monitoring have grown, a number of special techniques suited to
particular problems have been proposed and applied with varying degrees of success. The purpose
of this section is to mention a few of these techniques and discuss some topics that are of common
interest in all geophysical surveys.

5.11.1 Borehole magnetometer

T he measurement of self-potential, electrical resistivity and elastic wave velocities in boreholes


has already been mentioned (sections 5.6.1, 5.7.1 and 5.8.3 respectively). It may be added that
magnetic intensities too can be measured in a borehole by lowering a fluxgate or a proton
magnetometer into it.
The borehole magnetometer has been used chiefly in iron-ore prospecting as an auxiliary
instrument. If, for instance, a borehole has failed to encounter an expected ore zone, magnetic
measurements can frequently reveal whether the zone is present near the hole and also indicate its
distance. Weak and erratic mineralization often makes it difficult or impossible to establish the
precise limits of a magnetic impregnation zone by an inspection of the drill cores alone. In such
cases, the borehole magnetometer may be of considerably assistance.
The instrument can also be employed sometimes for correlating zones of the same grade in
different parts of a magnetite ore deposit.ii

5.11.2 Charged-body potential method

T his technique, also called mise-à-la-masse method, involves the placing of one current electrode
in a conducting body (e.g. an ore mass) with the other electrode at a distant point. The electric
potentials are measured at points on the ground surface or in boreholes when a current is passed.
Care must be taken before passing the current to compensate for any SP between the potential
electrodes. The measured values are divided by the current passed through the electrode and
expressed as mV A-1, a normalization procedure that avoids the need to keep the current strictly
constant. On homogeneous ground, the equipotentials on the surface will evidently be circles with
the epicenter of the electrode as the center. The form of the equipotential surfaces in three
dimensions may be visualized by imagining Figure 4.8 to be rotated around a vertical axis through the
buried point electrode. In general, however, the equipotentials will be distorted in practice from their
pattern in homogeneous ground due to electrically conducting inhomogeneities.
The charged-body potential method is particularly useful in finding out whether an ore outcrop (or
an ore width encountered in a borehole) is a small mass or part of a much larger mass. In the former
case, the equipotentials on the surface will be more or less circular, while their pattern in the latter
case will reflect the strike and geometry of the larger mass. The technique is also useful for testing
whether (conducting) ore widths encountered in different holes are interconnected or isolated from
each other (Parasnis, 1967). Several examples of the application of this technique in ore prospecting
have been given by Ketola (1972) and it has also been used to study the geometry of electrically well-
conducting fracture zones in a highly resistive environment (Jämtlid et al., 1984).ii

5.11.3 Logging in crystalline rocks and coal fields

Geophysical measurements in boreholes in crystalline rocks have become of great importance in


recent years with the need for finding safe repositories for used nuclear fuel (Parasnis and

153
Soonawala, 1984). All techniques of well logging described in the previous sections can be employed
in crystalline rocks but the object is to delineate fracture-free zones of fresh rock. Fractured rock is
indicated generally by relatively low values of apparent resistivity, density and seismic velocity. Of
particular use in this connection, because it can indicate even minor fractures, is the single-point
resistance log that simply measures the contact resistance between a short brass rod and the
borehole wall. Fractures are also indicated by characteristic irregularities in the temperature in a hole
provided the hole is water-filled.
Logging in coal fields is mainly used for geologic correlation and to estimate the shaliness of a
coal formation (Kayal, 1979). Shales and non-shales are distinguished in this case by the same
criteria as in oil fields (section 5.6).ii

5.11.4 Geothermal methods

O wing to the radioactivity of the crustal rocks, heat is being continuously transported to the surface
-2
of the earth from its interior at a mean rate of about 50 mW m . The lateral variations of this heat
flow over the surface of the earth are small. If there are any appreciable local variations, they are
superimposed on the thermal effects due to vegetation, microclimate, etc. The latter are so large that
surface temperature measurements cannot (as a rule) be used for deducing the thermal conductivity
of rocks at depth or for determining the position of structures such a buried domes, anticlines, etc.
They have, however, been successfully applied in finding fissures and cracks along which convective
transfer of heat has taken place from depth through the agency of water (Kappelmeyer and Hähnel,
1974). Also, Pooley and van Steveninck (1970) showed in an excellent paper that temperatures only
about 2 m below the ground surface are substantially free from the above effects, and measurements
in short drill holes are capable of revealing shallow salt domes and similar structures by anomalies on
the order of 1 – 2 ºC (50 – 100 times the measurement accuracy).
Contrary to the behavior of the total heat flow, the vertical gradient of the temperature in the earth
varies within wide limits (5 – 70 ºC km-1) depending upon the thermal conductivity of rock formations,
and temperature logs in deep boreholes can be used with advantage in correlating stratigraphic
horizons.

5.11.5 Geochemical prospecting

I ndications of oil, ore and other minerals can be sought by chemical analysis of the solid and
gaseous substances of which the earth’s crust is composed. Such geochemical prospecting can be
carried out in the bedrock, in loose overburden or in the uppermost layer of the earth’s crust, namely
the surface soils (Ginzburg, 1960); Hawkes and Webb, 1962).
The analysis of the gases slowly diffusing from very great depths to the surface of the earth has
been employed in petroleum prospecting for quite some time. In this case, the contents of
hydrocarbons, such as methane, ethane, propane, etc. (which are intimately associated with oil), are
determined in samples of gas collected just below the surface of the earth.
In mineral prospecting the object is to determine the metallic contents at different places on the
earth. Metal ions can migrate through a number of agencies like weathering, leaching, wind
transport, etc., several tens or even hundreds of meters from the parent ore body and create ‘halos’
of mineralization around it. The metal concentration in such halos is in general very small and refined
colorimetric and spectrographic methods are needed for reliable determinations. The tracing of halos
is reported to have led to the discovery of a large number of mineralized zones in different parts of the
world.
The phenomenon of the migration of metals within the earth’s crust is an exceedingly complex
one. Nevertheless, great progress has been made in its study. The importance of geochemical
prospecting will be apparent from the fact that many metals such as molybdenum, tungsten,
vanadium, etc., which are of vital importance in modern industry, occur in the earth’s crust in
quantities far too small to alter appreciably the physical properties of the host rock so that their
detection by geophysical means is out of the question.

Conspicuously absent from all the previous discussions are the topics of gravity and airborne
gravity, both important in modern oil and gas exploration and both to be discussed at a later date.

154
5.12 Well velocity surveys

5.12.1 Introduction

W ell velocity surveys are made by using a seismic energy source (dynamite, Vibroseis, air gun,
etc.) at or near the surface normally within 300 meters of a deep borehole and recording the
initial direct wave and short leg reflected waves with a special downhole geophone (Figure 5.13).

Vibrator
energy Wireline
source Recorder

Direct wave

Datum

Short leg
reflected Velocity tool
wave Reflector

Figure 5.13 Well velocity survey, vibrator source

The geophone (velocity tool) is lowered into the hole and clamped at incremental depths, the
interval of sounding ranging from 50 ft to 1000 feet, depending on the type of survey and the
geophysical requirements. Consideration is given to important geological markers by positioning the
sonde just above the horizon and just below it. Zones of borehole washout are investigated in the
previously run caliper / gamma-ray log so that when the sonde approaches that zone, its clamp depth
can be adjusted up or down to avoid it. Observed downhole arrival times are corrected to vertical
times from a selected datum elevation.
The two principal types of well velocity surveys, sometimes referred to as well seismic surveys,
are the checkshot and vertical seismic profiling (VSP) survey. In the checkshot survey, the source
(often air gun for economic reasons) is stationary and placed close to the well and only the direct
arrival first break is used. VSP, however, can be stationary or walkaway. In the stationary
configuration, the source: dynamite, air gun or vibrator, is placed either as close to the well as the
drilling location permits (vertical) or at a given distance from the well (offset). For walkaways, the
shots can extend successively from near the well to several kilometers away from it, Vibroseis and
dynamite being the most commonly used sources.
A well velocity survey is the most accurate method for determining vertical travel times to the
depths at which downhole times are recorded. It is therefore the best method of measuring average
velocities to those depths and the interval velocities between them. A time-depth graph from a well
velocity survey showing average and interval velocities is shown in Figure 2.10 on page 21.

5.12.2 Checkshot survey

C heck shots are often taken in the same borehole after having run an integrated sonic log in which
case the integrated sonic log times are adjusted to fit the checkshot survey (Figure 5.14). The
resulting time-depth graph provides the best available determination of both average and interval
velocities throughout the depth range of the combined survey.

155
The tick marks down the left side of the log in Figure 5.14 show the time-depth relationship
furnished by the unadjusted integrated sonic log. The two time values, 0.7105 s and 0.7420 s, seen
on the log at depths 8260 ft and 8840 ft respectively, are the vertical checkshot times recorded by the
borehole geophone at those depths.

Sonic Log
Depth Adjusted
below Check TT below
KB shot KB
Delta T (DT) time
200 µs / ft break 40

Integrated 8200
sonic log time Check shot time =
= .6967 .7105 s @ 8260 ft
@ 8260 ft
8300

8400

8500

8600

8700

Integrated 8800 Check shot time =


sonic log time
.7420 s @ 8840 ft
= .7280
@ 8840 ft
8900

Figure 5.14 Integrated sonic log with checkshot corrections

The tick marks down the right side of the log show the time-depth relationship from the sonic log
after adjustment to fit the checkshot survey. For example, the adjustment at depth 8260 ft is (0.7105
– 0.6967) = 0.0138 sec. The adjustment at depth 8840 ft is (0.7420 – 0.7280) = 0.0140 sec. The
differences in adjustment values are distributed linearly between checkshot depths. The resulting
derived interval velocities are only as valid as their proximity to the well.

5.12.2.1 Checkshot QC

A s mentioned previously, only the first break of the direct arrival is used in the checkshot survey
and only one or two good shots are needed at each level to provide information for time-depth
corrections. Prior to the start of actual recording, several test shots should be fired and recorded in
order to check for timing breaks and system and environmental noise. In both types of velocity
surveys, some shots should be taken on the way down to total log depth (downshots) in order to
compare the data with the shots taken at exactly the same levels coming up the hole (upshots). The
resulting two same-level transit times should agree within 1 msec.
While recording, the well seismic transit times should be compared with the integrated sonic log
transit times and should closely match if the well is vertical, the source is close to it and the sonic log

156
is good. The differences between the two, if any, will normally appear linear through the depth of log
since accumulated transit-time integration errors are generally linear. In addition, the seismic source
signature should be checked for repeatability. If the source is dynamite, the charge size, charge
depth, and if possible, the medium surrounding the charge should all be uniform. If air gun is used
the source should be held at a constant depth and sleeve pressure.

Figure 5.15 Checkshot QC record

The logging truck operator will normally provide 100% camera records of the shots, an example
of which is seen in Figure 5.15. The upper curve shows the source signature with the thin vertical line
indicating the surface geophone break, which in this dynamite survey, is the uphole time (6.60 ms); in
airgun surveys, it would be the hydrophone break. All shots should indicate a shot time break, but its
absence does not necessarily nullify the shot if the recording software routine contains a backup
algorithm for registering time breaks. The lower curve is the direct arrival wave registered at the
downhole phone, the thin vertical line indicating the time of the first break, in this case 356.77 ms.
The seismic tool used in Figure 5.15 was a Schlumberger CSAT tool with a three-axis geophone
(x, y, z), however, only the ‘z’ signal was processed for this checkshot survey. In addition to the
three-axis geophone recording ability, the SAT tools can digitally transmit the seismic signal up the
hole eliminating signal distortion due to cable transmission.
Other Schlumberger tools include the WST well seismic tool, a single axis tool designed
especially for checkshots, the QSST quick shot seismic tool, similar to the WST but can be combined
with other logging tools, the MWST monocable WST which has a 2” reduced diameter, and the DSA
downhole seismic array which is a multilevel, single-axis array where a single shot is simultaneously
recorded from each of eight shuttles separated by an equal length of bridle cable, generally 20 m.
Each shuttle can be damped magnetically to the casing. The array is connected to the SAT tool,
which digitizes the waveform and sends the data to the surface by telemetry link.
All well seismic tools have a spring-loaded clamp that allows the geophone to be properly
coupled with the borehole wall. Good coupling in uncased boreholes is essential in well seismic
surveys. Prior to running the survey, the gamma-ray/caliper log should be reviewed for locations of
possible washouts. Where they exist, the operator can raise or lower the geophone from the
preselected level as necessary to find better coupling. Figure 5.16 shows a washout zone at 3300 ft
on the caliper curve of a combined sonic log run. The well velocity-logging operator, having previous
knowledge of this zone, clamped the geophone at 3340 ft and avoided the washout.

157
Borehole diameter Caliper Delta T

Figure 5.16 Sonic log with γ-ray and caliper curves with borehole diameter

5.12.2.2 Checkshot summary

CHECKSHOT SUMMARY LISTING CHECKSHOT SUMMARY LISTING

Stack Measured Measured Stack Measured Measured


number depth transit number depth transit
(ft) time (ft) time
(ms) (ms)

Figure 5.17 (a) Checkshot summary listing of depth-time pairs

158
A t the conclusion of the checkshot run, the operator initially produces two records: A checkshot
summary listing (Figure 5.17 (a)) and depth-versus-transit time plot (Figure 5.17 (b)). The
checkshot transit times are then used to correct the integrated times of the sonic log as shown in
section 5.12.2 and along with the true vertical depths corrected to seismic datum, are used to
calculate interval and average velocities. The depth-time curve should be generally smooth; a
deviation from this could be the result of incorrect depth correlation for a particular level(s) or incorrect
hydrophone or geophone break detection.

Figure 5.17 (b) Depth versus transit-time plot

Another important record of the checkshot survey is the overall shot-receiver geometry. It is
necessary immediately for corrections to recorded transit times due to offset and is used later in the
processing center. At greater depths and with minimal offsets, the correction is usually less than 0.5
ms, however for shallow shots and longer source offsets the correction can be considerably more.
Required information consists of the distance and azimuth from the source to the wellhead and in the
case of a dynamite spread, the UTM coordinates of the individual shots. After these data are
acquired and applied to the recorded checkshot data, a corrected depth-transit time chart can be
produced along with accurate interval and average velocities. Appendix K contains a discussion of
checkshot geometry and provides examples of supporting documents such as tables of corrected
transit times, interval and average velocities, operators log, frequency spectra, and frequency
attenuation chart.

5.12.3 VSP surveys

I n general, the methods used in a checkshot survey, as discussed in sections 5.12.2 and Appendix
K, apply as well to a VSP survey. The differences are that in the checkshot survey, the source is in
a stationary position whereas in VSP, the source can be stationary or sequentially offset such as is
the case of walkaways; the primary energy source used in checkshots is the airgun and for VSP,
airgun, vibrator and dynamite are all used; for checkshots, one or two shots per depth level is usually
sufficient, while in VSP, stacks of 3 to 7 shots per level (preferably an odd number since most

159
stacking routines stack peripheral traces to a weighted center trace) are needed, primarily for noise
cancellation; only the first break is of interest in the checkshot survey while in VSP, the first break as
well as the short-leg reflected arrivals are both of interest.
As in other types of well logging, the data acquired from a VSP survey is valid only the vicinity of
the well. The benefit of VSP is that in ideal conditions (e.g. noise-free surface and an uncased
borehole) the first break is multiple free and the reflections recorded are, in general, more reliable as
they are short-leg reflections and not subjected to the same signal deviations as in surface offset
seismic surveys. Whether to use a stationary or walkaway VSP is a matter of debate and
geophysical end use. Originally, the principal behind VSP was that its verticality and proximity to the
well produced a geophysical advantage in that it could extract pure, multiple and noise-free data with
geological reliability. Later, it was discovered that by sequentially offsetting the shots (walkaway), the
data was more correlative to surface seismics, including the presence of multiples, but had the
advantage of known geology in the vicinity of the well.

5.12.3.1 VSP acquisition and field plots

T he VSP survey generally follows the same methodology as in checkshot surveys the primary
difference being the source used and its physical relationship to the well. Vibroseis and airgun
offer an efficient means of multiple shots per depth level while dynamite, probably more costly,
provides the best repeatability and the widest frequency spectrum of the three sources. In addition,
consideration is given to the seismic data set available for the area around the well. If, for example,
the well is located on or near Vibroseis data, than a vibrator source might provide a better correlation
to the seismic data with the same logic applied to dynamite data.
An observer’s log is required for each job. This consists of a ‘seismic job summary’ and ‘job data’
sheets. The seismic job summary provides information about the equipment used, well data, source
details, acquisition geometry, survey details and surface velocity. Land jobs require determination of
the velocity between the source and the seismic reference datum (SRD). When this cannot be
determined by a shot at SRD, a weathering zone survey should be performed, and the weathered
zone thickness and velocity computed at the well site. The job data sheet is a sequential shot record
similar to Figure K.5 in Appendix K with explanations given for anomalies in the comments section.
Figures K.6 and K.7 are frequency analyses of the stacks are also performed in the field.
Real-time VSP plots are recommended if an array processor is available. Examples of these are
shown in Figures 5.18 (a) to 5.18 (h).

400 -500

(a) (b)

2400 1500

Figure 5.18 (a) Stacked data with band (b) Downgoing aligned without a
pass filter and true amplitude recovery velocity filter

160
800 -500

(c) (d)

2800 1500

Figure 5.18 (c) Upgoing aligned without (d) Downgoing events velocity filtered
velocity filtering

Plot (a) consists of band-pass filtered stacks (with true amplitude recovery applied) that have
been corrected for gun geometry, true vertical depth and seismic datum. This plot is in one-way time,
and along with plots (b) to (g) is plotted in SEG normal polarity. It is useful for observing data quality
and to check if coherent noise is present. In addition, any gross error in the reference signal time for
a level will be apparent.
Plot (b) is the same data as plot (a) that have been aligned along the first-arrival times and is in
an arbitrary time scale. This plot highlights any gross error in a transit time pick of a level.
Additionally, source consistency between levels can be examined for quality control. Downgoing
primary and multiple events have been aligned.
Plot (c) consists of plot (a) data where each individual level has been shifted by its transit time to
give a two-way time scale. Major reflected events now appear vertically aligned and dipping events
can be recognized. Direct comparison with surface seismics can now be made.
800 -500
Zero phase

(e) (f)
Normal polarity

2800 1500

(e) Upgoing events velocity filtered (f) Downgoing events velocity filtered waveshape decon

161
Plot (d) consists of plot (b) data on which a nine level median velocity filter has been applied.
This filter attenuates upgoing energy resulting in downgoing primary and multiple events in an
arbitrary time scale. Comparison of this plot with plot (b) indicates how effective the median filter is in
eliminating any remaining noise spikes that appear on individual levels. A ‘phase-in/phase-out’ option
can also be used. This option reduces the number of ‘lost’ levels at the beginning and end of the data
set to as few as two by reducing the number of levels in the median filter at the beginning and end of
the survey. The number of levels used in the velocity filter can be reduced to seven or even five.
Plot (e) is the same data as plot (c) on which a nine level median velocity filter has been applied.
This filter attenuates downgoing energy resulting in upgoing primary and multiple events in a two-way
time scale. With the downgoing events now removed, upgoing events can be clearly seen and a
rough comparison with the surface seismics is now possible. A ‘phase-in/phase-out’ option, as
described above can also be used.
Plot (f) consists of plot (d) data on which a band-limited waveshape deconvolution operator of
specified phase, high and low cut has been applied. The operator has been designed and applied on
each of the levels of plot (d) and it attenuates downgoing multiples resulting in downgoing wave-
shaped events in an arbitrary time scale. It should be chosen by reference to the clients seismic
section. This provides a quality check on the deconvolution operator in attenuating multiple events
and replacing the source signature with a minimum- or zero-phase wavelet. The deconvolution
operator has a 500 ms window beginning 60 ms before the first break. It has 5% white noise added
to stabilize the operation.

800 800 800

Zero phase
Zero phase

Zero phase
(g) (h)

Reverse polarity
Normal polarity
Normal polarity

2800 2800 2800

Figure 5.18 (g) Upgoing events, velocity (h) Corridor stack, two-way time
filtered and waveshape decon applied

Plot (g) consists of plot (e) data on which the band-limited wave shape deconvolution operator
computed on each of the levels in plot (f) is now applied to the corresponding level of plot (e). The
deconvolution attenuates upgoing multiple events resulting in upgoing wave-shaped primary events in
two-way time. As in plot (f), it should be chosen to match the client’s seismic section.
This display can now be compared directly to the surface seismic data. Short-order multiples
generated in layers traversed by the VSP survey can be pinpointed on both surface seismic and plot
(g). Higher frequencies will also be enhanced on the VSP for better bed resolution. Time to reflected
events below TD (total depth) could now be determined.
Plot (h) is a corridor bounded by the first-break time of each level and extending for a specified
time window down the traces and stacked to produce a single trace repeated several times for display
visual enhancement. The VSP data can now be considered as an ideal multiple-free seismic record
at the location of the well. This is plotted in two-way time in both SEG normal and reverse polarities

162
for ease of comparison with the surface seismic and synthetic seismograms (section 5.8.3.2).
The choice of real-time VSP plots can be limited to plots (g) and (h) if a rapid evaluation is
needed or to plots (a) and (b) for quality control. The preceding eight real-time VSP plots are from
the Schlumberger logging help file.

5.12.3.2 VSP quality control checks

(a) All sources: 1) Check that the downhole signal break is present on the record. If the geophone
break is not recorded, the data is unusable. Care should be taken while setting blanking times. A
minimum of 150 ms between the end of blanking time and the break of the geophone is
recommended. On the other hand, a too short blanking time may produce a long section of
waveform without signal before the geophone break.

