You are on page 1of 145

Environmental Geophysics Reader

Compiled from the internet


by Christopher L. Liner

December 2000

Contents

I Gravity Surveying 5

1 Geophysical Surveying Using Gravity (T. Boyd) 5


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Gravitational Force . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Gravitational Acceleration . . . . . . . . . . . . . . . . 5
1.1.3 Units Associated With Gravitational Acceleration . . 6
1.2 Gravity and Geology . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.1 How is the Gravitational Acceleration, g, Related to
Geology? . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 The Relevant Geologic Parameter is not Density, but
Density Contrast . . . . . . . . . . . . . . . . . . . . . 12
1.2.3 Density Variations of Earth Materials . . . . . . . . . 14
1.2.4 A Simple Model . . . . . . . . . . . . . . . . . . . . . 14
1.3 Measuring Gravitational Acceleration . . . . . . . . . . . . . 15
1.3.1 How do we Measure Gravity? . . . . . . . . . . . . . . 15
1.3.2 Falling Body Measurements . . . . . . . . . . . . . . . 17
1.3.3 Pendulum Measurements . . . . . . . . . . . . . . . . 19
1.3.4 Mass and Spring Measurements . . . . . . . . . . . . . 21
1.4 Factors That Affect Gravitational Acceleration . . . . . . . . 24
1.4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.4.2 Temporal Based Variations . . . . . . . . . . . . . . . 26
1.4.3 Spatial Based Variations . . . . . . . . . . . . . . . . . 35
1.4.4 Summary of Gravity Types . . . . . . . . . . . . . . . 48
1.5 Isolating Gravity Anomalies of Interest . . . . . . . . . . . . . 50

1
Liner Env Geophysics Reader 2

1.5.1 Local and Regional Gravity Anomalies . . . . . . . . . 50


1.5.2 Sources of the Local and Regional Gravity Anomalies 52
1.5.3 Separating Local and Regional Gravity Anomalies . . 55
1.5.4 Local/Regional Gravity Anomaly Separation Example 57
1.6 Gravity Anomalies Over Bodies With Simple Shapes . . . . . 59
1.6.1 Gravity Anomaly Over a Buried Point Mass . . . . . . 59
1.6.2 Gravity Anomaly Over a Buried Sphere . . . . . . . . 62
1.6.3 Model Indeterminancy . . . . . . . . . . . . . . . . . . 64
1.6.4 Gravity Calculations over Bodies with more Complex
Shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

II Magnetic Surveying 67

2 Geophysical Surveying Using Magnetic Methods (T. Boyd) 68


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.1.1 Historical Overview . . . . . . . . . . . . . . . . . . . 68
2.1.2 Similarities Between Gravity and Magnetics . . . . . . 69
2.1.3 Differences Between Gravity and Magnetics . . . . . . 70
2.1.4 Magnetic Monopoles . . . . . . . . . . . . . . . . . . . 71
2.1.5 Forces Associated with Magnetic Monopoles . . . . . . 72
2.1.6 Magnetic Dipoles . . . . . . . . . . . . . . . . . . . . . 73
2.1.7 Field Lines for a Magnetic Dipole . . . . . . . . . . . . 75
2.1.8 Units Associated with Magnetic Poles . . . . . . . . . 76
2.2 Magnetization of Materials . . . . . . . . . . . . . . . . . . . 78
2.2.1 Induced Magnetization . . . . . . . . . . . . . . . . . . 78
2.2.2 Magnetic Susceptibility . . . . . . . . . . . . . . . . . 78
2.2.3 Mechanisms of Magnetic Induction . . . . . . . . . . . 79
2.2.4 Suseptibilities of Common Rocks and Minerals . . . . 81
2.2.5 Remanent Magnetism . . . . . . . . . . . . . . . . . . 83
2.3 The Earths Magnetic Field . . . . . . . . . . . . . . . . . . . 85
2.3.1 Magnetic Field Nomenclature . . . . . . . . . . . . . . 85
2.3.2 The Earths Main Field . . . . . . . . . . . . . . . . . 86
2.3.3 Magnetics and Geology - A Simple Example . . . . . . 88
2.3.4 Temporal Variations of the Earths Main Field - Overview 91
2.3.5 Secular Variations . . . . . . . . . . . . . . . . . . . . 92
2.3.6 Diurnal Variations . . . . . . . . . . . . . . . . . . . . 94
2.3.7 Magnetic Storms . . . . . . . . . . . . . . . . . . . . . 94
2.4 Magnetometers . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.4.1 Instrumentation Overview . . . . . . . . . . . . . . . . 95
Liner Env Geophysics Reader 3

2.4.2 Fluxgate Magnetometers . . . . . . . . . . . . . . . . . 97


2.4.3 Proton Precession Magnetomenters . . . . . . . . . . . 98
2.4.4 Total Field Measurements . . . . . . . . . . . . . . . . 100
2.5 FieldProcedures . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.5.1 Modes of Acquiring Magnetic Observations . . . . . . 102
2.5.2 Assuring High-Quality Observations - Magnetic Clean-
liness . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
2.5.3 Strategies for Dealing with Temporal Variations . . . 104
2.5.4 Spatially Varying Corrections? . . . . . . . . . . . . . 106
2.5.5 Correcting for the Main-Field Contributions . . . . . . 107
2.6 Magnetic Anomalies Over Simple Shapes . . . . . . . . . . . 108
2.6.1 Comparison Between Gravity and Magnetic Anomalies 108
2.6.2 Magnetic Anomaly: Magnetized Sphere at the North
Pole . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
2.6.3 Magnetic Anomaly: Magnetized Sphere at the Equator 111
2.6.4 Magnetic Anomaly: Magnetized Sphere in the North-
ern Hemisphere . . . . . . . . . . . . . . . . . . . . . . 113

III Seismic Refraction Surveying 116

3 My section 116
3.1 My subsection . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
3.1.1 My subsubsection . . . . . . . . . . . . . . . . . . . . 116

IV Electrical Conductivity Surveying 116

4 Electromagnetic techniques 116

5 Conductivity overview 117

6 Terrain conductivity 118


6.1 Temperature effect . . . . . . . . . . . . . . . . . . . . . . . . 120
6.2 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

7 Linearity and resistivity 121

8 Conductivity and water quality 122


Liner Env Geophysics Reader 4

9 EM31 conductivity meter 123


9.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
9.2 Basic Interpretation . . . . . . . . . . . . . . . . . . . . . . . 125
9.3 Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
9.4 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
9.5 Specifications . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
9.6 Frequently asked questions . . . . . . . . . . . . . . . . . . . 135
9.7 Field procedure . . . . . . . . . . . . . . . . . . . . . . . . . . 135
9.8 Case history 1: Soil and rock characterization . . . . . . . . . 136
9.9 Case history 2: Industrial waste . . . . . . . . . . . . . . . . . 136
9.10 Case history 3: Contaminate plume . . . . . . . . . . . . . . . 137

V Ground Penetrating Radar Surveying 145

10 My section 145
10.1 My subsection . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
10.1.1 My subsubsection . . . . . . . . . . . . . . . . . . . . 145
5

Part I
Gravity Surveying
1 Geophysical Surveying Using Gravity (T. Boyd)

1.1 Introduction
1.1.1 Gravitational Force

Geophysical interpretations from gravity surveys are based on the mutual


attraction experienced between two masses1 as first expressed by Isaac New-
ton in his classic work Philosophiae naturalis principa mathematica (The
mathematical principles of natural philosophy).
Newtons law of gravitation
G m1 m2
F = (1)
r2
states that the mutual attractive force between two point masses2 , m1 and
m2, is proportional to one over the square of the distance between them.
The constant of proportionality is usually specified as G, the gravitational
constant. Thus, we usually see the law of gravitation written as shown to
the right where F is the force of attraction, G is the gravitational constant,
and r is the distance between the two masses, m1 and m2 .