2) Check that the signal is free of noise (Figure 5.19). If noise is present, confirm that its
frequency is outside the useful range (in general, 10 – 70 Hz for airgun and Vibroseis, 10 – 120
Hz for dynamite). In the case where the noise is outside these ranges, it will be possible to filter
it out during processing without any noticeable loss of data quality. Otherwise, try to find the
source of noise and eliminate it. Common sources of noise are:

Electrical noise: 50 or 60 Hz or harmonics (Figure 5.19 (a)). If possible, request a shutdown


of the rig generators during the shooting. Turn off the electrical motors in sequence until the
cause of the noise is discovered

Borehole casing ringing: This noise, originating mostly from uncemented and/or concentric
casings, occurs within the usable frequency range of the signal. Shot stacking is of little help.
Data will not be usable for VSP, but if the break is detectable, it can still be used for
checkshots. Wait until the cement is totally cured (often one or two days) before starting the
seismic job and then run a CBL log to check cement quality. This is of particular importance
when running a large offset (walkaway) VSP with an array tool.

Casing arrivals: These are generally of high frequency and often occur before the normal
geophone break.

Poor tool-formation coupling (Figure 5.19 (b)): Refer to the caliper log to preselect shot
levels. Avoid trying to clamp in caves or washouts. If the tool unwittingly enters one of these
zones, move it up or down a few feet to locate a more solid clamping location. If a movement
of a few feet doesn’t work, keep trying further up or down or go on to the next preselected
level as a shot without proper coupling is generally unusable.

Figure 5.19 (a) 60 Hz electrical noise

Figure 5.19 (b) Bad coupling to borehole wall

Road noise/cable noise: Let the noise decrease before firing. Slack off the cable after the
tool is anchored (10 ft per 1000 ft of well with an upper limit of 15 ft / 1000 ft in high deviation)
and as a last resort, use a bridle.

163
High-frequency bursts (Figure 5.19 (c): Analyze the frequency content of the noise and
select a filter to eliminate it (a high-cut of 90 Hz is generally adequate). If the surface seismic
target data is shallow and of a higher frequency than 90 Hz then the HF bursts should be
investigated. If possible, monitor the activity, scanning the whole waveform, and wait it out,
shooting during ‘quiet’ periods. For checkshots, where only the first break is needed, the HF
burst occurring after the break may not matter.

Figure 5.19 (c) High-frequency ‘bursts’

3) Verify that enough shots are recorded for each level. For checkshots, two shots are
recommended for airgun and vibrator while for dynamite, one may suffice. For VSP, a minimum
of five shots is recommended for airgun and Vibroseis and a minimum of three shots for
dynamite. If the signal-to-noise ratio is low (Figure 5.19 (d)), more shots need to be taken.
When there is no coherent noise on all shots, the signal-to-noise improvement is of the order of
the square root of the number of shots stacked (see section 2.5). However, it may be more cost
effective to modify the job setup than to increase the number of shots.

Figure 5.19 (d) S/N ratio too low

4) Check for the presence of tube waves (Figure 5.19 (e). Tube waves are coherent noise
that spans all frequencies (typically, they can be observed with high amplitude traveling at about
5000 f/s, after the geophone break). As they are difficult to filter out, they need to be avoided
during recording by optimizing the job setup.

Mud arrival

Figure 5.19 (e) Tube wave (mud arrival)

Lower the level of fluid in the borehole by a minimum of 50 ft. On jackup rigs, position the
gun between the well and the rig to avoid direct transfer to the mud column. On land, create an
impedance barrier by digging a water-filled trench between the wellhead and the seismic source.
The source pit could be placed behind an existing drainage pit.
Keep a large distance between the source and the well so that the ground roll is mostly
reduced before reaching the wellhead. A too-large distance (>100 m) is not recommended for a
VSP, as the vertical ray-path assumption would no longer be met, especially in shallow wells.
Section 5.12.3 argues the case of vertical ray path vs. offset.
Spot viscous fluids in the borehole close to the surface. For example, in the top part of
Figure 5.20 (a), no useful information is recorded. 10 bbls of viscous mud were spotted and 300
feet of borehole left empty. The resulting record (Figure 5.20 (b) was dramatically improved.

5) Check that the stack transit time matches the individual shot transit times. Any difference
should be less than 1 ms.

6) Check that the signal does not saturate. This can be verified by checking the peak-to-

164
peak value. It should not exceed 60,000 bits. Saturation may be directly visible on the monitor.

(a)

(b)

Figure 5.20 Viscous mud spotted near the surface

7) Check that the level separation is close enough to avoid energy aliasing. Referring to the
integrated sonic time provided by the sonic log, a shot should be taken every 7 ms or less to give
an alias frequency of at least 71 Hz. Important levels are the top and bottom of each sonic run in
the well, total depth, seismic reference datum (SRD), 5 m below major lithology changes, and
below and above major washouts where the sonic may not be reliable.

8) Check for repeatability of source signature and transit time. For airgun source, the canon
should be held at a constant depth and to a constant pressure; for dynamite, same charge size,
depth and medium; for Vibroseis, vibrator accelerometer, timing, and wireline and radio similarity.
For transit time repeatability of multiple shots at the same level, the difference should be less
than 1 ms. Check a level at a given depth on the way down the borehole and also coming back
up. The two transit times should agree within 1 to 2 milliseconds. In deeper wells, several tie
points should taken. Between different surveys in the same hole under the same source
conditions at the same depth, the times should agree within 2 ms. Also compare the well seismic
transit times with integrated sonic transit times, they should closely match.

9) Select proper filtering. It should be noted that the maximum frequency for a bandpass
filter high-cut is a function of the source, formation velocity and distance between levels. For
constant depth spacing of levels, the maximum frequency for a high-cut filter is determined by the
minimum formation velocity over the surveyed interval divided by twice the level spacing. For
constant time spacing, this relation simplifies to the reciprocal of twice the one-way time spacing
of the levels. These criteria should be used when selecting high-cut filters for the well site job
setup.

(b) Dynamite source: All of the previous quality control checks apply to dynamite as well as other
sources with a few special additions. Prior to the well seismic survey, charge size and depth
tests should be run to obtain an optimum in each case. Often, drilling deeper (preferably below
the weathering layer) may place charges in a more favorable shot medium, producing higher
amplitudes and wider bandwidth.
The selection of charge size affects the pulse width and amplitude of the pulse. If M = charge
size, the time duration of the pulse and its amplitude are both proportional to M1/3; the bandwidth
of the pulse is inversely proportional to M1/3; and the amplitude of the spectrum is proportional to
M2/3. For example, when a charge size of 2 pounds of dynamite is compared with a charge of 16
pounds (8 times greater), the time duration of the pulse with the larger charge will be doubled,
the amplitude of the pulse doubled, the bandwidth of the pulse of the larger charge halved, and
the amplitude of the spectrum quadrupled. In other words, lowering the charge size decreases
the pulse width and increases the frequency spectrum but at the cost of amplitude. One way to
generate higher frequencies and higher amplitudes is to use a simultaneous array of 3 to 5
smaller charge shot holes instead of one with a large charge. However, this implies increased
drilling costs.
If the shot medium on one side of the well site is not favorable to optimum energy transfer, try

165
moving the shot field to another side of the well. When shooting several shots at each depth
level, the shot pattern (3 to 5 shotholes) should be drilled so that the individual shots are close
enough not to change the shooting geometry yet far enough apart not to sympathetically
detonate each other. In addition, the pattern should be in a uniform medium, drilled to identical
depths and the charges should be well tamped.
Extra shot holes should be drilled and loaded for initial testing and for replacement of the
inevitable misfires. Time breaks and uphole times should be checked prior to shooting with
uphole simulators and also monitored during the execution phase. During the VSP, activity in
and around the shot field should be at a minimum both for noise reduction and for safety.

(c) Airgun: Arguably the most cost-effective of the three main energy sources, airgun is the most
widely used in checkshot surveys and utilized in many VSP surveys. The energy is the result of
the pressure created by a sudden injection of air into the sleeved chambers of the canon barrels.
This pressure produces air bubbles, which in turn produce acoustic waves in a water medium.
The amplitude of the signal depends on the pressure and the pressure depends on the gun
volume (V). Specifically, the pressure is proportional to V1/3. The effect of increased pressure or
volume on the amplitude and frequency spectrum of the signal pulse is similar to that of dynamite.
In general, the frequency range of airguns is much less than that of dynamite, which for
checkshot is not as significant as it would be for a surface seismic correlative VSP survey.
Specific quality control checks for an airgun well velocity are:

1) Check that the hydrophone signal has a clear break. Data from a seismic job without a
good break is virtually useless. Common problems to look for are (one): crosstalk between the
gun fire and the hydrophone signal, the former being mistaken for the time break resulting in a
corrupted signal and a bad first break detection and (two): during an offset survey, the radio TB
signal may be recorded instead of the correct hydrophone break. In the latter case, a longer
window may be used.

2) Use the hydrophone signal for VSP. The accelerometer signal may be acceptable for
checkshot surveys where only the break time is required. Normal practice would imply the use of
two hydrophone signals or one gun accelerometer hydrophone. Both the hydrophone and
geophone should have the correct break detection. It is mandatory that the hydrophone
detection be correct before stacking. The type of break-pick algorithm used depends on the
source type and must be selected at the beginning of the survey. The bubble period can be
estimated from the hydrophone signal and filtered out during processing.

3) As well as checking for signal saturation (paragraph (a6) above), check that the signal is
not too weak. Peak to peak amplitude before filtering should be larger than 30,000 bits. If not,
(a): use a larger gun or gun array, (b): check the air pressure (maximum is 140 bars or 2000 psi,
(c): increase downhole gain for the channel, and/or (d): remove the debubbling kit from the gun (if
installed).

4) Check that the hydrophone break is constant within a given level. Check that the break
position on the operator unit is always occurring at the same place and that signal inputs are not
changed within a level

(d) Vibroseis: Check sweep quality. Cross correlation of the entire Vibroseis trace is not possible for
every sweep unless an array processor is present. Without an array processor, the ‘first break’
can be checked by setting the WCL to 300 or 400. It is also recommended to check at least one
sweep per level with a WCL of 1500 to 2000 to ensure a good signal-to-noise ratio and to check
for tube waves. It should be noted that similar checks should be performed several times during
the survey to confirm that the pilot sweep and base plate signals are correctly synchronized.
Check coupling between the Vibroseis unit and the earth. This is done to avoid baseplate
chatter. This can be monitored through direct observation of the baseplate when sweeping,
manual scan of the uncorrelated base plate sensor on the Optical Monitor Unit of the Central
Seismic Unit, or through the use of frequency spectra plots of raw and correlated data.

166
Appendix A

Seismic noise

Table A.1 Types of noise

Genre Type Description Appearance

Direct waves Waves propagating directly High amplitude first events


from shot on near traces
Ground roll Waves propagating along Usually low frequency and
the ground surface low apparent velocity
Air waves Sound waves from the shot Very low apparent velocity
Source propagating in the air (1100 ft/sec)
Generated Shallow refraction Near-surface refracted wave High amplitude, often first
(Coherent) events of far traces
Reverberations Ghosts, tails on shot pulse Tails on reflected events
mainly in marine data
Multiples (long period) Repeated reflections Reflections at even time
intervals after the primary
reflection
Diffractions Wave energy propagated in Coherent hyperbolic events
all directions from a point
such as a fault interface

Traffic Cars, animals, people and Often seen as spikes


objects in motion near line
Wind Air currents High frequency
Any motion produced from
Ambient Earth tremors within the earth by other than Usually very low frequency
(Random) seismic source
Shot hole Debris from shot blow-out High and low frequencies
emanating from source pt.
Highline noise Induced noise from power 50 or 60 Hz
lines
Instrument noise From recording instruments Usually high frequency and
low level

Noise is defined as any unwanted signal, that is, anything that alters the input seismic pulse other
than inherent earth effects with the ultimate goal of isolating reflection events. Table A.1 lists most
but not all of the many noise types.
Source-generated noise usually consists of events produced by the seismic energy source that
have some definable non-random pattern in space and/or time. They are normally described as
coherent events and include those shown in Figures A.1 and A3. Coherent noise events are often
distinguished by their apparent velocities. Apparent velocity is defined as the distance between two
receivers divided by the difference in the times a particular event arrives at each of these two
receivers. Figure A.2 shows a time-distance graph relating various coherent noises to their
respective velocities of propagation.

167
Direct
waves
R R R
R R R
S Refraction
Reflections Diffraction

Multiple

Diffraction Refractions

Figure A.1 (a) Illustration of source generated noise: diffraction, refraction and multiples

Air wave

R R R
R R R

Ground roll

(b) Illustration of source generated noise: air wave and ground roll

Air wave

Ground roll
Corrected time

Refracted (headwaves)
Reflected wave

Refracted wave overtakes direct wave


Direct arrivals

xc
Distance

Figure A.2 Time-distance graph illustrating relative velocities of source related noise

168
T he effect of water-layer multiples (reverberation) can be seen in Figure A.3 where the reflector
of interest is sufficiently deep compared with the source-hydrophone distance that all the rays
may be assumed to be nearly vertical, although for illustration purposes they are shown to be
slanting. Thus, S and C are almost identical points and so also are H and H’, and A and A’. It is
assumed for simplicity that the source pressure pulse is a single spike of unit amplitude and that the
reflector is a perfect one. The seabed reflection coefficient is assumed to be +Rc for pulses coming
from above and reflected back into the water. The water depth is h.
The direct reflection received at the hydrophone is SAH producing the signal 1.0 on the seismic
trace. Besides this, the hydrophone will receive the pulse SBCA’H’ reflected first at B on the sea
bottom, next at the air-water interface at C and then at A’. The first reflection reduces the pulse
amplitude to R. On reflection from the water-air interface (R = -1.0) the amplitude becomes –R while
there is no further change of amplitude at A’ as the deep reflector is perfect. This pulse, which is
double reflected within the water layer, arrives 2h / V seconds later that the primary reflection, which
is the additional time for the path SBC, since SB and BC are nearly vertical. Another double reflected
pulse, also arriving at the same time, is SAHB’H’. Its amplitude is also –R so that the total signal on
the trace at this time will be –2R.
Continuing in this way for pulses reflected three or more times between the water surface and the
sea bottom we find that the seismic trace will consist of spikes 1.0, -2R, 3R2, -4R3, … (Figure A.4)
following each other at the same interval 2h / V. Strictly speaking, all the amplitudes in this should be
multiplied by Tc,1, the transmission coefficient from sea to rock, and by Tc,2, the transmission
coefficient from rock to sea at the seabed.
S C H H’

+Rc
B B’ - Rc

Figure A.4 Spike series 1.0, -


Deep perfect 2R, 3R2, -4R3, … due to simple
A A’ reflector water layer multiples (ringing or
reverberation)
Figure A.3 Illustration of reverberation

This now becomes a problem for processing to remove the water-layer multiples (reverberation or
ringing as they are sometimes called) from the trace to obtain a series 1, 0, 0, … that corresponds
only to the response of the deep reflector. Multiple reflections as well as diffractions, refractions, the
change of pulse in transmission, etc. influence the final appearance of a seismic record but mainly the
record is the convolution of a basic wavelet with the long string (‘log’) of all various reflection
coefficients in the ground.
The procedure of convolution using discrete evenly spaced samples can be mathematically
written as follows. Let a = (a0, a1, a2, …) denote the series of reflection coefficients and b the pulse
series. Then the pth term in the final series c is given by

cp = Σm=0 bmap-m
p
(A.1) ξ

The desired series 1, 0, 0, … is treated as series c in the equation (A.1) and the known series 1.0, -
2R, 3R2, -4R3, … as a string of pulses b. Therefore the problem reduces to finding a series a such
that b convolved with a (b * a) gives c.

169
Appendix B

B.1 Rock density

Table B.1 Table of densities (kgm-3)

Material Density Material Density

Oil 900 Anhydrite 2960


Water 1000 Diabase 2500 – 3200
Sand, wet 1950 - 2050 Basalt 2700 – 3300
Sand, dry 1400 - 1650 Gabbro 2700 – 3500
Coal 1200 – 1500 Zincblende 4000
English Chalk 1940 Chalcopyrite 4200
Sandstone 1800 – 2700 Chromite 4500 – 4800
Rock salt 2100 – 2400 Pyrrhotite 4600
Keuper marl 2230 – 2600 Pyrite 5000
Limestone (compact) 2600 – 2700 Haematite 5100
Quartzite 2600 – 2700 Magnetite 5200
Gneiss 2700 Galena 7500
Granite 2500 – 2700

B.2 Acoustic impedance

T o illustrate the essential idea, only a compressional wave motion in a liquid will be considered.
Suppose the motion to be in the direction Ox. Each particle of the medium vibrates parallel to Ox
about a mean position. Let A and B be two planes, each of area α, with equilibrium positions x and x
+ dx. If the displacement at A at any instant is u, then that at B is u + (du / dx) dx and the linear
strain (change in length per unit length) is therefore du / dx. This is the total dilatational strain as
there is no displacement at right angles to Ox. By Hooke’s law (Section 1.0) the stress at A is P =
κ(du/dx) where κ is the compressional modulus, and the stress at B is P + (dP/dx)dx. The
corresponding forces on the planes are α times the stresses. Since the liquid between A and B is not
transported as a whole the forces on A and B must be oppositely directed, but this leaves an
unbalanced force

α(dP/dx)dx = κα(d2u/dx2)dx

which causes an acceleration d2u/dt2 of the mass ρα dx of the liquid between A and B, if ρ is the
density. Since force = mass x acceleration we get

d2u/dt2 = (κ / ρ)(d2u/dx2) (B.1) ξ

This is the equation of a wave motion with velocity V = (κ /ρ)1/2 and its general solution for
propagation in the positive Ox direction is

u = φ(x-Vt)

170
It can be verified that this satisfies the wave equation above. The velocity of the particles at x is

du/dt = -Vφ’

while we have seen that stress at x = κ(du/dx) = κφ’

The magnitude of the ratio stress/particle velocity is evidently

κ/V = κV/V2 = ρV. (B.2) ξ

This is independent of x or t and is a characteristic of the medium and the elastic wave in question
only. It is termed acoustic impedance, the units being pascal seconds per meter (Pa s m-1) or,
which is the same thing, newton seconds per cubic meter (N s m-3).

Appendix C
Logging methods

Table C.1 Logging methods and geophysical objectives

Log Symbol Parameter Target Geophysical Objective


Gamma-ray GR count of natural gamma natural radioactivity petrography, clay
radiation of rocks content, thin bed
Density D counts of Compton density of rocks fracturing, porosity
scattered rays
Neutron N counts of secondary lithology stratigraphy, porosity,
neutron-neutron rays permeability
Electric EL, ES apparent resisitivity true resistivity hydraulics, lithology

Microlog, ML, MLL unfocused electrode, true small scale hydraulics, lithology, thin
Microlaterolog focused electrode resistivity layers, fractures, fissures
Induction log IES, IEL focused electromagnetic true conductivity lithology, saline or
induction leachate infiltration
Focused electric FEL, LL focused electrode true resistivity of lithology
log, Laterolog rock
Self-potential SP probe to surface kinetic sources of electric oxidizing bodies, clay %,
& electrochemical potent. potential permeability
Salinometer SAL resistivity of borehole salinity total salt content of
fluid fluid
Temperature TEMP temperature of borehole geothermal field thermal gradient
fluid
Sonic Velocity SONIC travel time of seismic seismic velocity seismic data
waves interpretation
Caliper CAL borehole diameter shape of borehole corrections to other logs,
walls borehole deviations
Flowmeter FLOW revolutions of a spinner velocity of fluid flow zones of in and outflow
of water
Deviation DV compass & dipmeter inclination & azimuth spatial drill path
of borehole

171
Table C.2 Geophysical well log objective cross reference

GR D N EL ML MLL IEL FEL LL SP Sal Tmp SV Cal Flw DV

Petrography x
Clay content x x
Thin bed x x x
Fracturing x
Porosity x x
Permeability x x
Stratigraphy x
Hydraulics x x x
Lithology x x x x x x
Fissures x x
Saline infiltration x
Oxidizing bodies x
Salt content x
Thermal gradient x
Seismic data x
Log correction x
Borehole caving x
Water in/out flow x
Hole deviation (y,z) x

Appendix D
Resistivity Figure D.1 Resistivity of pore water with
dissolved salts
D.1 Pore water resistivity
50

P ure water is ionized to only a


small extent and the electrical
conductivity of pore waters
40
10K 5K 2K 1K 500 200 100 50 mg NaCl / l
30
depends on the presence of
Temperature (ºC)

dissolved salts, mainly sodium


chloride. Figure D.1 shows the 20
variation of water resistivity with
concentrations of dissolved NaCl
measured in milligrams per liter of 10
water. Also shown is the effect of
temperature. With any given
concentration of salts in water, as
0
the temperature increases, the 0.5 1 2 5 10 20 50 100 200
specific resistivity decreases. Resistivity (Ωm)

172
1.00

0.80

Fractional porosity
0.60

0.40
m = 1.8

m = 1.5
0.20
m = 1.2

0.00
0.00 0.20 0.40 0.60 0.80 1.00
ρw / ρ

Figure D.2 Archie’s law variation of bulk resistivity, ρ, for rocks with an insulating matrix and pore
water resistivity ρw. The index, m, is about 1.2 for spherical grains and about 1.8 for platy or tabular
materials.

D.2 Determination of resistivities

T he determination of the resistivity of rock samples need, in principle, only involve measuring the
current density in, and the potential difference across, a rectangular or cylindrical sample with
plane end faces. The application of equation (4.6) in section 4.5.1.3 then gives the resistivity. In
practice several problems arise due to the inhomogeneous nature of most natural samples and due to
electrode polarization caused by electrochemical action when a current is passed through a sample.
To overcome the latter problem it is necessary to make the measurements with sinusoidal current
or d.c. pulses of alternating polarity. Also, in order to diminish contact resistances between the
current electrodes and the sample faces, aluminum or silver foils, or wet sponges, are pressed
against the plane faces while the potential drop is measured between two electrodes placed on the
top or bottom faces of a rectangular sample or on the generator or of a cylindrical one. Another
method is to have a sample with one flat face and use the Wenner arrangement of four collinear
electrodes shown in Figure 4.14 (a) in which case equation (4.18) gives the resistivity, provided the
spread of the electrode array is small compared with the linear dimensions of the rock sample.
Methods of the above type, based on passing a continuous sinusoidal or direct current through a
sample, are not well suited for measurements of very high resistivities. For highly resistive rocks the
sample is, in principle, put in a circuit as a resistor in parallel with a capacitor of known capacitance C
that is initially charged to a voltage V0. The time constant τ of the arrangement, that is the time in
which the voltage across the capacitor falls to the value V0/e, is determined as the capacitor

173
discharges through the sample. The resistance of the sample R is then given by τ /C and the
resistivity can be obtained from equation (4.4).
High resistances, however, can be measured more efficiently by means of high-frequency
alternating current using a circuit that contains a known inductance L and a parallel-plate capacitor of
capacitance C. Here only the barest principle of the method can be described. The sample in the
form of a disc is placed as the dielectric in the capacitor. When the circuit is brought to resonance by
varying the frequency ν of the current passed through it, the condition 4π2ν2LC’ = 1 holds where C’ is
now the capacitance with the sample inserted between the plates. The ratio Q of the voltage at
resonance across the inductance (or the capacitor) to the applied voltage is measured. Circuit theory
shows that Q, which incidentally is also a measure of the sharpness of the resonance, is equal to
2πνL/R, where R is the resistance of the sample. As before, when R is known, equation (4.4) gives
the resistivity. This method also yields the ratio C’/C, which is the relative dielectric permittivity of the
sample.

D.3 Apparent resistivities for various dipole-dipole arrays

T he product of I and the length AB is called the electric moment of the current source. If AB is
made vanishingly small and I is increased at the same time, keeping I x AB constant, an ideal
point current dipole of moment I x AB ampere meters (A m) is obtained. In practice, the electric field
produced at a point P by a pair of current electrodes may be considered to be produced by a point
dipole if the distance of P from both A and B is large compared with the separation AB. If the
separation MN of the potential probes, between which P is situated, is likewise small compared to P’s
distance from A and B, then that is referred to as a potential dipole MN. If the lengths AB and MN,
instead of being small, are comparable to the distance between their center points, then that is a
bipole-bipole system. There also exists a hybrid system like bipole-dipole.
P

r
M

α θ
B(-) A(+)

Figure D.3 An electric dipole

The potential at point P of the current dipole in Figure D.3 on homogeneous ground is

Iρ 1 1
U= -
2π AP BP

Iρ BP - AP
U= (D.1) ξ
2π AP x BP

However, when the distance of P is large, AP ≈ BP = r and BP – AP ≈ BP – MP = BM = ABcosα, if


AM is the perpendicular from A to BP. When r is large the angle α may be put equal to θ, which is the

174
angle between BA and the line joining P to the dipole center. With all the substitutions made in
equation (D.1)
Iρ ABcosθ
U= (D.2) ξ
2π r2

as the dipole potential in uniform half-space. From this can be obtained the gradient of U in the
desired direction, namely the direction of the potential dipole. For example, if the potential dipole is in
line with the current dipole take the derivative of U with respect to r and since θ = 0 in this case

dU Iρ x AB
= -
dr πr3

Substituting for ( - dU / dr) the measured approximation to the gradient UMN / MN gives the
apparent resistivity corresponding to this dipole-dipole orientation. Several different dipole-dipole
configurations have been suggested and given characteristic names as shown in Figure 4.18 in
section 4.1.5.10. The derivation of the ρa formulae follows.

T he electric fields of a current dipole in the radial and azimuthal directions as well as parallel and
perpendicular to the dipole (provided AB, MN << r) are given as follows. For an azimuthal array

U/MN ≈ E = IρABsinθ/2πr3

for which the apparent resistivity for the configuration is defined as

2πr3 U (D.3a) ξ
ρa =
AB x MN sinθ Ι

Similarly for an axial configuration

πr3 U (D.3b) ξ
ρa =
AB x MN Ι

radial
πr3 U
ρa = (D.3c) ξ
AB x MN cosθ Ι

parallel
2πr3 1 U
ρa = (D.3d) ξ
AB x MN 3cos θ − 1 Ι
2

perpendicular
2πr3 1 U
ρa = (D.3e) ξ
3AB x MN sinθ cosθ Ι

equatorial
2πr3 U
ρa = (D.3f) ξ
AB x MN Ι

175
Appendix E
Dipole-dipole mapping

A collinear dipole-dipole configuration such as in Figure 4.9 (c) can be moved as a whole along
lines parallel to the array keeping the values of a and n fixed. The apparent resistivity (equation
(4.20) may be referred to the center of the array or to the center of one of the dipoles and plotted on a
map. The configuration is, however used in mapping in a somewhat different way as follows.

A B M N M N
V V

45º 45º

n=1

n=2

n=3

Figure E.1 Plotting of dipole-dipole ρa data

Measurements are made along a profile with a selected a (e.g. 60m) and with n = 1, at suitable
intervals (usually equal to a). In plotting the measurements, lines making an angle of 45º with the line
representing the profile are drawn from the centers of the current and potential dipoles in opposite but
converging directions, and the value of ρa obtained for that position of the array is plotted at the
intersection of these two lines (Figure E.1). The measurements along the profile are repeated for n =
2, 3, . . . ., and plotted in a similar way. It is easy to see that the measurement for n = 2 in such a plot
will appear along a line below the line on which those for n = 1 appear. The 45º line emanating from
the n = 2 line to the surface would intersect the surface horizontal line at the center of the MN dipole
in between the two shown. Those for n = 3 will be plotted along a line still ‘deeper’ and so on.
Contours of equal ρa are then drawn on this plot. The picture thus obtained is called a vertical
pseudo-section of the ground because, while measurements along a profile are indicative of lateral
conductivity variations, measurements for a larger value of n may be supposed to contain more
information about deeper inhomogeneities than those for a small n (see Figure 4.11). An example of
a pseudo section is found in Figure 4.28.
The idea of pseudo-section plotting is not exclusive to dipole-dipole measurements but it has
seldom been exploited for other arrays. For example, pseudo-sections can be easily constructed for
Schlumberger or Wenner mapping measurements, if made with different-sized arrays, by plotting
each value vertically below the array center at a distance proportional to AB / 2 or a respectively.

Appendix F
Fourier analysis

F.1 Time domain and frequency domain

T he output voltage from a geophone in seismic surveys resulting in a seismic trace or from a
voltmeter in electrical and induced polarization surveys yielding an electrical or IP map results in
a graph of that voltage ‘trace’ relative to the duration of time of the seismic signal or the flow of
current. That is, the graph appears as a series of voltage amplitudes relative to a time scale. Since

176
time is the independent variable, these graphs are said to exist in the time domain.
These time-domain data can be transformed into the frequency domain in order to analyze their
frequency content and enable the enhancement of the data by frequency domain filtering. In the case
of the latter, the amplitude and phase spectra will be changed in some way and then transformed
back into the time domain to produce a new and improved version of the original data. In the process
of the transformation, which is completely reversible, no information is lost. As shown in Figure F.1,
given the frequency amplitude and phase spectra, the information can be transformed back into the
time domain producing a time-dependent seismic wavelet such as that on the left-hand side of the
figure.

Time domain Frequency domain

100

Relative response percent 80


Amplitude
60 response

40
Impulse
response 20

0
20 40 60 80 100
Frequency (Hz)
Phase shift, degrees of lag

t 100
Phase response
0

-100

-200

20 40 60 80 100
(From Lindseth, 1970) Frequency (Hz)

Figure F.1 Time-domain signal and its equivalence in the frequency domain

The conversion from time to frequency domain is accomplished by a Fourier transform (also
called Fourier analysis). The reverse process, from frequency-phase to time domain, is called an
inverse Fourier transform (or Fourier synthesis). The same process applies whether the input data
are originally seismic, electric or electromagnetic; in this section the example of seismic traces will be
used.

F.2 Fourier analysis and filtering

A seismic trace is not a periodic function but it often does have a roughly rhythmic pattern,
especially where organized events appear, whether signal or unwanted noise (Figure F.2). If one
observes several reflections within a time span of, say, 500 ms (milliseconds), we will probably find
that time intervals between successive peaks or toughs tend to cluster fairly close to an average
value, perhaps 25 to 35 ms in an upper part of the trace and perhaps 40 to 50 ms in a lower part.
Since a period of 25 ms corresponds to a frequency of 1 / .025 = 40 Hz and a period of 40 s
corresponds to a frequency of 1 / .040 = 25 Hz, you can say that the shallow reflections are rich in 30-
40 Hz energy and the deeper ones peak in the 20-25 Hz range.
Also present in the field record of Figure F.2 are a couple of noise bursts, wind noise, and some
low-frequency noise. Not so evident but usually present is ground roll (8-12 Hz) and often, power line
pickup (50 or 60 Hz).

177
Assuming that all of these events are present and visible, it should become evident that the
esthetics of the raw data would be enhanced by band-pass filtering, using a passband of about 15-55
Hz or 12-48 Hz or maybe even a time-variant filter from 20 to 58 Hz in the upper half and 8-40 for the
bottom half of the trace. Sometimes a change of only 5 to 10 Hz can appreciably change the quality
of filtered data. What is needed then, in order to select the proper filter, is a procedure to analyze the
raw data for frequency content. This procedure is the Fourier analysis.

Noise burst
Reflection
Low-frequency Reflection
Noise burst Wind
noise noise

100 ms
Noise
burst

Figure F.2 Field record showing signal and noise of different frequencies

F.3 Fourier transforms

Given a time domain function f(t) which is defined in the interval t < t < (t + T), the function can be
approximated by the Fourier series, S (t) as follows
n
1 1

n
2πkt
f(t) ≈ Sn(t) = 2 + ∑ ak cos
a0 2πkt
+ bk sin (F.1) ξ
k=1 T T

where k = the harmonic index

The d.c. shift is defined as


t1+T
a0 1
2
=
T t1
f(t)dt (F.2) ξ

and the cosine transform as


t1+T
2 2πkt dt
ak = f(t) cos (F.3) ξ
T t1 T

178
and the sine transform as


t1+T
2 2πkt dt (F.4) ξ
ak = f(t) cos
T t1 T

The approximation can be improved by adding more terms (making n larger). At the limit, the
series converges on f(t) as

n→∞
a0
f(t) = lim Sn(t) = 2 + ∑ akcos 2πkt + b sin 2πkt
T
k
T
(F.5) ξ
k=1

The concept of adding together separately weighted sets of sine waves and cosine waves to
approximate an arbitrary function is suitable for synthesis but it is not very convenient for analysis.

Example F.1

aq = 1
ap = 2

apcos 2πpt aqcos 2πqt


T T

bq = 3
bp = 2

bpsin 2πpt bpsin 2πpt


T T

Figure F.3 Weighted sine and cosine waves for two harmonic frequencies in a Fourier series

If p and q are two different harmonics as shown in Figure F.3, calculate which harmonic has the
greater amplitude if equations (F.3) and (F.4) result respectively in the following values.

ap = 2 aq = 1

bp = 2 bq = 3

It turns out that the sum of a weighted sine wave and a weighted cosine wave of the same
frequency can be represented by a single weighted cosine wave (also of the same frequency) which
has been shifted along the time axis. Thus

akcos 2πkt + b sin 2πkt = Ak cos


2πkt - ϕk (F.6) ξ
k
T T T

where for the kth harmonic frequency, Ak is the amplitude and ϕk the phase.

179
The amplitude and phase lag are determined by ak and bk as follows:

Ak = √ ak2 + bk2 (F.7) ξ

bk
ϕk = tan-1 ak (F.8) ξ

Now the problem of which harmonic has the greater amplitude can be resolved:

Ap = √ 22 + 22 = 2.83 Aq = √ 12 + 32 = 3.16

Therefore, the q th harmonic has the greater amplitude.


In addition, the respective phase lags for the two harmonic frequencies are

2 3
ϕp = tan-1 = .25π ϕq = tan-1 = .40π
2 1

The expression for phase lag in equation (F.8) follows an arbitrary convention. Some prefer the
concept of phase angle, which is the negative of phase lag; that is

-bk
Phase angle = tan-1 ak
F.4 Fourier transform procedures

F igures F.4 and F.5 illustrate four steps of the Fourier transform. The ‘signal’ in Figure F.4 is
shown as a continuous function, but it has actually been sampled at intervals of 2 ms. It is 500
ms long and is shown here for convenience to be centered at time zero. If this signal is moved to a
different position on the time axis, the amplitude would be unaffected but the phase lag spectrum
would be tilted by a predictable amount.
-0.250 s 0 0.250 s

Signal

22 Hz Cosine wave

Figure F.4 Seismic signal with a 22 Hz dominant frequency and


a pure 22 Hz cosine wave

Since the length of the signal is 0.5 seconds, the period of the fundamental frequency component
(first harmonic) is also 0.5 sec., and the fundamental frequency is therefore (1/ 0.5) = 2 Hz.

180
(a) Cosine transform The first step is to multiply the signal by a 2 Hz cosine wave, sample by
sample, and add the products. The product-sum is then normalized to some convenient scale and is
plotted as the cosine transform value at 2 Hz on the frequency axis.
In a similar way, the given signal is multiplied with cosine waves of other harmonic frequencies (4,
6, 8, …, Hz etc.) and the normalized product-sums are plotted against frequency in the cosine
transform (Figure F.5). Figure F.4 shows an example of a harmonic frequency of 22 Hz.

40 40
Cosine transform Sine transform
20 20

0 0
Amplitude

Amplitude
20 40 60 80
Frequency (Hz) 20 40 60 80
Frequency (Hz)

40
Fourier amplitude 360
spectrum 180
20 20
0
0 40 60 80
-180
Frequency (Hz)
Amplitude

20 40 60 80
-360
Frequency (Hz) Fourier phase lag
-540
spectrum
-720

Figure F.5 Four products of the Fourier transform (cosine, sine, amplitude and phase).

(b) Sine transform In a like manner as that of the cosine waves, the signal is multiplied with sine
waves of the same harmonic frequencies. The normalized product-sums are then plotted against
frequency in the sine transform.
The maximum frequency that can be used in a Fourier analysis, without aliasing, is less than or
equal to the Nyquist frequency. In this case, the Nyquist frequency is 1 / (2 x 0.002) = 250 Hz.
However, in the above example it appears that frequency components above about 70 Hz are
negligible and the transforms were therefore computed only up to 80 Hz.

(c) Amplitude and phase lag spectra The amplitude spectrum can now be computed from the values
obtained from the cosine and sine transforms. Each amplitude value is the square root of the sum of
the squares of the corresponding cosine transform and sine transform values (equation (F.7)), again
normalized to a convenient scale.
Finally, the phase lag spectrum is computed (equation F.8)). Each point plotted represents the
angle whose tangent is the ratio of the corresponding sine transform and cosine transform values.
The 22 Hz cosine wave shown in Figure F.4 turns out to be a better fit to the signal than any of the
other harmonic cosine or sine waves (thus referred to as the dominant frequency). This is the reason
why the value plotted at 22 Hz in the cosine transform is larger than any other value in either
transform.
Note that the discrete values of the amplitude and phase lag spectra have been connected by
continuous line segments. This ‘interpolation’ is sometimes done for cosmetic effect only. The

181
resolving power of the spectral analysis is limited to 2 Hz because the given signal was only 500 ms
long. If the signal had been 2 seconds long, the resolution in the frequency domain would have been
0.5 Hz.

(d) Fast Fourier transform and the power spectrum It is evident from the preceding discussion that
the classical Fourier transform process involves a formidable number of calculations. For years, the
development of frequency domain processes was retarded by considerations of computer time and
cost. Since the late 1960s, however, several computing algorithms have been developed which take
advantage of certain symmetrical relationships in the cosine and sine transform steps to reduce
computing time greatly. The best known of these more efficient procedures is the Cooley-Tukey fast
Fourier transform (FFT). For and input signal of about 1000 samples, the Cooley-Tukey FFT is
about 1000 times faster than the conventional transform.
If the frequency domain data is to be transformed later back into the time domain, then both the
amplitude and phase spectra must be computed. To do that, both the cosine transform and sine
transform must be computed. If what is needed is solely an analysis of the frequency content of the
data, then time and money can be saved by computing the power spectrum.

Time domain Frequency domain

Amplitude spectrum = A(f)


Signal = f(t)
Phase spectrum = ϕ(f)

Autocorrelation

Amplitude spectrum of autocorrelation


Autocorrelation = power spectrum of signal = A2(f)
of f(t) = ϕff(t)
Phase spectrum = 0

Figure F.6 Relationships involved in the power spectrum

The power spectrum of a time domain signal is the Fourier transform of the signal’s
autocorrelation. The power spectrum is also the square of the signal’s amplitude spectrum. Since an
autocorrelation is an even function, its cosine transform is also its Fourier transform. Figure F.6
shows the relationships involved.

Appendix G

Interpretation of magnetic remanence

IfκT.T isIntheTablefieldG.1strength associated with the earth’s flux density, the induced magnetization will be
are given the values of the Koenigsberger ratio Q = M κT where M is the natural
n n n
remanent intensity. From this it is observed that in most igneous rocks like gabbro and basalts, in kiln
bricks or in iron objects, Mn completely dominates the intensity induced by the earth’s field strength.
From a study of the magnetic anomalies, it is possible to obtain an estimate of the effective
magnetic moment m of a mass of postulated geometry. This is the vector sum of the moments due to
induced and remanent magnetism. If V is the volume of the mass, then equation (4.40) can be
rewritten more generally as follows

m = VκT + VMn (G.1) ξ

182
In the simplest case, when T and Mn are parallel,

m = VκT + VQκT (G.2) ξ

from which is obtained

V = m / [κ(1 + Q) T] (G.3) ξ

If remanence is neglected (Q = 0) then the volume estimate m / κT is obtained which is larger than
the true volume given by (G.3). Similarly, if T and Mn are antiparallel, the volume estimate obtained
by neglecting remanence will be smaller than the true volume m / [κ(1-Q)T]. In the general case of
rocks, the dips of formations may be incorrectly estimated if the vector Mn in equation (G.1) is
disregarded. A discrepancy between observed dips and those obtained by magnetic interpretation,
especially in areas where igneous rocks occur, can be very frequently traced to the existence of a
remanent intensity.ii

Table G.1 Values of Qn for some rock specimens (Parasnis, 1997)

Rock Specimen Locality Qn


Basalt Mihare volcano, Japan 99-118
Gabbro Cuillin Hills, Scotland 29
Gabbro Småland, Sweden 9.5
Andesite Taga, Japan 4.9
Granite Madagascar 0.3-10
Quartz dolerite Whin sill, England 2-2.9
Diabase Astano Ticino, Switzerland 1.5
Tholeiite England 0.6-1.6
Dolerite Sutherland, Scotland 0.48-0.51
Magnetite ore Sweden 1-10
Manganese ore India 1-5
Sediments -- generally low

Appendix H

Quasi-static, near and far electromagnetic fields

H.1 Dielectric permittivity

I t can be shown that if there is a time-dependent electric current instead of a steady one, there is at
every point in a medium not only a magnetic field but also an electric one, both varying with time.
Both are dependent on the magnetic permeability µ (section 4.6.2) and the dielectric permittivity ε
(often also called by the older term ‘absolute dielectric constant’) of the medium.
The concept of the dielectric permittivity of a medium is most easily grasped by noting that it is
the property which determines the magnitude of the force (in newtons) between two ‘massless’
electrical point charges, Q1 and Q2 coulombs, situated in the medium. If the charges are separated
by a distance r, then the force is Q1Q2 / (4πεr2). The permittivity of free space is denoted by ε0 and its
magnitude is defined in SI as 107 (m Ω-1 s-1) divided by 4πc2 (m2 s-2), where c is the velocity of light in
vacuum. The dimensions of permittivity are therefore Ω-1 s m-1 (or S s m-1, siemen seconds per
meter, or F m-1, farads per meter) and for vacuum ε0 = 8.852 x 10-12 F m-1. The ratio ε/ε0, a pure

183
number denoted by εr, is called the relative permittivity and indicates how much weaker the electric
field is due to a point charge in a material medium compared to what it would be if the charge were
placed in a vacuum.ii

H.2 Field of an oscillating magnetic dipole in homogeneous space

(a) Complete field: Consider the oscillating magnetic or current dipole of section 4.6.6.4 and, using
one of the conventions mentioned at the end of section 4.6.3.2, express its magnetic moment as
m0exp(-iωt) rather that as m0cos(ωt). If the dipole is situated in a homogeneous isotropic medium of
resistivity ρ, permittivity ε and magnetic permeability µ, the fields Br and Bθ are found to be

Br = (µm0 / 2πr3) (1-iκr)e-i(ωt-κr) cos θ (H.1a) ξ

Bθ = (µm0 / 4πr3) (1-iκr-κ2r2)e-i(ωt-κr) sin θ (H.1b) ξ

where

κ = (ω2εµ + iωµ/ρ)1/2 = a + ib (H.2) ξ

and

|κ| = (a2 + b2)1/2 = (ωµ / ρ)1/2 (ρ2ε2ω2 + 1)1/4 (H.3) ξ

The real and imaginary components, a and b, of κ are calculated as follows

Let

ω2εµ = αcosψ (H.4) ξ

ωµ /ρ = αsinψ (H.5) ξ

Then

κ2 = αεiψ

so that

κ = α1/2eiψ/2

= α1/2cos(ψ / 2) + iα1/2sin(ψ / 2)

therefore

a = α1/2cos(ψ / 2) b = α1/2sin(ψ / 2) (H.6) ξ

From (H.4) and (H.5)