1.1.2 Gravitational Acceleration

When making measurements of the earths gravity, we usually dont measure


the gravitational force, F. Rather, we measure the gravitational acceleration,
g. The gravitational acceleration is the time rate of change of a bodys speed
1
As described on the next page, mass is formally defined as the proportionality constant
relating the force applied to a body and the accleration the body undergoes as given by
Newtons second law, usually written as F=ma. Therefore, mass is given as m=F/a and
has the units of force over acceleration.
2
A point mass specifies a body that has very small physical dimensions. That is, the
mass can be considered to be concentrated at a single point.
Liner Env Geophysics Reader 6

under the influence of the gravitational force. That is, if you drop a rock off
a cliff, it not only falls, but its speed increases as it falls.
In addition to defining the law of mutual attraction between masses,
Newton also defined the relationship between a force and an acceleration.
Newtons second law states that force is proportional to acceleration.

F = m2 g (2)

The constant of proportionality is the mass of the object.


Combining Newtons second law with his law of mutual attraction,
G m1
g= (3)
r2
the gravitational acceleration on the mass m2 can be shown to be equal to
the mass of attracting object, m1 , over the squared distance between the
center of the two masses, r.

1.1.3 Units Associated With Gravitational Acceleration

Units Associated with Gravitational Acceleration


As described earlier, acceleration is defined as the time rate of change
of the speed of a body. Speed, sometimes incorrectly referred to as velocity,
is the distance an object travels divided by the time it took to travel that
distance (i.e., meters per second (m/s)). Thus, we can measure the speed of
an object by observing the time it takes to travel a known distance, Figure 1
If the speed of the object changes as it travels, then this change in
speed with respect to time is referred to as acceleration. Positive accelera-
tion means the object is moving faster with time, and negative acceleration
means the object is slowing down with time. Acceleration can be measured
by determining the speed of an object at two different times and divid-
ing the speed by the time difference between the two observations, Figure 2.
Therefore, the units associated with acceleration is speed (distance per time)
divided by time; or distance per time per time, or distance per time squared.
Liner Env Geophysics Reader 7

If an object such as a ball is dropped, it falls under the influence of gravity


in such a way that its speed increases constantly with time. That is, the
object accelerates as it falls with constant acceleration. At sea level, the rate
of acceleration is about 9.8 meters per second squared. In gravity surveying,
we will measure variations in the acceleration due to the earths gravity.
As will be described next, variations in this acceleration can be caused by
variations in subsurface geology. Acceleration variations due to geology,
however, tend to be much smaller than 9.8 meters per second squared. Thus,
a meter per second squared is an inconvenient system of units to use when
discussing gravity surveys.
The units typically used in describing the graviational acceleration vari-
ations observed in exploration gravity surveys are specified in milliGals. A
Gal is defined as a centimeter per second squared. Thus, the Earths grav-
itational acceleration is approximately 980 Gals. The Gal is named after
Galileo Galilei. The milliGal (mgal) is one thousandth of a Gal. In milli-
Gals, the Earths gravitational acceleration is approximately 980,000.
The following relationships among units hold

1 gal = 102 m/s2


= 102 newton/m2
1 mgal = 105 m/s2
= 103 gal
1 gal = 108 m/s2
= 103 mgal .
Liner Env Geophysics Reader 8

Figure 1: The concept of velocity.

Figure 2: The concept of acceleration.


Liner Env Geophysics Reader 9

1.2 Gravity and Geology


1.2.1 How is the Gravitational Acceleration, g, Related to Geol-
ogy?

Density is defined as mass per unit volume. For example, if we were to


calculate the density of a room filled with people, the density would be given
by the average number of people per unit space (e.g., per cubic foot) and
would have the units of people per cubic foot. The higher the number, the
more closely spaced are the people. Thus, we would say the room is more
densely packed with people. The units typically used to describe density
of substances are grams per centimeter cubed (gm/cm3 ); mass per unit
volume. In relating our room analogy to substances, we can use the point
mass described earlier as we did the number of people.
Consider a simple geologic example of an ore body buried in soil, Fig-
ure 3. We would expect the density of the ore body, d2, to be greater than
the density of the surrounding soil, d1.
The density of the material can be thought of as a number that quan-
tifies the number of point masses needed to represent the material per unit
volume of the material just like the number of people per cubic foot in the
example given above described how crowded a particular room was. Thus,
to represent a high-density ore body, we need more point masses per unit
volume than we would for the lower density soil3 , Figure 4
Now, lets qualitatively describe the gravitational acceleration experi-
enced by a ball as it is dropped from a ladder, Figure 5. This acceleration
can be calculated by measuring the time rate of change of the speed of the
ball as it falls. The size of the acceleration the ball undergoes will be pro-
portional to the number of close point masses that are directly below it.
Were concerned with the close point masses because the magnitude of the
gravitational acceleration varies as one over the distance between the ball
and the point mass squared. The more close point masses there are directly
3
In this discussion we assume that all of the point masses have the same mass.
Liner Env Geophysics Reader 10