α = (ωµ /ρ) (ρ2ε2ω2 + 1)1/2 (H.7) ξ

Now

2cos2(ψ / 2) = 1+ cosψ (H.8) ξ

2sin2(ψ / 2) = 1 - cosψ (H.9) ξ

184
Substituting for cosψ from (H.4) in (H.8) and (H.9) gives cos(ψ / 2) and sin(ψ / 2), after which the
equations (H.6) and (H.7) yield

a = (ωµ / 2ρ)1/2 [(ρ2ε2ω2 + 1)1/2 + ρεω]1/2

b = (ωµ / 2ρ)1/2 [(ρ2ε2ω2 + 1)1/2 - ρεω]1/2

(b) Quasi-static field: This is special case where the frequency is so low that the product ρεω is
much less than unity for the medium in question. For practically all rocks, with a few notable
exceptions such as water-logged soils, wet clays, etc., ε is on the order of 90 pF m-1. If, now, the
frequency ν is chosen so that ρν < ~108 Ω m s-1, then ρεω = 2πρεν ≈ 0.056 << 1. With the frequencies
(maximum about 20 Hz) commonly employed in geophysical work that exploits the phenomenon of
induction, the condition ρν < ~ 108 Ω m s-1 is satisfied with a good margin in many, if not most, areas.
(The statement about the ‘good margin’ does not apply to radar methods as these do not exploit the
phenomenon of induction as such). For a and b if ρεω << 1

a = b = (ωµ / 2ρ)1/2 (H.10) ξ

and the fields given by equations (H.1a) and (H.1b) reduce to

Br = (µm0/2πr3) (1 + br – ibr) e-i(ωt-br) e-br cosθ (H.11a) ξ

Bθ = (µm0/4πr3) [1 + br – i(br + 2b2r2)] e-i(ωt-br) e-br sinθ (H.11b) ξ

These fields are called quasi-static fields, since the frequency is so low that the permittivity of the
medium is not of any practical consequence, just as is the case with a purely steady field. In more
technical language, at any such frequency the displacement current density in the medium in
question is negligible compared to the conduction current density.ii

(c) Near field: Another concept that occurs in connection with time-dependent electromagnetic fields
in a medium can be seen from equations (H.1) in that at small distances r from the dipole such that
|κ|r → 0,

Br ≈ (µm0 / 2πr3) e-iωt cosθ (H.12a) ξ

Bθ ≈ (µm0 / 4πr3) e-iωt sinθ (H.12b) ξ

These expressions can be derived by means of the Biot-Savat law for the field of a magnetic dipole in
a medium of permeability µ. The real components of (H.12a) and (H.12b) are identical to equations
(4.54a) and (4.54b) respectively if µ = µ0. Thus, for a given medium and a given frequency, there is a
region (|κ|r ≈ 0) in the vecinity of the dipole where we can, with reasonable accuracy, express the
magnetic field by equations (H.12). This is the so-called near-field or induction region in which the
field amplitude varies as 1/r3 and the field is in phase with the dipole.
It should be noted that the quasi-static field expressions (H.11a) and (H.11b) hold irrespective of
the distance r, in contrast to the near-field case for which r must satisfy the condition |κ|r ≈ 0.
Equation (H.3) shows that for the quasi-static field we have |κ| = (ωµ/ρ)1/2. Hence, the near-field or
induction region of a quasi-static field is defined by the condition that (ωµ/ρ)1/2r must be a small
quantity. It is usual in this connection to consider (ωµ/ρ)1/2r as small if it is much less than one.
Although this is an arbitrary definition, it is useful for estimating roughly the distance from the source
out to which the primary field may be calculated from the Biot-Savat law.
For a non-magnetic medium (µ = µ0 = 4π x 10-7 Ω s m-1), the near-field condition becomes 2.81 x
10-3 (ν/ρ)1/2 r << 1, or r << 356 (ρ / ν)1/2, where ν is the frequency. As an example, if ν = 1000 Hz and
ρ = 100 Ωm, the Biot-Savat law is adequate for calculating the primary field of the source up to
distances substantially smaller than 112 m.ii

185
(d) Far field: For very large distances |κ|r → ∞ we find from equations (H.1) that Br decreases as 1/r2
and Bθ as 1/r so that when ρ is sufficiently large the total field is simply Bθ. Thus, in this so-called far-
field region or radiation region, the field of the dipole varies as the inverse distance from it, that is,
much more slowly than in the near-field region.
The field structure is more complicated in the intermediate region. In particular, the field is
elliptically ploarized in this region, whereas in the near and far regions it is linearly polarized; that is,
at any point the field vector oscillates along a straight line. There are however, no sharp boundaries
between the various regions.
The importance of distinguishing the near- and far-field zones is that in the interpretation of
electromagnetic measurements in the near-field zone it is mostly sufficient to apply the law of
induction while for the interpretation of measurements in the far-field zone, it is necessary to take in
account phenomena like the refraction and reflection of electromagnetic waves.

Appendix I
Logarithms and decibels

I.1 Natural logarithms (ln)

Table I.1 Ln (y) returns the natural logarithm of a number (y)

y ln (y) y ln (y) y ln (y) y ln (y) y ln (y)


1E-12 -27.6310 0.01 -4.6052 1 0 11 2.3979 20 2.9957
1E-11 -25.3284 0.02 -3.9120 2 0.6931 12 2.4849 21 3.0445
1E-10 -23.0259 0.03 -3.5066 3 1.0986 13 2.5649 22 3.0910
1E-09 -20.7233 0.04 -3.2189 4 1.3863 14 2.6391 23 3.1355
1E-08 -18.4207 0.05 -2.9957 5 1.6094 15 2.7081 24 3.1781
1E-07 -16.1181 0.06 -2.8134 6 1.7918 16 2.7726 25 3.2189
1E-06 -13.8155 0.07 -2.6593 7 1.9459 17 2.8332 26 3.2581
1E-05 -11.5129 0.08 -2.5257 8 2.0794 18 2.8904 27 3.2958
1E-04 -9.2103 0.09 -2.4079 9 2.1972 19 2.9444 28 3.3322
0.001 -6.9078 0.1 -2.3026 10 2.3026 20 2.9957 29 3.3673

30 3.4012 130 4.8675 230 5.4381 600 6.3969 7000 8.8537


40 3.6889 140 4.9416 240 5.4806 700 6.5511 8000 8.9872
50 3.9120 150 5.0106 250 5.5215 800 6.6846 9000 9.1050
60 4.0943 160 5.0752 260 5.5607 900 6.8024 1.E+04 9.2103
70 4.2485 170 5.1358 270 5.5984 1000 6.9078 1.E+05 11.5129
80 4.3820 180 5.1930 280 5.6348 2000 7.6009 1.E+06 13.8155
90 4.4998 190 5.2470 290 5.6699 3000 8.0064 1.E+07 16.1181
100 4.6051 200 5.2983 300 5.7038 4000 8.2940 1.E+08 18.4207
110 4.7005 210 5.3471 400 5.9915 5000 8.5172 1.E+09 20.7233
120 4.7875 220 5.3936 500 6.2146 6000 8.6995 1.E+10 23.0259

Table I.2 Powers of e, the base of the natural logarithm raised to the power x. [x = ln (ex)]

x ex x ex x ex x ex x ex
-12
10 1.000000000001 0.01 1.0101 1 2.718281828 11 5.9874E+04 30 1.0686E+13
-11
10 1.000000000010 0.02 1.0202 2 7.3891 12 1.6275E+05 40 2.3539E+17
-10
10 1.000000000100 0.03 1.0305 3 20.0855 13 4.4241E+05 50 5.1847E+21
-9
10 1.000000001000 0.04 1.0408 4 54.5982 14 1.2026E+06 60 1.1420E+26
-8
10 1.000000010000 0.05 1.0513 5 148.4132 15 3.2690E+06 70 2.5154E+30
-7
10 1.000000100000 0.06 1.0618 6 403.4288 16 8.8861E+06 80 5.5406E+34
-6
10 1.000001000001 0.07 1.0725 7 1096.6332 17 2.4155E+07 90 1.2204E+39
-5
10 1.000010000050 0.08 1.0833 8 2980.9580 18 6.5660E+07 100 2.6881E+43
-4
10 1.000100005000 0.09 1.0942 9 8103.0839 19 1.7848E+08 110 5.9210E+47
-3
10 1.001000500167 0.1 1.1052 10 22026.4658 20 4.8517E+08 120 1.3042E+52

186
I.2 Base 10 logarithms (log10)

B ase 10 logarithms utilize exponents of 10 ranging from about –12 to +12 without losing the
dynamics of the number range they represent (10-12 - 1012). Log10(x) is the power of 10 required
to raise it to the value x. For example log10(2) ≈ 0.30103 so therefore, setting 0.30103 as the power
of 10, 100.30103 = 2. An even simpler example is log10 (100) = 2, i.e., 102 = 100.

Table I.3 Base 10 logarithms

x log10 (x) x log10 (x) x log10 (x) x log10 (x) x log10 (x)
0.001 -3 0.01 -2 0.1 -1 1 0 10 1
0.002 -2.6990 0.02 -1.6990 0.2 -0.6990 2 0.3010 20 1.3010
0.003 -2.5229 0.03 -1.5229 0.3 -0.5229 3 0.4771 30 1.4771
0.004 -2.3979 0.04 -1.3979 0.4 -0.3979 4 0.6021 40 1.6021
0.005 -2.3010 0.05 -1.3010 0.5 -0.3010 5 0.6990 50 1.6990
0.006 -2.2218 0.06 -1.2218 0.6 -0.2218 6 0.7782 60 1.7782
0.007 -2.1549 0.07 -1.1549 0.7 -0.1549 7 0.8451 70 1.8451
0.008 -2.0969 0.08 -1.0969 0.8 -0.0969 8 0.9031 80 1.9031
0.009 -2.0458 0.09 -1.0458 0.9 -0.0458 9 0.9542 90 1.9542

1.001 0.0004 1.01 0.0043 1.1 0.0414 11 1.0414 101 2.0043


1.002 0.0009 1.02 0.0086 1.2 0.0792 12 1.0792 102 2.0086
1.003 0.0013 1.03 0.0128 1.3 0.1139 13 1.1139 103 2.0128
1.004 0.0017 1.04 0.0170 1.4 0.1461 14 1.1461 104 2.0170
1.005 0.0022 1.05 0.0212 1.5 0.1761 15 1.1761 105 2.0212
1.006 0.0026 1.06 0.0253 1.6 0.2041 16 1.2041 106 2.0253
1.007 0.0030 1.07 0.0294 1.7 0.2304 17 1.2304 107 2.0294
1.008 0.0035 1.08 0.0334 1.8 0.2553 18 1.2553 108 2.0334
1.009 0.0039 1.09 0.0374 1.9 0.2788 19 1.2788 109 2.0374

21 1.3222 41 1.6128 61 1.7853 81 1.9085 100 2


22 1.3424 42 1.6232 62 1.7924 82 1.9138 200 2.3010
23 1.3617 43 1.6335 63 1.7993 83 1.9191 300 2.4771
24 1.3802 44 1.6435 64 1.8062 84 1.9243 400 2.6021
25 1.3979 45 1.6532 65 1.8129 85 1.9294 500 2.6990
26 1.4150 46 1.6628 66 1.8195 86 1.9345 600 2.7782
27 1.4314 47 1.6721 67 1.8261 87 1.9395 700 2.8451
28 1.4472 48 1.6812 68 1.8325 88 1.9445 800 2.9031
29 1.4624 49 1.6902 69 1.8388 89 1.9494 900 2.9542
31 1.4914 51 1.7076 71 1.8513 91 1.9590 1000 3.0000
32 1.5051 52 1.7160 72 1.8573 92 1.9638 2000 3.3010
33 1.5185 53 1.7243 73 1.8633 93 1.9685 3000 3.4771
34 1.5315 54 1.7324 74 1.8692 94 1.9731 4000 3.6021
35 1.5441 55 1.7404 75 1.8751 95 1.9777 5000 3.6990
36 1.5563 56 1.7482 76 1.8808 96 1.9823 6000 3.7782
37 1.5682 57 1.7559 77 1.8865 97 1.9868 7000 3.8451
38 1.5798 58 1.7634 78 1.8921 98 1.9912 8000 3.9031
39 1.5911 59 1.7709 79 1.8976 99 1.9956 9000 3.9542

4 5 6 7 12
10 4 10 5 10 6.0000 10 7 10 12
4 5 6 7 12
2x10 4.3010 2x10 5.3010 2x10 6.3010 2x10 7.3010 2x10 12.3010
4 5 6 7 12
3x10 4.4771 3x10 5.4771 3x10 6.4771 3x10 7.4771 3x10 12.4771
4 5 6 7 12
4x10 4.6021 4x10 5.6021 4x10 6.6021 4x10 7.6020 4x10 12.6021
4 5 6 7 12
5x10 4.6990 5x10 5.6990 5x10 6.6990 5x10 7.6989 5x10 12.6990
4 5 6 7 12
6x10 4.7782 6x10 5.7782 6x10 6.7782 6x10 7.7781 6x10 12.7782
4 5 6 7 12
7x10 4.8451 7x10 5.8451 7x10 6.8451 7x10 7.8450 7x10 12.8451
4 5 6 7 12
8x10 4.9031 8x10 5.9031 8x10 6.9031 8x10 7.9030 8x10 12.9031
4 5 6 7 12
9x10 4.9542 9x10 5.9542 9x10 6.9542 9x10 7.9542 9x10 12.9542

187
I.3 Decibels (dB)

T he analog signals registered by, for example, the geophones in the seismic method or the
oscilloscope in the radar method are often measured in nanovolts, microvolts or millivolts and are
generally too weak to be converted directly into digital signals. Recording systems commonly use
preamplifiers in the analog to digital (A/D) conversion, that apply gains ranging typically from 0 dB to
48 dB to the input signal, sample by sample. After digital conversion, the signal is again amplified (or
‘gained’) to improve the signal to (instrument) noise ratio and to improve the A/D converter resolution.
The decibel (dB) is the unit of measurement of the gain or attenuation of an input signal. It is
sometimes used arbitrarily in sound systems to indicate volume levels but it is more precisely
described as the ratio of the amplitudes (volts) or powers (watts) between two signals. Sub-section
4.6.11.1 (f) discusses the decibel as it applies to the power ratios of radar signals. An example of the
measurement in decibels as a ratio of signal voltages is given below in Figure I.1. Here, a 16 bit IFP
(instantaneous floating point) A/D converter receives an analog input signal, preamplifies it, converts
it to digital and then applies an IFP gain after conversion. In the figure, the bar of 16 squares is a
binary word representing the mantissa of the signal strength. The left-most bit is the sign (±) of the
word (1 = positive) and the remaining 15 comprise the binary code (1 = on, 0 = off) for the voltage of
the IFP gained signal. Each of the 15 squares is a power of 2 mV ranging from 213 (8192) on the left
to 20 (1) on the right with 0.5 mV as the last bit. The ‘on’ bits (ones) in the mantissa are summed to
describe the signal voltage.
The bar of 4 squares describes the operator-selectable gain in decibels to apply to the digital
signal. Each of the squares is a power of 2 from 28 on the left to 21 on the right. The exponents of 2
in the ‘on’ squares are summed to give the gain value (e.g. 2(1+2) = 23 = 8). In this case, 18 dB, equal
to an 8-fold gain in signal strength was applied. Thus, a preamplified analog signal of 733.2125 mV
results in an analog value of 5865.69 mV after the IFPgain.

Input signal: +45.825781 mV

Preamp gain: 16 (24 dB)

Signal before IFP: +733.2125 mV


mV

mV

mV

mV

mV

mV
mV

mV

mV

mV

mV

mV

mV

mV

mV
±

8192

4096

2048

1024

512

256
128

64

32

16

1
sign

0.5

28

24

22

21

X X X X X X X X X X X X

Mantissa: 1010110111001101 Gain: 8 = 23

IFP gain: 8 (18 dB)

Output analog value: +5865.69 mV

Figure I.1 Instantaneous floating point (IFP) 16 bit A/D converter with a 24 dB preamp gain and
an 18 dB IFP final gain.

188
The decibel, as a measure of the ratio of two amplitudes, is given by

dB = 20 log10 (A2 / A1) (I.1a) ξ

where A2 is the comparison (output) amplitude and A1 is the reference (input) amplitude. If A2 > A1,
there is a gain or amplification and the dB is positive. If A2 < A1, that is A2 / A1 < 1, there is
attenuation and the dB is negative. Conversely, if the dB is known and set equal to 20 log10 (x), then
the amplitude ratio x (= A2 / A1) is given by

x = 10(dB/20) (I.1b) ξ

Example I.1

A reference signal with an amplitude of 256 mV is attenuated -48 dB. To calculate the amplitude of
the output signal, set x = (A2 / A1). Using equation (I.1b),

A2 / A1 = 10(-48/20) = 0.00398 or 0.398% attenuation.

The output voltage is 256 x 0.398% = 1.0192 mV

Some of the more commonly used actual decibel values are given in Table I.4. To obtain
intermediate values that may not appear in the table, recall one of the laws of exponents which states
that if y = (a + b) and xy = x(a + b), then xy = xa · xb. From this, find two values of x in Table I.4 whose
product equals the desired intermediate value of x and add their dB values.

Example I.2

If an input signal A1 to an amplifier has strength of 3.2 µV and is gained resulting in an output signal
A2 that measures 230.4 µV, the ratio of amplitudes is equal to 72. The value of x = 72 does not
appear in the table but the values for x of 8 and 9 do. Where x = 8, 20 log10 (x) = 18.0618 dB and for
x = 9, 20 log10 (x) = 19.0849 dB. Summing the two shows that a 72-fold gain is equal to 37.1467 dB.

Table I.4 Table of decibels (actual values)

20 log10 (x) 20 log10 (x) 20 log10 (x) 20 log10 (x)


x x x x
(dB) (dB) (dB) (dB)
-6
10 -120 1 0 15 23.5218 150 43.5218
10-5 -100 2 6.0206 20 26.0206 200 46.0206
-4
10 -80 3 9.5424 30 29.5424 300 49.5424
-3
10 -60 4 12.0412 40 32.0412 400 52.0412
0.0039063 -48.1654 5 13.9794 50 33.9794 500 53.9794
0.015625 -36.1236 6 15.5630 60 35.5630 600 55.5630
0.0625 -24.0824 7 16.9020 70 36.9020 700 56.9020
0.125 -18.0618 8 18.0618 80 38.0618 800 58.0618
0.25 -12.0412 9 19.0849 90 39.0848 900 59.0849
0.5 -6.0206 10 20 100 40 1000 60

1500 63.5218 15000 83.5218 150000 103.5218 1500000 123.5218


2000 66.0206 20000 86.0206 200000 106.0206 2000000 126.0206
3000 69.5424 30000 89.5424 300000 109.5424 3000000 129.5424
4000 72.0412 40000 92.0412 400000 112.0412 4000000 132.0412
5000 73.9794 50000 93.9794 500000 113.9794 5000000 133.9794
6000 75.5630 60000 95.5630 600000 115.5630 6000000 135.5630
7000 76.9020 70000 96.9020 700000 116.9020 7000000 136.9020
8000 78.0618 80000 98.0618 800000 118.0618 8000000 138.0618
9000 79.0849 90000 99.0849 900000 119.0849 9000000 139.0849
10000 80 100000 100 1000000 120 10000000 140

189
The decibel values commonly used in binary-gain recording appear in Table I.5. In many
recording instruments, gain selections are preset to increments of 6 dB and as such are associated
with fixed gain levels. For example, 6 dB, the ‘universal doubler’, actually computes to a gain of
1.995, while 60 dB, which appears in table 1.5 as 1024 is actually a 1000 fold gain. The reason for
this is that binary-based systems such as the 16-bit IFP one in Figure I.1 use powers of 2 as gain
multipliers. 60 dB, actually equal to 20 log10 (1000), then becomes equal to 1024 (210) in instrument
tables.
Table I.6 contains decibel values of attenuations or ‘inverse gains’. It is seen, for example, that an
attenuation of –18 dB reduces the signal by 12.5% which is the reciprocal of an 18 dB gain of factor 8
(8-1 = 0.125).

Table I.5 Decibel values for gains in binary systems

dB Gain dB Gain dB Gain dB Gain


0 1 36 64 72 4,096 108 262,144
6 2 42 128 78 8,192 114 524,288
12 4 48 256 84 16,384 120 1,048,576
18 8 54 512 90 32,768 126 2,097,152
24 16 60 1,024 96 65,536 132 4,194,304
30 32 66 2,048 102 131,072 138 8,388,608

Table I.6 Decibel values for attenuations in binary systems

Percent Percent Percent


dB Attenuation dB Attenuation dB Attenuation
% % %
0 1 100 -36 0.016 1.6 -1 0.891 90
-3 0.7071 70.7 -42 0.008 0.8 -2 0.794 80
-6 0.50 50 -48 0.004 0.4 -5 0.56 56
-9 .355 35.5 -54 0.002 0.2 -10 0.316 32
-12 .25 25 -60 0.001 0.1 -20 0.1 10
-15 .178 17.8 -66 0.0005 0.05 -40 0.01 1
-18 .125 12.5 -72 0.00025 0.025 -50 .00316 .32
-21 0.089 8.9 -78 0.000125 0.0125 -70 .000316 .032
-24 0 .063 6.3 -84 0.0000631 0.0063 -80 10-4 0.01
-27 0 .045 4.5 -90 0.0000316 0.0032 -100 10-5 0.001
-30 0.0316 3.2 -96 0.0000158 0.0016 -120 10-6 0.0001
-33 0.022 2.2 -102 0.00000794 0.0008 -140 10-7 0.00001

Appendix J
Tables of symbols, units and definitions

Table J.1 Exponential symbols

Exponent Decimal Symbol Prefix Exponent Decimal Symbol Prefix

1012 1,000,000,000,000 T tera - 10-1 0.1 d deci -


109 1,000,000,000 G giga - 10-2 0.01 c centi -
106 1,000,000 M mega - 10-3 0.001 m milli -
103 1,000 k kilo - 10-6 0.000 001 µ micro -
2
10 100 h hecta - 10-9 0.000 000 001 n nano -
101 10 da deca - 10-12 0.000 000 000 001 p pico -

190
Table J.2 Symbols and units for seismic methods

Sym Unit Meas. Name Definition Equivalence Comments


-2
(N m x newtons Derived from Hooke’s law, which states
κ -10 Bulk Modulus Incompressibility
10 ) per meter Vp = [(κ + that stress is proportional to strain where
4/3µ) / ρ]
-2 1/2
(N m x newtons the elastic constant is the constant of
µ -10 Shear Modulus Rigidity proportionality.
10 ) per meter
meters P-wave Velocity of Also called the longitudinal wave or the
Vp = (j / ρ)
-1 1/2
Vp (m s )
per sec. velocity compressional wave primary wave
meters S-wave Velocity of Also called the shear wave or the
Vs = (µ / ρ)
-1 1/2
Vs (m s )
per sec. velocity transverse wave secondary wave
kilograms mass of matter in a
ρ ρ = j / Vp
-3 2
(kg m ) Density
per cu. m medium
Elongational µ = q / 2(1+σ)
j = κ + (4µ) / 3
Reflection

j
elasticity κ = q / 3(1-2σ)
Young’s j = q(1-σ) /
q
modulus (1+σ)(1-2σ)
σ Poisson ratio Always less than 0.5
Depth of burial of
z (m) meters Depth
sedimentary rocks V = 46.5 P-wave velocity in shales and sands as
-1
Geologic age of (zT)1.6 m s a function of age and depositonal time
T (y) years Age
sedimentary rocks
Angle between P-
Angle of
i (º) degrees wave reflection and
incidence
normal incidence sin i / sin r =
Seismic methods

Snell’s law of refraction


Angle between V 1 / V2
Angle of
r (º) degres propagating wave
refraction
and normal incidence
-1
ic = sin (V1 / Angle at which refraction is parallel to
ic (º) degrees Critical angle Occurs when r = 90º
V2) the interface and propagation is nill
Length of one cycle
λ (m) meters Wavelength λ=V/ν Inverse of wavenumber K
of a seismic wave
Number of cycles
ν (Hz) Hertz Frequency that pass an arbitrary ν=V/λ Formally cycles per second
point in one second
-1 U = V – λ(dV /
U (m s ) m / sec Group velocity Constant velocity
dλ)
Logarithmic Coefficient of energy A = A0 exp(- Loss of energy through absorption in
δ
decrement absorption (.02-.03) δνx / V) rocks of a seismic wave of amplitude A0
Distance at which a
Minimum At lesser distance, there is no refraction
xm (m) meters refraction can be xm = 2h tan ic
distance recorded.
Reflection

recorded at surface
Distance at which
Crossover xc = 2h [(V2 +
xc (m) meters refraction overtakes 1/2
distance V1) / (V2-V1)]
the direct wave
2
Two-way travel time t = (2/V)(h +
tx (s) seconds Travel time 2 1/2 Time to travel the path SAG in Fig. 2.1
for a reflected wave x /4)
Normal incidence
h (m) meters Depth h = (t0V) / 2
depth to reflector
Two-way time from
Normal Also called vertical time or zero-offset
t0 (s) seconds surface to reflector t0 = (2h) / V
incidence time time
along the normal
Difference between
reflected time t and Defines NMO, normal moveout
∆t (s) seconds Change in time ∆t = t – t0
normal incidence NMO = Tx – T0
time t0
2 2
-1 Tx = T0 + (x / An empirical value that yields a
VNMO (m s ) m / sec NMO velocity Stacking velocity 2
VNMO) satisfactory relationship between Tx & T0
T (s) sec Period Seconds per cycle T=1/ν Reciprocal of frequency

191
Sym Unit Meas. Name Definition Equivalence Comments

K=1/λ
-1
K (m ) / meter Wavenumber Cycles per meter Reciprocal of wavelength
Calculated from V =
rms, n
V ∆tk) / For multi-layer stacking
-1 n 2
Vrms (m s ) m / sec RMS velocity borehole velocity [(Σ k=1 k
1/2
measurements t] n
2 2
V DIX = (Vb Tb
-1 2 Average speed of wavefront between
VDIX (m s ) m / sec Dix velocity Interval velocity - Va Ta / (Tb -
two points
Ta)
Reflection

Interval Velocity in one ∆d is difference between shot depths


Vint = ∆d / ∆t
-1
Vint (m s ) m / sec
velocity homogeneous layer and ∆t is difference of respective tuh
-1 Weathering Average velocity of
Vw (m s ) m / sec Weathering layer also called LVL
velocity wave in weathering
-1 V, Sub- Average velocity
Vsw (m s ) m / sec
weathering below LVL
ds (m) meters Depth of shot
Depth of
dw (m) meters
weathering
Time from shot to tuh = (ds / Vsw) -
tuh (msec) millisec. Uphole time
uphole geophone Wc
Weathering To correct for error in Wc = (ds / Vsw)
Wc (msec) millisec
correction time due to LVL - tuh
Seismic methods

Elevation
To correct for Ec,s = (Ed – Es) Ed and Es are elevations of datum and
Ec,s (m) meters correction,
elevation differences / Vsw shot respectively
shot
Elevation
To correct for Ec,r = (Ed – Er)
Ec,r (m) meters correction, Er is elevation of receiver
elevation differences / Vsw
receiver
Total static To correct for
Tc (ms) millisec.
correction weathering and elev
Acoustic Magnitude of ratio of Rc = (I2 – I1) /
I
impedance stress : point velocity (I2 + I1)
Displacement Rc = (ρ2V2 –
Reflection Where displacement amplitude of
Rc amplitude of ρ1V1) / (ρ2V2 +
coefficient incident wave is empirically 1.0
reflected wave ρ1V1)
Displacement
Transmission
Tc amplitude of Tc = 1 - Rc = 2I1 / (I2 + I1)
coefficient
transmitte wave
Surface projection of In refraction, the distance from a point
a (m) meters Distance a = h tan ic
N1A below the shot to point of ic on reflector
h = V1∆c / 2
∆c (ms) millisec Delay time See Fig. 3.7
Refraction

cos ic

θ (º) degrees Dip angle See Fig 3.8 To calculate two-way times and interval
ζ = x sin θ velocities for refraction along a dipping
Differencial refractor
ζ (m) meters See Fig 3.8
difference
Time Used in mean-minus-T method for
∆T (ms) msec See Fig 3.10
difference velocity estimates in refraction surveys

Table J.3 Symbols and units for radioactivity method

Sym Unit Meas. Name Definition Equivalence Comments

Joule,
(J), Measure of electrical The energy acquired by an electron
eV Kilowatt Electron volt
(kWh) energy falling through potential diff. of one volt
hour

192
Sym Unit Meas. Name Definition Equivalence Comments

1 eV = 1.602 x
J Joule -19
10 J
1 eV = 4.45 x
kWh Kilowatt hour -26
10 kWh
Particles of helium Large kinetic energy but slow rapidly
α alpha 4
nucleus 2He upon collision with other particles
Particles of electrons Lower atomic mass, higher velocity than
β beta
and positrons alpha particles
Electromagnetic energy = hc Also called quantum consisting of
γ gamma
radiation /λ discrete particles called photons
Planck’s h = 6.625 x
h -34
constant 10 J s
8
-1 Velocity of c = 3 x 10 m
c (m s ) m / sec -1
light s
λ = 10 to
-11
Wavelength of (0.1 – 0.01 Å) Energy of a photon is
Radioactivity methods

λ (m) meters Wavelength


1.24 x 10 / λ eV
-12 -6
gamma rays 10 m
Rate of decrease in
t1/2 = (ln 2) / λ mass decay: m = m0e λτ
-
t1/2 (y) years Half-life
mass of radioactivity
9
238 Primeval radioactive Half-life = 4.51 x 10 y For complete list
U Uranium-238
isotope of radioactive decay series, see pg. 58
Activity of number of
1 Bq = 1
(Bq) becquerel Radiation unit disintegration events
decay / sec
(decays per second)
Activity of 1 gram of 1 Ci = 3.7 x
(Ci) curie Radiation unit 10
radium 10 Bq
Quantity of x-
1 R = 8.38 x
radiation producing -6 -1
(R) roentgen Radiation unit 9 10 J kg (of
2.083 x 10 ion pairs
air)
/ cm3 of air
-1
Dose of ionizing 1 Gy = 1 J kg
(Gy) gray Radiation unit
radiation per kg = 100 rad
Old unit of energy -2
rad Radiation unit 1 rad = 10 Gy
dose
-1
1 Sv = 1 J kg
(Sv) sievert Radiation unit Equivalent dose
= 100 rem
Intensity of Emitted by a point C = constant, µ = absorption coefficient,
I = (C/ r )e µ
2 - r
I
light source at distance r = distance
grams Radiation The earth absorbs γ-rays so radiation
ρ
-3
(g cm )
per cu m density count depends upon density
2
Number of quanta N = (N0 / N0 = number of quanta emitted, r = (x +
N
4πr )µe µ z ) , µ = absorption coefficient
2 - r 2 1/2
reaching detector

Table J.4 Symbols and units for electrical methods

Sym Unit Meas. Name Definition Equivalence Comments

ampere Electrical
I (A) Electron flow I=U/R In short, thin linear conductor
or amp current
Voltage drop or
U (V) volt Potential U = IR Also, V, dV
potential difference
In short, thin linear conductor. Also as
Opposition to current I = - U / R where minus expresses
R (Ω) ohm Resistance R=U/I
flow direction of current flow from high to low
potential (opposite potential increase).

193
Sym Unit Meas. Name Definition Equivalence Comments

R is directly proportional to length L (dL)


L (m) meters Length Length of conductor R = ρ (L/ s) meters of the conductor and inversely
2
proportional to the cross-section s (m )
Area of conductor
s = ρ (L / R)
2
s (m ) sq. meter Cross-section
face
Unlike resistance which is characteristic
Constant of
ohm of a particular path of electric current,
ρ (Ωm) Resistivity proportionality,
meter resistivity is a physical property of a
volumetric resistance
material.
amps per
Current Current per unit area
Also, j = E / ρ and j = σE
-2
j (A m ) square j=I/s
density of cross-section
meter
m-1 volts per E=U/L Also (-dV / dL, In the direction of the
E (V ) Electric field
meter =ρj current density vector)
siemens
σ σ=1/ρ
-1 -1 -1
(S m ) Conductivity Inverse of resistivity Also mhos per meter (Ω m )
per meter
Ah = area of hemisphere. For potential
Radial distance of a at distance r from a point current
r = (2Ah / π)
1/2
r (m) meters Radius
hemispherical shell source, U = (Iρ / 2π) (1 / r). For two
electrodes, U = (Iρ / 2π) [(1 / r) – (1 / r’)]
E Activation energy also called energy gap
Electrical methods

Boltzmann’s -(E / κT) Measure of increasing conductivity of


κ σ = σ0e
constant semiconductors with increasing temp.
Absolute
T (K) º Kelvin
temperature
ohm of water filling pores n is usually close to 2.0 if more than
ρ0 (Ωm) Resistivity
meter in porous rocks about 30% of pore space is water-filled
Volume fraction of but can be greater. m depends upon
φ ρ = ρ0φ s
-m -n
Porosity
pores degree of cementation or geologic age
varies from 1.3 (loose Tertiary
Fraction of pore
s sediments) to about 1.95 (well cemented
space filled by water
Palaeozoic)
A capacitor initially charged to U0, dis-
C Capacitance charges to voltage (U0 / e) = τ (time
constant). The resistance is then τ / C.
L Inductance
ohm Apparent Inhomogeneous General equation for apparent resistivity
ρa (Ωm) ρa = K (UMN / I)
meter resistivity resistivity where K is the geometric constant
farads Dielectric Also referred to as
ε E = (ερζ / η) p Electric field along a capillary
-1
(F m )
per meter constant dielectric permittivity
Streaming Also called zeta potentials, it is the
ζ (V) volts
potentials potential across ends of a capillary tube
-1 pascal Pressure C = (ερζ / η) is the streaming potential
p (Pa m )
per meter gradient coefficient for earth material and ranges
pascal Dynamic from a fraction of a millivolt per
η (Pa s) -1
seconds viscosity kilopascal to about 1.3 mV kPa
Universal gas -1 -1
R = 8.314 J K mol
constant
-1
F Faraday’s constant = 96.487 C mol

Table J.5 Symbols and units for magnetic method

Sym Unit Meas. Name Definition Equivalence Comments

nano- Magnetic flux Measurement of magnetic field also


B (nT) Flux per unit area B = µH
tesla density called magnetic induction, a vector

194
Sym Unit Meas. Name Definition Equivalence Comments

amps per Magnetic field B per


H=B/µ
-1
H (A m ) Field strength
meter permeability µ
ohm sec Absolute Also a vector, it is the absolute
µ µ=B/H
-1
(Ω s m )
per meter permeability permeability of a medium.
Magnetic field strength at center of one-
-1 Unit field
(1 A m ) turn circular of 1 meter diameter,
strength
carrying current of 1 ampere
volt Unit of magnetic flux
(V s) Magnetic flux
second also called weber
-2 volt sec Unit of flux density B
(V s m ) Flux density -2
per sq m = (Wb m )
-2 weber Unit of flux density B
(Wb m ) Flux density
per sq m = (nT)
-9
nano- Unit of flux density 1 nT = 10 tesla. Exactly equal to the
(nT) Flux density -2
tesla equal to (Wb m ) older unit gamma (γ)
Field strength H will create in vacuum a
Absolute
ohm sec flux density B0. For most geophysical
µ0 = 4 π x 10 Ω s m B0 = µ0 H
-1 -7 -1
(Ω s m ) permeability of
per mete purposes, absolute permeability in air
vacuum
and most rocks = µ0.
A pure number. B = µH = µrµ0H = µ0H +
Magnetic methods

ohm sec Relative The ratio of two


µr µr = µ / µ0
-1
(Ω s m )
per m permeability permeabilities µ0(µr-1) H = µ0H + µ0kH
k Susceptibility Equal to µr - 1 µr = 1 + k A pure number. In vacuum µr = 1, k = 0
Intensity of Additional field strength present at
kH magnetization M points of space occupied by a medium
induced by H subject to field strength H
amps per Intensity of
B = µ0 (H + M)
-1
M (A m ) M = kH A vector
meter magnetizaton
Volume of a body in
V Volume VM = m
a magnetic field
2 amps per Magnetic A vector. Direction of magnetic moment
m (A m ) m = VkH
sq meter moment of dipole is the dipole axis.
Susceptibility
∆k
difference
Vertical magnetic field (flux dens) of
Bz (T) tesla
earth
Total change in vertical magnetic field in
∆Bz ∆Bz = ½ ∆kBz
traversing a contact
Field strength
Induced
associated with
T magnetization
earth’s flux = kT
density
Koenigsberger
Qn Mn / kT
ratio
Natural m is magnetic
remanent moment of a
VkT is induced moment, VMn is
Mn intensity of the m = VkT + VMn mass of
remanent mag. moment
earth’s field postulated
intensity geometry

Table J.6 Symbols and units for electromagnetic methods

Sym Unit Meas. Name Definition Equivalence Comments

Resultant-field
R0 = (P0 + S0 + 2P0S0 cos ψ)
2 2 1/2
R0
amplitude

195
Sym Unit Meas. Name Definition Equivalence Comments

Resultant-field
α α = tan [S0 sin ψ / (P0 + S0 cos ψ)]
-1
phase
Secondary-
S0 = (P0 + R0 + 2P0R0 cos α)
2 2 1/2
S0
field amplitude
Secondary-
ψ ψ = tan [R0 sin α / (R0 cos α - P0)]
-1
field phase
Magnetic-field
ν ν = ω / 2π
frequency
φ = tan (ω L /
-1
Secondary-
φ ψ = -(π / 2 + φ)
field lag R)
Response l = self inductance, R = resistance, ν =
τ t = 2πνl / R
parameter frequency
Attenuation
α α = ω{µε[(1 + σ /ω ε )
2 2 2 1/2 1/2
– 1] / 2}
constant
δ = 503.29
δ
Electromagnetic methods

(m) meters Skin depth 1/2


(ρ/ν) meters
Angular
ω (Hz) Hertz ω = 2πν
frequency
µ0= 4 π x 10
-7
ohm sec Permeability of
µ0
-1
(Ω s m )
Ωsm
-1
per mete vacuum
siemens
σ σ=1/ρ
-1 -1 -1
(S m ) Conductivity Inverse of resistivity Also mhos per meter (Ω m )
per meter
ohm Specific Inverse of
ρ (Ωm) ρ=1/σ
meter resistivity conductivity
ν (Hz) Hertz Frequency Inverse of period ν=1/T
T (s) sec Period Inverse of frequency T=1/ν
Induced RI + L(dI / dt) Induced current in loop when current is
I (A) amps [-(R / L) t]
current =0 switched off. I = I0
amps per Magnetic field Magnetic field B per
H=B/µ
-1
H (A m )
meter strength permeability µ
ohm sec Magnetic
µ µ=B/H
-1
(Ω s m )
per meter permeability
farads Dielectric Also referred to as Magnitude of force in newtons between
ε
-1
(F m )
per meter permittivity dielectric constant two massless point charges, Q1 and Q2
VLF methods

siemen Free-space Permittivity in ε0 = 8.852 x


ε0 Also Ω s m and F m
-1 -1 -1 -1
(S s m ) -12 -1
sec / m permittivity vacuum 10 F m
Relative
εr Ratio of permittivities εr = ε / ε0 A pure number
permittivity
Electric field
Etx Transmitted wave
strength | Etx / Hty |
= (µωρ)
1/2
Magnetic field
Hty Transmitted wave
strength
ohm Specific ρ = (1/2πµ0ν) The ratio Ex / Hy is called the wave
ρ (Ωm) 2
meter resistivity | Ex / Hy | impedance and has the unit of ‘ohm’
τ (s) sec. Time Duration of square- wave pulse
A(f) =
Amplitude Of a square-wave
A(f) τ [(sin(πfτ) / f is the component frequency
spectrum pulse
(πfτ)]
GPR methods

= 10 log10 (P1/ 2
Decibel Power ratio Also 10 log10 (A1 / A2)
P2)
dB
= 20 log10 (A1 /
Decibel Amplitude ratio
A2)
Quality Efficiency of EM
Q Q = ρεω Also = ωε / σ, where σ is conductivity
indicator wave propatation
Also = tan δ. If tan δ << 1, Q >> 1 and
1/Q Loss tangent Inverse Q 1 / Q = (σ / ωε)
the wave propagates more efficiently.

196
Sym Unit Meas. Name Definition Equivalence Comments

Angular
ω ω = 2πr
frequency
Attenuation
α See above EM methods
constant
farads Dielectric Relative permittivity
ε εwater is about 80
-1
(F m )
per meter permittivity used in GPR
Relative
µr magnetic
permeability
ohm sec Absolute
µ
-1
(Ω s m )
GPR methods

per meter permeability


ohm sec Free-space
µ0 µ0 = 4π x 10
-1 -7
(Ω s m )
per meter permeability
meters Velocity of EM
v = c / (εrµr)
-1 1/2
v (m s ) c = velocity of light
per sec Wave
Wavelength of
λ (m) meters λ = c / ν √εr if µ is taken as unity
radar wave
Complex
K K = ε + jσ / ω
permittivity
Z (W) ohms Impedance Z = ωµ / κ
Reflection Rc = (Z2 – Z1) /
Rc Also, Rc = (√ e1r - √ ε2r) / (√ e1r + √ ε2r)
coefficient (Z2 + Z1)
Transmission Tc = (2Z2) / (Z2 Z1 and Z2 are impedances of the two
Tc
coefficient + Z1) layers of media
Radar wave two-way t = (2/v) [h2 +
t (s) sec Two-way time 2 1/2 v = electromagnetic wave velocity
travel time (x / 2) ]
2 2
Depth of h = (v t –
h (m) meters 2 1/2
reflector x) /2
Fraction of total pores containing 3 Oil in place. A is area, h is thickness and
Sw Ah(1-Sw)φ m
water φ is porosity of an oil reservoir
[P(273 + t0) /
Gas volume. t0 is standard (surface)
P0 (Pa) Pascal Standard pressure (atmospheric) P0(273 + t)](1-
3 temperature
Sw)φAh m
-12
Well logging

2 = 0.987 x 10
(m ) darcy Permeability in oil reservoir 2 Length squared. Also common millidarcy
m
Rt ohm True resistivity of a formation
Rt = Rw / φ Sw
m n
(Ωm)
Rw meters Resistivity of water in pore spaces
Esh e.m.f. across shale/nonshale border Emw + Esh + UB
Kirchoff’s law applied to SP boundaries
Emw e.m.f. across water/mud interface – UA = 0
ρma kilograms Matrix density of a clean formation
ρb = φρf + (1 –
Also, φ = (ρma – ρb) / (ρma – ρf)
-3
ρf (kg m ) per cubic Density of fluid filling pores
meter φ)ρma
ρb Bulk density

Appendix K
Well velocity surveys

K.1 Checkshot geometry

F igure K.1 shows an exaggerated case of a source offset x from the well. Generally, this offset is
small compared to the depth of log however at shallow depths and longer offsets, the difference
between recorded travel time and the vertical transit time (tx) can be significant. Figure K.3 shows an
actual example where the x offset is on the average of 56 m and the shallowest level (500 ft) has a
shot-receiver angle (θ) of 83º resulting in a time correction of over 5 ms (Figure K.4).

197
s
x dist

tx
s – r dist

shot depth

Figure K.1 Source-receiver well velocity geometry

If x is the horizontal distance from the source to the well head, h the vertical logging depth to the
downhole geophone less charge depth and TT the recorded transit time of the direct wave, then the
corrected transit time tx (Figure K.1) can be calculated as follows:

tx = TT sin [tan-1(h / x)] (K.1) ξ

The resultant time is then added to tuh and corrected to seismic datum.

Example K.1

From Figure 5.17 (a), at depth level 1500 ft, the recorded transit time TT for stack number 40 was
214.08 ms. From Figure K.3, the offset distance x of SP 45 on the upshots is 52.66 m (172.8 ft).
Using equation (K.1) and a charge depth of 15 ft, the vertically corrected travel time tx = 214.08 x sin
[tan-1 (1485 / 172.8)] = 212.65 ms. This was a relatively shallow shot that resulted in a transit time
correction of 1.43 ms.

2m

10
11
30
31
50

2m
1
20
21
40
41

Figure K.2 Shot array to wellhead geometric relationship

The geometric relationship of the shot array to the wellhead is shown in Figure K.2. Here the

198
distance and azimuth from the center of the array to the well is measured. Then, starting with the
UTM coordinates of the well and working backward, the UTM coordinates are calculated for each of
the shotpoints in the array by computing each distance and azimuth from the array center point.
In this particular survey, one shot was taken at each depth level. The charges should be
sufficiently separated to avoid sympathetic detonations of nearby shotpoints. The separation
depends upon the charge size and the type of near surface geology. As will be seen in the Dynamite
check shot sequence log in Figure K.5, the separation for this survey was insufficient for the loose
and moist conditions in the weathering and resulted in several sympathetic detonations. When
planning a checkshot or VSP dynamite survey, provide extra shots to allow for initial testing, misfires,
double detonations, etc.
Figure K.3 is a previously referenced table of shotpoint coordinates with shot to receiver distances
and azimuths given. Note that, for the major portion of the survey, the shot to receiver angles were
nearly vertical (from 88º to 89º). The only deviations from near vertical occurred between 500 ft and
about 2000 ft (83º to 87º).

DYNAMITE CHECK SHOT

depth (ft) sp utm e utm n x-dist (m) x-azm s-r dist s-r angle
1500 47 782034.15225 1799098.10686 52.73539 118.0112 1468.9469 87.942632
5200 4 782024.37513 1799096.00727 60.72415 111.9182 5168.3567 89.326804
5125 5,16 782027.02874 1799096.98624 58.65852 113.7732 5093.3378 89.340126
5050 6 782026.04977 1799099.63985 60.65825 115.6946 5018.3666 89.307435
4975 7,14 782028.70338 1799100.61882 58.72668 117.6782 4943.3488 89.319314
4900 8,13 782029.54070 1799102.43511 58.86274 119.6228 4868.3559 89.307226
4800 9 782028.56173 1799105.08872 61.05263 121.3337 4768.3909 89.266387
4670 10,11 782031.21534 1799106.06769 59.33652 123.4741 4638.3795 89.267023
4600 12 782030.37802 1799104.25140 59.06626 121.5562 4568.3819 89.259182
4500 15 782027.86606 1799098.80253 58.65852 115.7268 4468.3850 89.247831
4400 17 782026.19142 1799095.16995 58.72667 111.8218 4368.3948 89.229719
4300 18 782025.35410 1799093.35366 58.86274 109.8772 4268.4059 89.209847
4200 19,20 782023.67946 1799089.72108 59.33652 106.0258 4168.4223 89.184380
4100 21 782025.49575 1799088.88377 57.36046 105.7228 4068.4044 89.192160
4000 22 782026.33307 1799090.70006 57.08084 107.7059 3968.4105 89.175840
3900 23 782027.17039 1799092.51635 56.87022 109.7061 3868.4181 89.157655
3810 24 782028.00771 1799094.33264 56.72938 111.7186 3778.4259 89.139727
3700 25 782028.84503 1799096.14893 56.65882 113.7387 3668.4376 89.115035
3600 26 782029.68235 1799097.96521 56.65882 115.7613 3568.4498 89.090236
3500 28 782031.35699 1799101.59779 56.87022 119.7939 3468.4663 89.060516
3430 29 782032.19431 1799103.41408 57.08084 121.7941 3398.4794 89.037615
3340 30 782033.03163 1799105.23037 57.36046 123.7772 3308.4973 89.006595
3200 31 782034.84792 1799104.39305 55.38612 124.1018 3168.4841 88.998400
3100 32 782034.01060 1799102.57676 55.09648 122.0492 3068.4947 88.971168
3000 33 782033.17328 1799100.76047 54.87825 119.9775 2968.5073 88.940723
2900 34 782032.33596 1799098.94418 54.73227 117.8921 2868.5222 88.906713
2800 35 782031.49864 1799097.12789 54.65914 115.7983 2768.5396 88.868739
2650 36 782030.66132 1799095.31160 54.65914 113.7017 2618.5705 88.803941
2500 37 782029.82400 1799093.49531 54.73227 111.6079 2468.6068 88.729573
2347 38 782028.98668 1799091.67902 54.87824 109.5225 2315.6504 88.642029
2050 39 782028.14936 1799089.86273 55.09647 107.4508 2018.7520 88.436070
1950 40 782027.31204 1799088.04644 55.38611 105.3982 1918.7995 88.345929
1925 41 782029.12833 1799087.20912 53.41368 105.0496 1893.7534 88.383747
1800 42 782029.96565 1799089.02541 53.11329 107.1767 1768.7976 88.279269
1700 43 782030.80297 1799090.84170 52.88687 109.3251 1668.8382 88.183945

199
1600 44 782031.64029 1799092.65799 52.73539 111.4888 1568.8866 88.073739
1500 45 782032.47761 1799094.47428 52.65948 113.6619 1468.9442 87.945591
500 46 782033.31493 1799096.29057 52.65948 115.8381 470.9533 83.580067

Figure K.3 Source-receiver geometry for dynamite check shot

K.2 Average and interval velocities

F igure K.4 shows a table of adjusted transit times and resulting interval and average velocities
based on the shot-receiver geometry of Figure K.3 and the tx correction of equation (K.1). The
values should generally agree with the Delta-T curve in the sonic log. Anomalous interval velocities
may be the result of borehole inconsistencies, improper geometry or the failure of first break detection
or uphole time.

Stack Depth Offset x Tuh TT Tx Tx + Tuh Int Tint Vint Vavg


SP -1 -1
No. (ft) (ft) (ms) (ms) (ms) (ms) (ft) (ms) (f s ) (f s )

41 46 500 172.72 8.28 96.01 90.75 99.03 500 99.03 5049 5049
40 45 1500 172.72 8.22 214.08 212.67 220.89 1000 121.93 8202 7053
39 44 1600 172.97 8.17 243.75 242.34 250.51 100 29.66 3371 6602
38 43 1700 173.47 10.33 250.62 249.33 259.66 100 6.99 14312 6818
37 42 1800 174.21 10.12 262.49 261.27 271.39 100 11.94 8373 6889
36 41 1925 175.20 11.33 273.94 272.81 284.14 125 11.54 10829 7056
35 40 1950 181.67 10.33 275.67 274.48 284.81 25 1.67 14979 7104
34 39 2050 180.72 10.13 281.87 280.78 290.91 100 6.30 15874 7301
33 38 2347 180.00 8.89 300.11 299.23 308.12 297 18.45 16097 7843
32 37 2500 179.52 8.40 308.60 307.81 316.21 153 8.58 17840 8122
31 36 2650 179.28 7.43 319.57 318.84 326.27 150 11.03 13595 8311
30 35 2800 179.28 8.39 327.61 326.94 335.33 150 8.10 18520 8564
29 34 2900 179.52 8.58 336.42 335.78 344.36 100 8.84 11316 8637
28 33 3000 180.00 8.58 343.01 342.39 350.97 100 6.62 15113 8762
27 32 3100 180.72 6.56 352.01 351.41 357.97 100 9.02 11088 8822
26 31 3200 181.67 7.57 357.43 356.86 364.43 100 5.44 18376 8967
25 30 3340 188.14 6.60 368.40 367.82 374.42 140 10.96 12772 9081
24 29 3430 187.23 7.66 373.72 373.16 380.82 90 5.35 16830 9192
23 28 3500 186.53 7.50 381.42 380.88 388.38 70 7.71 9073 9189
22 26 3600 185.84 8.41 386.59 386.08 394.49 100 5.20 19244 9325
21 24 3700 185.84 8.52 391.48 390.99 399.51 100 4.91 20362 9463
20 25 3810 186.07 8.36 403.53 403.05 411.41 110 12.06 9119 9453
19 23 3900 186.53 9.63 408.35 407.88 417.51 90 4.83 18618 9562
18 22 4000 187.23 9.44 426.83 426.36 435.80 100 18.48 5411 9382
17 21 4100 188.14 8.75 431.25 430.80 439.55 100 4.43 22556 9517
16 20 4200 194.62 8.63 437.60 437.13 445.76 100 6.33 15787 9608
15 18 4300 193.07 8.50 444.50 444.05 452.55 100 6.92 14447 9684
14 17 4400 192.62 8.09 447.91 447.48 455.57 100 3.43 29165 9833
13 15 4500 192.40 8.41 455.59 455.17 463.58 100 7.69 12999 9886
12 12 4600 193.74 7.36 458.20 457.79 465.15 100 2.62 38168 10048
11 10 4670 194.62 8.56 462.57 462.17 470.73 70 4.37 16001 10105

200
10 9 4800 200.25 8.53 471.20 470.79 479.32 130 8.62 15078 10196
9 8 4900 193.07 8.40 475.72 475.35 483.75 100 4.56 21927 10308
8 7 4975 192.62 7.70 479.79 479.43 487.13 75 4.08 18384 10377
7 6 5050 198.96 7.33 491.31 490.93 498.26 75 11.50 6523 10287
6 5 5125 192.40 7.68 494.96 494.61 502.29 75 3.68 20367 10362
5 4 5200 199.18 8.42 499.26 498.89 507.31 75 4.28 17513 10423

Figure K.4 Table of adjusted transit times, interval and average velocities

For stack number 16 at depth 4200, there was no uphole time recorded so an average was taken
of the two nearby shots. The computed interval velocities generally agreed with the sonic log. The
low interval velocity between 3900 and 4000 feet was attributed to a low velocity stringer within a
limestone matrix. The anomalous interval velocity between 4500 and 4600 feet was due to improper
first break detection. Figure K.5, below, shows an example of a checkshot sequence log.

DYNAMITE CHECK SHOT SEQUENCE 31 JAN 1999

SEQ DEPTH Tuh TT


SP COMMENTS
NO (ft) (ms) (ms)
1 50 500 0.00 0.00 NO SIGNAL 29 JAN 99 07:01
2 49 1500 0.00 0.00 NO T.B. 31 JAN 99 03:25
3 48 2500 0.00 0.00 NO T.B. 04:14
4 1 5200 0.00 0.00 NO T.B., NO Z SIGNAL 05:19
5 2 4900 0.00 0.00 NO T.B., NO Z SIGNAL 05:28
6 3 4900 0.00 0.00 8.5 Hz INTERNAL NOISE
7 47 1500 8.49 219.00 NEW CSAT TOOL OK 06:47
8 4 5200 8.42 219.38 07:32
9 16 & 5 5125 7.68 494.96 DOUBLE DET 07:37
10 6 5050 7.33 491.31 07:44
11 14 & 7 4975 7.70 479.79 DOUBLE DET 07:48
12 13 & 8 4900 8.40 475.72 DOUBLE DET 07:53
13 9 4800 8.53 471.20 07:58
14 11 & 10 4670 8.56 462.57 DOUBLE DET 08:04
15 12 4600 7.36 458.20 08:08
16 15 4500 8.41 455.59 08:13
17 17 4400 8.09 447.91 08:19
18 18 4300 8.50 444.50 08:25
19 19 & 20 4200 0.00 437.60 NO UH BREAK 08:29
20 21 4100 8.75 431.25 08:36
21 22 4000 9.44 426.83 08:42
22 23 3900 9.63 408.35 08:48
23 25 3810 8.36 403.53 08:54
24 24 3700 8.52 391.48 08:59
25 26 3600 8.41 386.59 09:05
26 28 3500 7.50 381.42 27 MISSFIRED 09:17
27 29 3430 7.66 373.72 09:21
28 30 3340 6.60 368.40 09:26
29 31 3200 7.57 357.43 09:32
30 32 3100 6.56 352.01 09:36

201
31 33 3000 8.58 343.01 09:42
32 34 2900 8.58 336.42 09:49
33 35 2800 8.39 327.61 09:54
34 36 2650 7.43 319.57 10:01
35 37 2500 8.40 308.60 60 Hz NOISE, TURNED OFF GENERATORS
36 38 2347 8.89 300.11 10:19
37 39 2050 10.13 281.87 10:28
38 40 1950 10.33 275.67 10:33
39 41 1925 11.33 273.94 10:37
40 42 1800 10.12 262.49 CEMENTED CASING 10:44
41 43 1700 10.33 250.62 CEMENTED CASING 10:48
42 44 1600 8.17 243.75 CEMENTED CASING 10:54
43 45 1500 8.22 214.08 CEMENTED CASING 10:59
44 46 500 8.28 96.01 CASING ONLY 11:13

Figure K.5 Checkshot sequence log

K.3 VSP frequency analysis

F requency analyses are offered in both checkshot and VSP surveys but are more often requested
in the latter. Figure K.6 is a frequency analysis plot of a dynamite shot taken at depth 2500 ft.
Being relatively shallow, the frequency content is good with amplitudes of –18 dB (relative to the peak
frequency of about 18 Hz) over the range of 6 Hz to about 125 Hz. It comes as no surprise that at
greater depths the frequency bandwidth is increasingly subdued with the higher frequencies greatly
attenuated.

Figure K.6 Frequency spectrum of a single dynamite shot taken at 2500 ft.

202
Figure K.7 shows a frequency spectrum plot of a shot at 5050 ft during the same survey. Here, the
highest frequency obtained at –18 dB is about 65 Hz and a total bandwidth of only about 80 Hz.
Figure K.8 is a plot of the frequency attenuation with depth effect of the above survey with curves at –
36 dB and –30 dB.

Figure K.7 Frequency spectrum plot of shot taken at 5050 ft.

HIGH FREQUENCY ATTENUATION WITH DEPTH

260

-36 dB
240
freq (-30 dB)

220
freq (-36 dB)
Linear (freq (-30 dB))
200 Linear (freq (-36 dB))
FREQUENCY (Hz)

180

-30 dB
160

140

120

100

80

60
1200 1700 2200 2700 3200 3700 4200 4700 5200
DEPTH (FT)

Figure K.8 Frequency attenuation with depth at -36 and -30 dB

203
Indices to Figures, Tables, Examples and Equations

Index of Figures

actual seismic section and synthetic seismogram Figure 5.11 151


analog refraction record Figure 3.3 (a) 37
apparent conductivity map of archaeological survey in
Copan, Honduras Figure 4.52 112
apparent rates of change in conductivity with increasing
radar frequencies for some common rock types Figure 4.62 130
Archie's law variation of bulk resistivity Figure D.2 173
bad coupling to borehole wall Figure 5.19 (b) 163
basic time-distance graph Figure 1.6 11
broadening of a seismic pulse Figure 1.5 10
broadside tilt-angle technique Figure 4.48 107
buried point electrode Figure 4.8 63
calculating geophone static correction from the uphole time Figure 2.12 23
calculation of low induction number apparent resistivity for a
layered earth Figure 4.46 105
cancellation of headwave by a nine detector geophone array Figure 3.1 33
chargeability profile across Pyramide ore body Figure 4.27 84
checkshot QC record Figure 5.15 157
checkshot sequence log Figure K.5 202
checkshot summary listing of depth-time pairs Figure 5.17 (a) 158
CMP gather of 6 traces Figure 2.6 18
coil orientations in various moving source-receiver systems Figure 4.49 110
combination of alternating vertical and horizontal magnetic
field vectors Figure 4.57 121
combination of magnetic dipoles to form a magnet with
moment 2Lm Figure 4.31 91
comparison of different logs with their relation to core lithology Figure 5.1 137
conductor striking parallel to the propagation direction of the
VLF wave (E polarization) Figure 4.58 123
conductor striking perpendicular to the propagation direction
of the VLF ground wave (H polarization) Figure 4.59 124
construction of a normal-incidence time section over a
curved reflector Figure 2.15 27
corridor stack, two-way time Figure 5.18 (h) 162
critical refractions Figure 3.2 35
cross-line and in-line bistatic GPR configurations for
transmitter (Tx,) and receiver (Rx,) Figure 4.63 132
current flow patterns for uniform half-space and two-layer ground Figure 4.15 73
current-carrying coil as a dipole Figure 4.42 102
definition of NMO Figure 2.4 (a) 16
density and neutron logs with crossover Figure 5.9 149
density determinations by radioactivity Figure 4.5 59
depth versus transit-time plot Figure 5.17 (b) 159
derivation of expression relating tx to x with dipping reflector Figure 2.16 28
determination of radar pulse velocity in ground Figure 4.65 134
diagram of an photomultiplier tube (electron multiplier) Figure 4.4 56
diffraction hyperbola Figure 2.17 29
diffraction of a radar wave by a buried culvert Figure 4.64 133
diffraction of an edge Figure 1.4 9

204
diffusion of ions producing IP voltage Figure 4.24 82
dipping interface Figure 3.8 44
displacement of MN segments in Schlumberger sounding Figure 4.13 72
distribution of resistivities around a borehole Figure 5.2 138
downgoing VSP events velocity filtered Figure 5.18 (d) 161
downgoing VSP events velocity filtered with waveshape decon Figure 5.18 (f) 161
downgoing VSP signal without a velocity filter Figure 5.18 (b) 160
effect of coil position on anomaly detection Figure 4.54 115
electric dipole Figure D.3 174
electrolytic concentration cell Figure 4.20 78
electron and ion flow in and around an ore body Figure 4.22 79
elevation corrections Figure 2.13 24
exponential law with parameters for radioactive decay and
radio wave attenuation Figure 4.1 51
field record showing signal and noise of different frequencies Figure F.2 178
Fourier amplitude spectrum of a square wave pulse Figure 4.61 127
Fourier transform products: (cosine, sine, amplitude and phase) Figure F.5 181
frequency attenuation with depth at –36 dB and –30 dB Figure K.8 203
frequency spectrum of a single dynamite shot taken at 2500 ft Figure K.6 202
frequency spectrum plot of shot taken at 5050 ft Figure K.7 203
frequency-domain IP survey contour plot in a dipole-dipole
configuration Figure 4.28 85
Fresnel zones Figure 2.18 30
general phase diagram: primary, secondary and resultant field Figure 4.36 96
GPR pulse, idealized asymmetric ‘square’ GPR pulse, power
spectrum of pulse Figure 4.60 126
ground response to a square wave signal and to a spike impulse Figure 4.25 83
high-frequency bursts Figure 5.19 (c) 164
hysteresis loop Figure 4.32 92
identification of wave time arrivals Figure 3.3 (b) 37
IFP 16 bit A/D converter Figure I.1 188
illustration of Schlumberger sounding with resistivity graph Figure 4.12 71
instantaneous direction of rate of flux increase with time Figure 4.34 95
integrated sonic log with checkshot corrections Figure 5.14 156
IP measurement in the time domain Figure 4.26 83
long-term SP stability over the same 1.5 km track Figure 4.17 75
low-frequency model Figure 5.12 151
magnetic dipole field Figure 4.33 95
magnetic fields for horizontal cable layouts Figure 4.41 101
mapping fracture zone using mean-minus-T method Figure 3.11 47
matrix correction to φa limestone chart Figure 5.8 148
mean-minus-T method for estimating velocities in refractors Figure 3.10 47
migration required in GPR measurements Figure 4.66 135
moving source-receiver system configuration Figure 4.50 111
n layer example of hyperbolic raypaths Figure 2.9 (b) 20
n layer example of normal raypaths Figure 2.9 (a) 20
natural gamma-ray spectrum Figure 4.3 54
NMO corrected CMP gather Figure 2.7(c) 19
nomogram relating resistivity, frequency and skin depth Figure 4.47 106
operators log for a Wenner electric sounding in Copan, Honduras Figure 4.10 69
origin of negative induced polarization Figure 4.30 88
origin of the electromagnetic anomaly in the moving
source-receiver system Figure 4.53 114

205
permeable bed invaded uniformly Figure 5.6 145
PFE due to EM coupling as a function of overburden thickness
in frequency domain IP for a dipole-dipole configuration Figure 4.29 87
phase relations for a single-turn loop Figure 4.38 98
plotting of dipole-dipole ra data Figure E.1 176
point electrodes on the surface of a homogeneous earth Figure 4.7 62
potential drops in SP observations Figure 4.23 80
probes for geophysical well logging Figure 5.3 139
quadrature relationship Figure 4.35 96
ranges of P-wave velocities and rippabilities in common rocks Figure 1.1 5
ray path Figure 2.5 17
reflected and refracted rays in a multi-layer medium Figure 2.8 19
reflection Figure 1.3 (a) 8
reflection hyperbola Figure 2.1 (b) 13
reflection path Figure 2.1 (a) 13
reflections of a spike from various reflectors Figure 2.20 31
refraction Figure 1.3 (b) 8
refraction equation Figure 3.5 39
relationship of Vs to NMO Figure 2.4 (b) 16
relationships involved in the power spectrum Figure F.6 182
resistivity and induction log Figure 5.7 146
resistivity of a parallelpiped Figure 4.6 61
resistivity of pore water with dissolved salts Figure D.1 172
response of a horizontal-loop EM system to a vertical-loop
target as a function of the response parameter (t) Figure 4.40 100
reverberation Figure A.3 169
schematic of electrofiltration SP anomalies (profiles and surface
maps) on some flow models Figure 4.19 77
secondary-field response of a single-turn loop to a sinusoidal field Figure 4.39 98
seismic signal with a 22 Hz dominant frequency and a pure
22 Hz cosine wave Figure F.4 180
seismic time section showing air wave, ground roll, direct arrivals,
refractions and reflection hyperbola Figure 2.2 14
shot array-to-wellhead geometric relationship Figure K.2 198
signal to noise ratio to low Figure 5.19 (d) 164
simplified electrical circuit to explain SP jump across shale /
non-shale boundary Figure 5.5 143
single and multi-layer refraction time-distance graph Figure 3.7 42
single-turn coil in an alternating magnetic field Figure 4.37 97
six trace gather after both static and NMO correction Figure 2.14 (c) 25
six trace gather after static correction Figure 2.14 (b) 25
six trace gather before static and NMO correction Figure 2.14 (a) 25
sixty hertz electrical noise Figure 5.19 (a) 163
slingram EM-31 in an archaeological survey in Copan, Honduras Figure 4.51 111
Snell right-triangle Figure 3.6 41
Snell's law Figure 1.2 7
some common electrode arrays Figure 4.9 67
sonic log sonde Figure 5.10 150
sonic log with gamma-ray, caliper and borehole diameter curves Figure 5.16 158
source generated noise: air wave, ground roll Figure A.1 (b) 168
source generated noise: diffraction, refraction and multiples Figure A.1 (a) 168
source-receiver geometry for dynamite checkshot Figure K.3 200
source-receiver well velocity geometry Figure K.1 198

206
SP and γ-ray logs Figure 5.4 143
SP due to ion adsorption on pegmatite dikes Figure 4.21 79
SP profile across a sulphide body with hydrological influence of
ground-water flow Figure 4.16 74
spacing and depth penetration Figure 4.44 104
spike pulse Figure 2.19 31
spike series Figure A.4 169
stacked trace Figure 2.7(d) 19
stacked VSP data with bandpass filter and true amplitude recovery Figure 5.18 (a) 160
superposition of a series of reflection coefficients on an original
source pulse Figure 2.21 32
t - x graph representing the actual time section Figure 2.3 (b) 15
table of adjusted transit times, interval and average velocities Figure K.4 201
time picks Figure 3.3 (c) 37
time section showing both sides of the shot Figure 2.3 (a) 15
time-depth graph from well velocity survey (check-shot) with Vavg
and Vint Figure 2.10 21
time-depth plot Figure 2.11 (b) 22
time-distance graph for complex near-surface geology Figure 3.9 45
time-distance graph of relative velocities of noise Figure A.2 168
time-distance plot from refraction survey Figure 3.4 38
time-domain signal and its equivalence in the frequency domain Figure F.1 177
transient-field (TEM) measurement layouts Figure 4.55 117
tube wave Figure 5.19 (e) 164
two VES methods Figure 4.11 70
U-238 radioactive decay series Figure 4.2 53
uncorrected CMP gather Figure 2.7(a) 18
upgoing aligned VSP signal without velocity filtering Figure 5.18 (c) 161
upgoing VSP events filtered Figure 5.18 (e) 161
upgoing VSP events, velocity filtered and waveshape decon
applied Figure 5.18 (g) 162
uphole survey Figure 2.11 (a) 22
using sub-bases in an SP survey Figure 4.18 76
variation of induced current with depth in homogeneous ground for
co-planar coil systems operating at low induction numbers Figure 4.45 105
variation of skin depth, d, with frequency and resistivity Figure 4.43 103
various dipole-dipole configurations Figure 4.14 72
viscous mud spotted near the surface Figure 5.20 165
VLF field at a distant point Figure 4.56 120
weighted sine and cosine waves for two harmonic frequencies in
a Fourier series Figure F.3 179
well velocity survey, vibrator source Figure 5.13 155

Index of Tables and Examples

base 10 logarithms Table I.3 187


decibel values for attenuations in binary systems Table I.6 190
decibel values for gains in binary systems Table I.5 190
decibels (actual values) Table I.4 189
elastic constants Table 1.1 3
elastic velocities in m/s Table 1.2 6
electrical resistivities of some rocks and minerals Table 4.4 64

207
exponential symbols Table J.1 190
geophysical well log objective cross reference Table C.2 172
logging methods and geophysical objectives Table C.1 171
magnetic susceptibilities of common rocks and ores Table 4.5 94
natural logarithms ln (y) Table I.1 186
powers of e, base of natural logarithm Table I.2 186
radiation units of measurement Table 4.3 57
radioactive contents of some natural materials Table 4.2 55
radioactive decay series 238Th Table 4.1 (c) 53
radioactive decay series 238U Table 4.1 (b) 52
radioactive decay series 40K Table 4.1 (a) 52
rock density Table B.1 170
symbols and units for electrical methods Table J.4 193
symbols and units for electromagnetic methods Table J.6 195
symbols and units for magnetic method Table J.5 194
symbols and units for radioactivity method Table J.3 192
symbols and units for seismic methods Table J.2 191
types of seismic noise Table A.1 167
typical reflection coefficients Rc Table 2.1 26
typical values of radar parameters for some common materials Table 4.7 129
values of Qn for some rock specimens Table G.1 183
world-wide VLF radio transmitters Table 4.6 119

calculating amplitudes of weighted sine and cosine waves Example F.1 179
calculating output voltage of a signal with given dB Example I.1 189
calculating the radius of area of reflection Example 2.2 31
calculating tn given depth, velocity and group interval Example 2.1 18
corrected vertical travel time Example K.1 198
decibel calculation by interpolation of table values Example I.2 189
depth penetration with current dipole Example 4.1 104
estimating critical angle, minium distance, direct arrival, refraction
and reflection times Example 3.1 35
finding slope and intercept of refraction line Example 3.4 40
locating time arrival time anomalies in complex t-d graph Example 3.5 46
picking arrival times Example 3.2 38
skin depth determination from nomogram and equation for δ Example 4.2 106
time averaging Example 1.1 6
verification of refraction times Example 3.3 40

Index of Equations

acoustic impedance ξ B.2 171


amplitude lag ξ F.7 180
amplitude ratio x, given dB ξ I.1b 189
amplitude spectrum of a square-wave pulse ξ 127
apparent chargeability, Ma ξ 4.32 84
apparent resistivity for a dipole in a parallel array ξ D.3d 175
apparent resistivity for a dipole in a perpendicular array ξ D.3e 175
apparent resistivity for a dipole in a radial array ξ D.3c 175
apparent resistivity for a dipole in an axial array ξ D.3b 175
apparent resistivity for a dipole in an azimuthal array ξ D.3a 175

208
apparent resistivity for a dipole in an equatorial array ξ D.3f 175
apparent resistivity in a pole-dipole array ξ 4.21 68
apparent resistivity in a pole-pole array ξ 4.22 68
apparent resistivity in a Schlumberger array ξ 4.19a 67
apparent resistivity in a Schlumberger array with small M-N ξ 4.19b 68
apparent resistivity in a Wenner array ξ 4.18 66
apparent resistivity in an axial-dipole array ξ 4.20 68
Archie’s law adaptation for pore-space water fraction ξ 5.4 142
Archie’s law for true resistivity ξ 5.3 142
Archie's equation ξ 4.13 65
B-field in induced magnetization ξ 4.39, 4.39a 90
Biot-Savat law ξ 4.49 101
change in time (∆t) ξ 2.7, 2.8 15
complete field of an oscillating magnetic dipole in homogeneous
space ξ H.1, H.2, H.3 184
components of elliptical polarization ξ 4.47a, 4.47b 99
conductivity in semi-conductors ξ 4.12 65
conductivity, σ [f (E, j )] ξ 4.7 61
convolution of a with b = convolution of b with a ξ 32
convolution to pth term in the final series ξ A.1 169
corrected transit time for well velocity survey ξ K.1 198
cosine transform ξ F.3 178
critical angle ξ 1.7 9
critical angle with dip ξ 3.16 44
crossover or critical distance xc ξ 3.6 41
current density, j [f (E, ρ)] ξ 4.6 61
d.c. shift in frequency domain ξ F.2 178
decibel as a measure of ratio of two amplitudes ξ I.1a 189
depth calculation in second refraction layer ξ 3.10 43
diffracted ttwo with coincident source-receiver ξ 2.29 29
diffraction hyperbola ξ 2.28 29
dip angle θ ξ 3.15 44
dipole magnetic moment ξ 4.40 91
dipole potential in uniform half-space ξ D.2 175
direct arrival time ξ 3.3 35
distance a with respect to critical angle ic ξ 3.1 35
distance from reflector to GPR antenna ξ 4.71b 133
Dix formula for calculating Vinterval from VRMS ξ 2.15 22
down-dip refraction time ξ 3.14 44
effect of coil separation on field strength ξ 4.60a, 4.60 b 113
effect of dip on NMO velocity ξ 2.27 27
effect of dip on total travel time ξ 2.26 27
elastic constants ξ 1.3 5
elastic P-wave velocity [f ( j, ρ)] ξ 5
elastic S-wave velocity [f (µ, ρ)] ξ 5
electric field in flowing electrolyte (streaming potential) ξ 4.23 76
electric field strength Etx of a transmitted VLF wave ξ 4.63 122
electrode distance relation to separation ξ 4.15b 66
electromotive force in electrolyte solution ξ 4.24 77
elevation correction ξ 2.21, 2.22 24
exponential law of attenuation of a penetrating wave ξ 4.55 103
field acting on receiver in moving-source EM systems ξ 4.59 113
fold ξ 18

209
Fourier series approximation of a time-domain function ξ F.1 178
Fresnel zone radius of the area of reflection ξ 2.30 31
gas volume in a reservoir ξ 5.2 141
general equation of an ellipse ξ 4.48 99
group velocity - constant velocity relationship ξ 1.8 10
horizontal refraction time tAB ξ 40
hyperbolic equation for two-way time (t2) ξ 2.4 14
hyperbolic equation rewritten for t ξ 2.6 15
imaginary component of R and S fields ξ 4.46 97
impedance (Z) of any medium to electromagnetic waves ξ 4.67 130
induced current in a transient field of a single-turn circular loop ξ 4.61 116
intensity I at a distance r from a point source ξ 4.1 58
intensity of magnetization, M [f (κ, H)] ξ 4.38 90
IP frequency effect ξ 4.33 85
K factor in electrode arrays ξ 4.17 66
Kirchoff’s law applied to self-potential ξ 5.5 143
Kirchoff's law in electric theory ξ 4.28 80
Koenigsberger ratio Qn = MnκT ξ 182
limit of Fourier series ξ F.5 179
liquid-junction potential ξ 4.26 78
Maclaurin's (infinite) series ξ 2.13 20
magnetic field in direction of radius vector ξ 4.54a, 4.54b 102
magnetic field in horizontal cable ξ 4.50 101
magnetic field in horizontal rectangular loop ξ 4.52 102
magnetic field in infinitely long cable ξ 4.51 101
magnetic field strength Htx of a transmitted VLF wave ξ 4.64 122
magnetic flux density (magnetic induction, B) [f (µ, H)] ξ 4.36 89
magnetic flux density as function of susceptibility ξ 4.37 90
magnetic moment in general ξ G.1 182
magnetic moment when T and Mn are parallel ξ G.2 183
magnetic parallel field in current carrying coil as a dipole ξ 4.53a 102
magnetic perpendicular field in current carrying coil as a dipole ξ 4.53b 102
maximum reflection stretch ξ 2.16 15
mean velocity ξ 3.20 47
mean-minus-T refraction time estimate ξ 3.19 46
metal factor as function of conductivity, σ ξ 4.35 86
metal factor as function of PFE ξ 4.34 86
minimum refraction distance ξ 3.2 35
near field, small distance r ξ H.12 185
NMO ξ 2.9 16
NMO estimation from RMS velocities ξ 2.14 21
normal incidence time ξ 2.3 14
Ohm's law ξ 4.3 60
oil in place in a reservoir ξ 5.1 141
parabolic approximation ξ 2.5 14
parabolic approximation for ∆t ξ 2.10 16
phase lag ξ F.8 180
polariziblilty in inhomogeneous ground ξ 4.31 83
porosity from formation resistivity ξ 5.9 146
porosity of formation from Wylie relation ξ 5.11 150
potential at buried electrode ξ 4.11 63
potential at ground surface ξ 4.10 63

210
potential at point electrodes ξ 4.9 62
potential difference between four electrode array ξ 4.16 66
potential drop across an hemispherical shell ξ 4.8 62
potential drop within an ore in static equilibrium ξ 4.29 80
potential of a dipole on homogeneous ground ξ D.1 174
potential of copper electrode in contact with electrolyte ξ 4.25 78
quanta detection, N ξ 4.2 59
quasi-static field for a and b when ρεω << 1 ξ H.10 185
quasi-static field of an oscillating magnetic dipole, low frequency,
no permittivity ξ H.11 185
ratio of amplitudes Etx and Htx as relates to (µωρ) ξ 4.65 122
ray path time [f (V, h, x)] ξ 2.1 13
real and imaginary components of a magnetic dipole ξ H.4 - H.9 184
real component of resultant field ξ 4.45 97
reflection and transmission coefficients of GPR waves ξ 4.68 130
reflection coefficient ξ 2.24 26
reflection coefficient of GPR wave [f (ω,ε,σ)] ξ 4.69 131
reflection coefficient of GPR wave in non-conducting media ξ 4.70 131
reflector depth [f (V, t, x)] ξ 2.2 13
refraction angle ξ 1.6 8
resistance, R [f (ρ, L, s)] ξ 4.4 61
resistivity at ground electrode array ξ 4.15a 66
resistivity in homogeneous ground ξ 4.14 65
resistivity log ξ 5.8 145
resistivity of a parallelpiped ξ 4.5 61
resultant-field amplitude EM frequency domain ξ 4.41 96
resultant-field phase EM frequency domain ξ 4.42 96
RMS velocity, 2 layer ξ 2.12 20
rock velocity [f (z, T)] ξ 1.4 6
second critical distance xc2 ξ 3.11 43
secondary-field amplitude EM frequency domain ξ 4.43 97
secondary-field phase EM frequency domain ξ 4.44 97
sine transform ξ F.4 179
skin depth as function of resistivity and frequency ξ 4.56 103
skin depth estimate [f (ρεω)] ξ 4.58 104
skin depths with large displacement currents ξ 4.57a 104
skin depths with representative permittivity ξ 4.57b 104
slope and intercept of refraction line ξ 3.4 40
Snell's law ξ 1.5 8, 34
Snell's law applied to multiple layers ξ 3.8 43
SP [f(t,Rmf, Rw)] ξ 5.7 144
SP jump ξ 5.6 143
spacing between GPR scans relative to antenna spacing (u) ξ 132
stacking velocity - NMO relationship ξ 2.11 17
static equilibrium ξ 4.27 80
thickness of top refraction layer h1 ξ 3.7 42
time equation for third straight-line segment in multiple-layer
time-distance curve ξ 3.9 43
total ohmic drop in electroyte ξ 4.30 81
total refraction time T, [f (cos ic)] ξ 3.5 41
total refraction time T, [f (x, h, V)] ξ 3.4 40
total refraction time with dip ξ 3.12 44
total static correction ξ 2.23 24

211
transmission coefficient ξ 2.25 26
two way refraction vertical component time tm ξ 40
two-way time of a GPR reflection ξ 4.71a 133
tx to x relationship with dipping reflector ξ 28
up-dip refraction time ξ 44
uphole time [f (d, V)] ξ 2.17 23
uphole time [f (d, V, Wc)] ξ 2.18 23
velocity of a P-wave [f (κ, µ, ρ)] ξ 1.1 4
velocity of an S-wave [f (µ, ρ)] ξ 1.2 4
velocity V2 [f (cos θ)] ξ 3.17 45
velocity V2 [f (Vu, Vd)] ξ 3.18 45
VLF field in air ξ 4.62 120
VLF measurement of resistivity as function of the Etx:Htx ratio ξ 4.66 124
volume of magnetic mass ξ G.3 183
wave motion with velocity V = (κ/ρ)1/2 ξ B.1 170
weathering correction [f (d, V)] ξ 2.19 23
weathering correction [f (d, V, tuh)] ξ 2.20 23
weighted cosine wave shifted along the time axis ξ F.6 179
Wylie relation (total tavel time t for wave to cross a block of unit
thickness) ξ 5.10 150

212
References

Baule, H. (1953) Geophysical Prospecting, 1, 111-24.


Bertin, J. and Loeb, J. (1976) Experimental and Theoretical Aspects of Induced Polarization, vols 1
and 2, Gebrüder Borntraeger, Berlin.
Coffeen, J.A. (1978) Seismic Exploration Fundamentals, Penn Well, Tulsa.
Datta, S. (1968) Absorption of dilatational waves in rocks, Geoexploration, 6, 127-39.
Dewan, W. (1983) Essentials of Modern Open-Hole Log Interpretation, Penn Well, Tulsa.
Dolan, W. (1970) Geophysical detection of deeply buried sulfide bodies in weathered regions,
Mining and Groundwater Geophysics 1967.
Faust, L.Y. (1951) Seismic velocity as a function of depth and geologic time, Geophysics, 16, 192- .
Gay, S.P. (1967) An 1800 millivolt self-potential anomaly near Hualgayoc, Peru, Geophysical
Prospecting, 15, 236-45.
Hughes, D.S. and Cross, J.H. (1951) Elastic wave velocities in rocks at high pressures and
temperatures, Geophysics 16, 577-93 .
Kirkpatrick, R.L. (1999) Dynamite Check Shot Survey, Basic Resources, Seismic Division.
Mayne, H. (1962) Common reflection point horizontal data stacking techniques, Geophysics, 27,
927-39.
Milsom, John (1996) Seismic Methods: General Considerations, Field Geophysics. (ref i)
Parasnis, D.S. (1997) Seismic Methods, Principles of Applied Geophysics. (ref ii)
Ricker, N. (1953) The form and laws of propagation of seismic wavelets, Geophysics, 18, 10-40.
Schlumberger (1999) Well Seismic Field User File.
Sheriff, R.E. (1973) Encyclopedic Dictionary of Exploration Geophysics, Society of Exploration
Geophysicists, Tulsa.
Sheriff, R.E. and Geldart, L.P. (1982) Exploration Seismology, vols 1 and 2, Cambridge University
Press, Cambridge.
Shumway, G. (1956) A resonant chamber method for sound velocity and attenuation measurements
in sediments, Geophysics, 21, 305-319.
Université Paris 7 (2000) Tercera Escuela de Geofísica Aplicada a la Arqueología y al Medio
Ambiente, Honduras.
Volgelsang, Dieter Seismic Methods, Environmental Geophysics: A Practical Guide. (ref iii)
Ginzburg, I.I (1960) Principals of Geochemical Prospecting, Pergamon Press, New York.
Hallof, P. (1967) IP Newsletter Cases xxvii and xviii, McPhar Geophysics, Canada.
Harrison, C.H. (1970) Reconstruction of subglacial relief from radio echosounding records,
Geophysics, 35, 1099-1115.
Hartshorn, L. (1925) A critical résumé of recent work on dielectrics, The Institute of Electrical
Engineering, 64, 1152.
Hawkes, H.E. and Webb, J.S. (1962) Geochemistry in Mineral Exploration, Harper, New York.
Hedström, H. (1937) Phase measurements in electrical prospecting, American Institute of Mining
and Metallurigical Engineering, Technical Publication 827.
Hohmann, G.W. (1973) Electromagnetic coupling between grounded wires at the surface of a two-
layered earth, Geophysics, 38, 854-63.
Jackson, P.D. (1981) Focused electrical resistivity surveys, Geophysical Prospecting, 29, 601-26.

213
Jämtlid, A., et al. (1984) Electrical borehole measurements for the mapping of fracture zones in
crystalline rock, Geoexploration, 31, 203-16.
Kayal, J.R. (1979) Electrical and gamma-ray logging in Gondwana and Tertiary coal fields,
Geoexploration, 17, 243-58.
Ketola, M. (1972) Some points of view concerning mise-à-la-masse measurements, Geoexploration,
10, 1-21.
Kilty, K.T. (1984) On the origin and interpretation of self-potential anomalies, Geophysical
Prospecting, 32, 51-62.
Kristiansson, K. and Malmqvist, L. (1982) Evidence for non-diffusive transport of 86Radon-222 in
the ground and a new physical model for the transport, Geophysics, 47, 1444-52.
Nayak, P.R. (1981) Electromechanical potential in surveys for sulphides, Geoexploration, 18, 311-.
Ogilvy, A.A., Ayed, M.A. and Bogoslovsky, V.A. (1969) Geophysical studies of water leakages
from reservoirs, Geophysical Prospecting, 17, 36-62.
Parasnis, D.S. (1967) Three-dimensional electric mise-à-la-masse survey of an irregular lead-zinc-
copper deposit in Central Sweden, Geophysical Prospecting, 15, 407-37.
Parasnis, D.S. and Soonawala, N.M. (ed) (1984) Nuclear fuel waste management, Geoexploration,
3/4, Special Issue.
Pedersen, L.B., et al. (1994) An airborne tensor VLF system. From concept to realization,
Geophysical Prospecting, 42, 863-83.
Pooley, J. Ph and van Steveninck, J. (1970) Geothermal prospecting – delineation of shallow salt-
domes and surface faults by temperature measurements at a depth of approximately 2 metres,
Geophysical Prospecting, 18, 666-700.
Rathor, B.S. (1977) Transient electromagnetic field over a two-layer ploarizable earth,
Geoexploration, 15, 137-49.
Sato, M. and Mooney, H.M. (1960) The electrochemical mechanism of sulphide self-potentials,
Geophysics, 25, 226-49.
Seigel, H.O., et al. (1968) Discovery and case history of the Pyramide ore bodies, Pine Point,
Northwest Territories, Canada, Geophysics, 33, 645-56.
Stacey, F.D. and Bandrjee, S.K. (1974) The Physical Principles of Rock Magnetism, Amsterdam
Sumner J.S. (1976) Principles of Induced Polarization for Geophysical Exploration, Amsterdam
Sundberg, K. (1931) Principles of the Swedish geo-electrical methods, Gerlands Beitr. Geophys.,
Erg. Hft., 1, 298-361.

214
Index

a α particle monitors 58 apparent metal factor (MF)a 86


a α, β and γ radiation 56 apparent polarizability 82, 83
absolute electrical permittivity 103 apparent resistivities for dipole-dipole 174
absolute magnetic permeability 103 apparent resistivity (ρa) 65, 66, 124, 138
absolute permeability (µ) 89, 90, 129 apparent velocity 167
absolute permeability of a vacuum (µ0) 90 Archie’s law (equation) 65, 142, 173
absolute permittivity 131 artificial neutron source 137
absolute temperature (K) 78 artificial radioactive tracers 49
absorption 8, 11 attenuation (Sec 1.7) 10-11
absorption coefficients 58 attenuation constant (α) 51, 103
absorption of EM waves 103 autocorrelation 182
absorption of α, β and γ rays 55 average / interval velocities 200, 201
acoustic impedance 26, 150, 170, 171 average velocity 7, 20-22, 23, 31, 155
acoustic log 140 axial configuration 72
activation energy 65 axial dipole 175
adsorption of SP 78 axial dipole-dipole array 68
air wave 14, 33, 37 – 38, 167 azimuthal configuration 72
airgun 166
alpha (α), beta (β) and gamma (γ) rays 50 back - e.m.f. 82
alpha particle 50 background radiation 58
alphacard monitor 59 base 10 logarithms 187
alpha-track method 59 base e 186
alternating vertical and horizontal magnetic becquerel 57
field vectors 121 bedrock 34, 46
ambient noise 167 best-fit hyperbola 18
amperes or amps (A) 60 best-fit lines 38
amperes per meter (A m-1) 89 Beta particles 50
amplitude 180 binary gain/attenuation 190
amplitude of magnetic field 113 Biot-Savart law 100, 101
amplitude ratios (dB) 128 bistatic antenna arrangement 131
amplitude spectrum 181 blind zone 43
analog refraction record 37 body waves 3-4, 10 (see also P-,S-waves)
analog to digital (A/D) conversion 188 Boltzmann's constant 65
angle of incidence (i or θi) 8-9, 11, 13, 26,43 borehole casing ringing 163
angle of refraction (r or θr) 8, 43 borehole compensated 140
angular frequency (ω) 96, 103 borehole diameter 141
anisotropy 7, 69 borehole magnetometer 153
antiferromagnetism 92, 93 break point 41, 46
apparent chargeability 84 broadside technique 107
apparent conductivities 86, 146 bulk modulus (κ) 4-5
apparent conductivity map 112 bulk resistivity 173

215
buried current electrode 62 continuous wave (CW) 95, 107
buried point electrode 63 continuous-velocity logging 150
convolution (Sec 2.13) 31-32, 169
cable noise 163 coplanar Tx and Rx coils 110
calculation of low induction number apparent corkscrew rule 95
resistivity for a layered earth 105 corridor stack 162
caliper log 152, 158, 171 cosine transform 178, 181
caliper tool (CAL) 141 coupling factor 100
capacitance critical angle (ic ) 8, 34
capacitance (C) 81, 173, 174 critical distance 33, 41
capacitor 129 critical refraction 34, 35
capillary flow 76 cross-line antenna configuration 131
carrier waves 119 crossover 148-149
casing arrivals 163 crossover distance 41
Cäsium-137 137 cross-planar Tx and Rx coils 110
CDP (see CMP) CRP (see CMP)
center frequency 126 curie 57
center shots 34 Curie temperature 90, 93
chargeability 84 Curie-Weiss law 93
chargeability profile 84 current density (j) 61, 63
charged-body potential 153 current flow patterns 73
check shot 7, 21, 155 current per unit area 91
checkshot geometry 197 - 199 current-carrying coil as a dipole 102
checkshot QC 156
checkshot summary 158 d.c. shift 178
clay percentage 140 darcy 142
CMP gather 18 daughter nucleus 50
CMP gather NMO corrected 19 decay constant 51
CMP stacking (Sec 2.5) 18 decay series 51
coaxial Tx and Rx coils 110 decibel (dB) 188 – 190, 128
coherent noise 17, 167 delay times 42
coil orientations 110 density (ρ) 2, 4-5, 26
common electrode arrays 67 density log (D) 137, 146, 171
common mid-point (CMP) 16 , 132 density-neutron crossover 149
compensated density log 147 depositional (detrital) remanent magnet-
compensated neutron log 148 ization (DRM) 94
compensator method 108 depth of penetration 73, 103
complex plane 97 depth versus transit time 159
complex refraction 45 deviation tool 141, 171
compressional wave, see P-wave 3 diamagnetism 90, 92
compressive (or tensile) stress, see stress 3 dielectric constant 76, 129, 152
Compton effect 55,137 dielectric permittivity 174, 183
Compton scattering 54, 147 dielectric permittivity, definition 129
concentration differences in electrolytes 77 differential spectrometer 57
conductive mineralization 125 diffraction GPR 132
conductive mud 143 diffraction (Sec 1.5) 9, 167
conductivity (σ) 61, 64, 88, 103, 140 diffraction hyperbola (Sec 2.12) 28-29, 132
conductivity rates of change with increasing radar diffraction of a radar wave 133
frequencies 130 diffusion 77
conductor detection by an H-field anomaly 122 diffusion current 82
conductor responses 113 dip 27-28
contact resistances 64 dip of a magnetic field 99
continuous profiling 132 dip-angle technique 107

216
dipmeter log 152 electrofiltration 76
dipmeter probe 141 electrofiltration SP anomalies schematic 77
dipole-dipole configurations 72 electrolytic concentration cell 78
dipole-dipole mapping 176 electromagnetic coupling 87
dipole-dipole sounding 72 electromagnetic induction 87, 94
dipping interface 44 electromagnetic methods (EM) 88
direct arrivals 14, 33, 37 electromagnetic radiation 50
direct ray slope 44, 45 electromagnetic response 100
direct wave 37, 38, 167 electromagnetic symbols and units 195
direct wave velocity 34 electromagnetic waves 140
disintegration of radioactive elements 51 electron multiplier 56
dispersion 10 electrons and positrons 50
displacement currents 104 electron-volts (eV) 49, 50
Dix formula 22 elevation corrections 24
dolomitization 148 elliptical polarization 99, 120, 121
donor or acceptor impurities 65 ellipticity 123
double-ended walk away 33 elongational elasticity 5
down-dip ray 44 EM phase 95
down-dip velocity 44 EM propagation travel-time 151
dynamic correction 16 EPT log 151
dynamite energy source 165 equatorial dipole configuration 72, 175
dynodes 56 equipotential lines 63
equivalent cavity 10
E polarization 122, 123 equivalent content 55
e.m.f. 78 equivalent resistivity 144
earth resistivity 60 exothermal chemical reaction 140
eddy currents 105, 121, 140 exponential law of radioactive decay 51
effects of coil separation 113, 115 exponential symbols 190
E-field anomaly and apparent resistivity 123
elastic constant 3, 5 far shots 33
elastic moduli 3, 5 Faraday’s constant (F) 78
elastic wave velocity (V) 5 far-field 119, 186
elastic waves 2 fast Fourier transform (FFT) 182
electric charge flux (current) 89 ferrimagnetism 90, 92, 93
electric dipole 119, 174 ferromagnetism 90, 92, 93
electric field (E) 61, 119 field in direction of radius vector 102
electric field effects 121 field of a current dipole 102
electric field moment 119 field of horizontal cable on the ground 101
electric field strength (Etx) 122 field of horizontal rectangular loop 102
electric log (EL, ES) 138, 171 field of oscillating magnetic dipole 184
electric mapping 73 field perpendicular to radius vector 102
electric moment 174 field records 36
electric sounding 70 field strength (H) 89, 113
electrical methods 60 filtering 177, 178
electrical noise 163 first arrivals 36, 43, 46
electrical symbols and units 193 first breaks 33
electrical vertical section 86 fixed-source systems 108
electrically polarized rock masses 81 flowmeter (FLOW) 140, 171
electrochemical mechanisms 76 - 79 flux density, B = µH (Wb m-2) 119
electrochemical potential 140 focused electric log, laterolog (FEL, LL)
electrode array geometric factors 66 140, 171
electrode polarization 82 focused induction log 140

217
focused microlaterolog 139 ground roll 3, 4, 14, 33, 37-38, 167
fold 18 ground velocities 134
formation coupling 163 group velocity 10
formation resistivity 138 guided waves 4
four electrode array 66
Fourier amplitude spectrum of a square-wave H polarization 124
pulse 127 half-life, 51
Fourier analysis 86, 176 - 182 harmonic frequencies 179
Fourier transform 117, 177 harmonic index 178
Fourier’s theorem 10 harmonic-field methods 107
fractional porosity 65 headwave 11, 33-34
free space (vacuum) field 100 hidden layer 43
free-space magnetic permeability (µ0) 119 high frequency bursts 164
free-space permittivity (ε0) 119, 120 high-impedance voltmeter 64
frequency analysis 202 - 203 highline noise 167
frequency domain 177 Hooke’s Law 3, 170
frequency domain IP, 81 horizontal coils 113
frequency domain IP measurement 85 horizontal-loop systems 111, 114
frequency effect 85 Huygens’ principle 9, 30
frequency spectrum 4 hydrogen content 137
frequency-domain IP survey contour plot 85 hydrogen nuclei 147
Fresnel zone 30 hydrostatic pressure 142
hyperbolic equation 14
gamma (1γ = 10-5 gauss) 89 hysteresis 91
gamma photons 54 hysteresis loop 91, 92
gamma radiation 136,137
gamma ray 50 – 59, 136 igneous and crystalline rocks 6
gamma ray scintillation counters 49 imaginary component 95, 96, 97
gamma-ray log (GR) 136, 171 impedance 95
gamma-ray logging 144 incident electric field 120
gas volume 141 incident magnetic field strength 120
Geiger-Müller counters 56 incompressibility 4 (see also bulk modulus)
geochemical 154 induced current in single-turn loop 116
geometric attenuation of EM waves 103 induced currents 103
geometric divergence 11 induced currents VLF, 122
Geonics EM-31 111 induced magnetism 92
Geonics EM-34-3 113 induced polarization 81
geophone array 33 inductance 174
geophone static correction 23 induction log (IEL, IES) 140, 146, 171
geophysical well log objective 172 induction number 105
geothermal 154 induction region 185
geothermal gradient 140 inhomogeneity 69
geothermal SP anomalies 75 in-line antenna configuration 131
GPR – reflection seismics similarities 125 in-phase component 95
GPR antennae 126 instantaneous floating point (IFP) 188
GPR distance determination 133 instantaneous velocity 150
GPR pulse, power spectrum 126 instrument noise 167
GPR symbols and units 197 integral spectrometer 57
GPR time window 127 integrated sonic log 156
gradient array 68 integrated sonic time 156
ground penetrating radar (GPR) 125 intensity factors 58
ground response to a square wave signal 83 interstitial water 64

218
interval velocity 21-22, 155 metal factor 86
ion flow in and around an ore body 79 mho per meter 62
ionic conductivity 64 microlog, (ML) 139, 171
ionic current density 79 microlaterolog, (MLL) 139, 171
isothermal remanent magnetization (IRM) 94 migration in GPR 135
isotopes 50 millivolt second per volt (mV s V-1) 84
isotropic 62 minimum distance 34, 38
mise-à-la-masse 153
joule 49, 57, 192, 193 mixed resistivities 138
monoelectrode log (PR) 138
K-40 decay series 52 monostatic antenna arrangement 131
kinetic energies 50 moving source-receiver systems 109, 111
kinetic potential 140 mud logging 136
multi-electrode array 69
laterolog 145 multiple point arrays 138
law of induction 94 multiple-level refractions 42
law of radioactivity and half-life 50 multiples 167
lines of force 102 mutual inductance 95, 100
liquid-junction potential 78
litho-density tool 147 nanotesla 89
lithologic column 136, 137(Fig 5.1), 151 natural gamma ray spectrum 54
logging methods 136, 171 natural logarithms 186
longitudinal wave see P-wave 3 natural remanent magnetization (NRM) 94
long-term SP stability 75 near field 185
loss tangent 128 near shots 33
Love waves 3 near-field artificial source CW methods 107
low velocity layer (see LVL) negative induced polarization 88
low-frequency model 151, Fig. 5.14 neutron log (N) 137, 147, 171
LVL 22 noise 17, 167
nomogram: specific resistivity, conductivity,
Maclaurin’s (infinite) series 20 frequency, period ,skin depth, wavelength
magnetic dipole 91, 119 106
magnetic dipole field 94, 95 non-parallel interfaces 43
magnetic field 119 normal incidence 8, 26-27, 29, 31
magnetic field effects 120 normal moveout (NMO) (Sec 2.3) 15-16
magnetic field strength Hty 122 normal-incidence ray 8, 27
magnetic field strength, H (A m-1) 89, 119 normal-incidence time 14, 19
magnetic flux 89 normal-incidence time section 27, 135
magnetic flux density (B) 89 nucleons 49
magnetic induction 89 Nyquist frequency 181
magnetic moment (m) 91
magnetic moment of current-carrying loop 99 ohm meters (Ωm) 61
magnetic permeability (µ) 87 Ohm’s law 60
magnetic remanence 182 ohmic drop 80
magnetic susceptibilities of common rocks 94 oil in place 141
magnetic symbols and units 194 origin of IP 81
magnetization (M) 90 origin of negative induced polarization 88
mass number 49 origin of the electromagnetic anomaly in the
matrix correction 148 moving source-receiver system 114
matrix density 146 orthogonal Tx and Rx coils 110
mean-minus-T 46 oscillating electric dipole 119
membrane polarization 81 oscillating transmitting dipole 114

219
oscillatory stress 34 power reflection coefficient 131
out-of-phase component 95 power spectrum 182
overburden 34 practical depth penetration 104
overvoltage 81 pressure wave, see P-wave 3
oxidation/reduction 140 primary field (P0) 96
oxidation-reduction currents 75 primary porosity 142
primary wave, see P-wave 3
parallel configuration 72 primeval isotopes 51
parallel dipole 175 principal refractors 34
paramagnetism 90, 92, 93 product divided by mean 45
percent frequency effect (PFE) 85, 87 propagation of a VLF ground wave 120
permeability 129, 137, 140 pseudo-section 176
permeability–zone logs 143 pseudo-section 86
permeable zones 142 pulse duration (τ) 127
permittivity (ε) 129 pulse width 126
permittivity measurements 129 pulse-height analyzer 57
permittivity of free space 183 push wave, see P-wave 3
permittivity, value for most earth materials 104 P-wave 3, 33
perpendicular array 72 P-wave velocity 34
perpendicular dipole 175
PFE due to electromagnetic coupling 87 Q (quality ratio) 128
phase difference (φ) 86 Q value 183
phase lag 180 quadrature 95
phase lag spectrum 181 quadrature relationship 96
phase relations for a single-turn loop 98 quanta 50
phase shift, IP 86 quantum theory 50
phase velocity 10, 26 Quasi-static EM fields 185
photoelectric effect 55
photomultiplier tubes 56 radar parameters, common materials 129
photons 50 radar pulse velocity in ground 134
Planck’s constant 50 radar wavelength 129
point diffractions 135 radial dipole 175
point electrode 62 radiation detection and measurement 56
Poisson ratio (σ) 5 radiation log 137
polarizability 84 radiation region 186
polarization voltage (Up) 83, 84 radiation threshold 57
pole-dipole (three electrode) and pole-pole radiation units of measurement 57
(two electrode) arrays 68 radioactive absorption 51
pore space 65 radioactive contents, natural materials 55
pore water 60 radioactive decay 50
pore-water resistivity 172, 173 radioactive density determinations 59
porosity 6, 65, 140 radioactive equilibrium 54
porosity log 146 radioactive geometrical attenuation 51
positrons 54 radioactive symbols and units 192
potassium isotope 40K 136 radioactivity 50
potassium-40 51 radioactivity of rocks 49, 55
potential difference 60 radioatoms 49
potential drops in SP observations 80 radionuclids 49
potential gradient (jρ) 79 radon measurement 58
power amplification (dB) 128 radon-222 58
power attenuation (dB) 128 random noise 17, 167
power ratios expressed in decibels (dB) 128 range of α and β particles 55

220
ratio of magnitudes 99 Schlumberger array 67
ratio of the amplitudes, Etx : Hty 122 Schlumberger sounding 71
Rayleigh waves 3 (see also ground roll) scintillation counter 56
ray-path theory 7 scintillometer 137, 144
rays 7 secondary currents 89
real component 95, 96, 97 secondary electromagnetic fields 88
receiver spread 33 secondary field (S0) 96
redundancy 17 secondary field values 115
reflection 2, 8, 14, 33, 37, 132 secondary magnetic field, VLF 122
reflection coefficient 11,26 secondary porosity 142
reflection coefficients, GPR 130 secondary source 9
reflection hyperbola (Sec 2.1) 13 secondary wave, see S-wave 3
reflection stretch 15 secondary-field amplitude (S0) 97
refraction 2, 8, 14, 33 secondary-field phase (ψ) 97
refraction equation 39 secondary-field response 117
refraction time intercept 43, 46 secondary-field response of a single-turn
refraction velocity 34, 36, 38, 39, 42 - 47 loop to a sinusoidal field 98, 100
relative magnetic permeability (µ) 129 sedimentary rocks 6
relative permeability (µr) 90 seismic energy source 155
relative permittivity 131, 134, 184 seismic noise 167
remanence 91 seismic symbols and units 191
remanent magnetism 94 seismic velocities (Sec 1.1) 5
repetition frequency (νr) 126 seismic wavelengths 7
reservoir estimate 141 seismometer 7
resistance in ohms (Ω) 60, 61 self-inductance 95, 100
resistivity dispersion 81 self-potential (SP) 138, 140, 143
resistivity graph 71 self-potential log 171
resistivity log 145 self-potential, origins 74
resistivity of a parallelpiped 61 semiconductors 65
resistivity of rocks and minerals 64 SH wave 4 (see also S-wave)
resistivity, ρ (Ωm) 60, 61, 87, 172 shake wave (see S-wave) 3
response functions 100 shallow refraction 167
response of a horizontal-loop EM system to a shear wave (see S-wave) 3
vertical-loop target [f (τ)] 100 shoot-back technique 108, 110
response parameter (τ) 100 siemens 140
resultant field amplitude (R0) 96 siemens per meter 62
resultant-field phase (α) 96 signal-to-noise ratio 164
reverberations 167 sine transform 179, 181
rigidity 4, (see shear modulus) single frequency EM wave (ν = ω / 2π) 128
ringing 169 single square loop 117
rippabilities 5-6 single-pointer electrode array 138
RMS velocity (Sec 2.6) 19 single-turn circular coil 97
rock density 137, 170 skin depth (δ) 51, 103
rock porosity 137 slingram 110
roentgen 57 Snell’s law 7-9, 34, 39, 41
soil air 59
salinity 140 solid friction 11
salinity log 140 sonic log 140, 149 - 151
salinity of leachates 138 sonic velocity 140, 149
salinometer 140, 171 sonic velocity log 171
sampling frequency 127 source generated noise 167
sampling interval 127 source static correction 23

221
SP field prodecure 75 time domain 177
SP jump 143 time domain IP measurement 82, 83
spacing and depth penetration 104 time domain, IP 81
specific resistivity 140 time picks 37
specific times 150 time section (Fig 2.3) 15
specular reflection 131 time-depth plot (Fig 2.11) 22
spike 31 time-distance graph (Sec 1.8) 11, 38, 42,
spike pulse 31 45
spike reflections 31 time-domain electromagnetism (TDEM) 115
spike series 169 time-varying electromagnetic field 88
spread 33 total static correction 24
spread design 33 transient field of a single-turn circular loop
square array 69 116
stacked trace 19 transient-field measurement layouts 117
stacking velocity 16 transient-field methods (TEM) 95, 115
stacking, GPR 132 transit times 6, 201
static correction (Sec 2.8) 22 transmission coefficient 11, 26
strain 3 transmission coefficients GPR, 130
streaming potential coefficient 76 transverse wave, see S-wave 3
streaming potentials 75, 76 travel paths 7
stress 3 travel time 13, 140, 149
stress / point velocity 26 true porosity 148
stripping 57 true resistivity 140, 142
stripping ratios 57 tube waves 164
Sundberg method 108 Turam method 109
superposition (see convolution) 31-32 two-coil fixed source low-frequency
surface waves 4 systems in free-space magnetic fields 100
susceptibility (κ) 90 two-way time, GPR 133
susceptibility of rocks 93 types of radiation 49
SV wave 4 (see also S-wave)
S-wave 3 U-238 decay series 52, 53
synthetic seismogram 151 ultra-high frequency (UHF) 108
unfocused microlog 139
t – x graph (Fig 2.3) 15 universal gas constant (R) 78
TEM – CW comparison 117 up-dip velocity 44, 45
TEM – IP comparison 116 up-dip ray 44
TEM survey procedures 117 uphole geophone 23
temperature 140, 152 uphole survey (Fig 2.11) 22-23
temperature log 140, 171 uphole time 23
terrestrial and cosmic radiation 54 uranium and thorium decay series 136
tesla (T) 89 uranium-235 51
Th-232 decay series 53 uranium-238 51
thallium activated crystals 56
theory of elasticity 3 valence of ions 78
thermal neutrons 147 variable-area display (VAR) 128
thermoremanent magnetization (TRM) 94 variation of induced current with depth 105
thin conductor anomaly 121 variation of skin depth, δ, with frequency
thin layer resolution 139 and resistivity 103
thorium-232 51 velocities in rocks (Sec 1.3) 6
tilt of a magnetic field 99 velocity of an electromagnetic wave 129
tilt-angle methods 107 velocity of light 129
time averaging (Sec 1.2) 6 vertical dipole 109

222
vertical electric sounding VES, 70 VSP field plot 160
vertical electromagnetic field amplitude 109 VSP noise 163
vertical electromagnetic field phase 109 VSP quality control 163
vertical fluid flow 141
vertical secondary field in phase quadrature walkaway 155
with the primary field 121 water-layer multiples 169
vertical seismic profiling, see VSP 7 wave impedance 124
very low frequency (VLF) 118 wave motion equation 170
VES field procedure 71 wavefront 7
Vibroseis 166 waveguide 118
viscosity 11 wavelength of the radiation 50
viscous remanent magnetization (VRM) 94 wavelength of γ rays 50
VLF – EM comparison 125 weathered layer 22
VLF coupling 122 weathering 6, 34
VLF field effects at a distant point 120 weathering correction 22
VLF measurement in strike direction 122 weber (Wb) 89
VLF measurement with station perpendicular weighted sine and cosine wave 179
to strike 124 well logging 136 (Chap. 5)
VLF methods 118 well logging symbols and units 197
VLF phase 123 well velocity survey 7, 21, 155, 197
VLF radio transmitters 119 well velocity tools 157
VLF symbols and units 196 Wenner array 66
VLF wave in the ground 122 Wenner electric sounding 69
volt second (V s) 89 Wenner sounding 71
voltage (U) 60 world-wide VLF radio transmitters 119
voltage differences 87
voltage drop 60 Young’s modulus (q) 5
VSP 7
VSP (vertical seismic profiling) 155, 159 zeta potentials 76

223
Notes
Notes

You might also like