below the ball, the larger its acceleration will be.


We could, therefore, drop the ball from a number of different locations
(Figure 6), and, because the number of point masses below the ball varies
with the location at which it is dropped, map out differences in the size of
the gravitational acceleration experienced by the ball caused by variations
in the underlying geology. A plot of the gravitational acceleration versus
location is commonly referred to as a gravity profile.
This simple thought experiment forms the physical basis on which gravity
surveying rests.

Figure 3: Earth density model of an ore body.

Figure 4: Point mass representation of the ore body density model.


Liner Env Geophysics Reader 11

Figure 5: More point masses mean more acceleration.

Figure 6: Building a gravity profile.


Liner Env Geophysics Reader 12

1.2.2 The Relevant Geologic Parameter is not Density, but Den-


sity Contrast

Contrary to what you might first think, the shape of the curve describing
the variation in gravitational acceleration is not dependent on the absolute
densities of the rocks. It is only dependent on the density difference (usually
referred to as density contrast) between the ore body and the surrounding
soil. That is, the spatial variation in the gravitational acceleration4 gener-
ated from our previous example would be exactly the same if we were to
assume different densities for the ore body and the surrounding soil, as long
as the density contrast, d2 - d1, between the ore body and the surrounding
soil were constant. One example of a model that satisfies this condition is
to let the density of the soil be zero and the density of the ore body be d2 -
d1, Figure 7.
The only difference in the gravitational accelerations produced by the
two structures shown above (one given by the original model and one given
by setting the density of the soil to zero and the ore body to d2 - d1)
is an offset in the curve derived from the two models. The offset is such
that at great distances from the ore body, the gravitational acceleration
approaches zero in the model which uses a soil density of zero rather than the
non-zero constant value the acceleration approaches in the original model.
For identifying the location of the ore body, the fact that the gravitational
accelerations approach zero away from the ore body instead of some non-zero
number is unimportant. What is important is the size of the difference in the
gravitational acceleration near the ore body and away from the ore body and
the shape of the spatial variation in the gravitational acceleration. Thus, the
latter model that employs only the density contrast of the ore body to the
surrounding soil contains all of the relevant information needed to identify
the location and shape of the ore body.
4
It is common to use expressions like Gravity Field as a synonym for gravitational
acceleration.
Liner Env Geophysics Reader 13

Figure 7: Relative density contrast is the important quantity.


Liner Env Geophysics Reader 14

1.2.3 Density Variations of Earth Materials

Thus far it sounds like a fairly simple proposition to estimate the variation
in density of the earth due to local changes in geology. There are, how-
ever, several significant complications. The first has to do with the density
contrasts measured for various earth materials.
The densities associated with various earth materials are shown in Ta-
ble 1.
Notice that the relative variation in rock density is quite small, 0.8 gm/cm3 ,
and there is considerable overlap in the measured densities. Hence, a knowl-
edge of rock density alone will not be sufficient to determine rock type.
This small variation in rock density also implies that the spatial varia-
tions in the observed gravitational acceleration caused by geologic structures
will be quite small and thus difficult to detect.

Material Density g/cm3


Air 0
Water 1
Sediments 1.7
Sandstone 2.0-2.6
Shale 2.0-2.7
Limestone 2.5-2.8
Granite 2.5-2.8
Basalt 2.7-3.1
Metamorphic Rocks 2.6-3.0

Table 1: Density range of typical earth materials.

1.2.4 A Simple Model

Consider the variation in gravitational acceleration that would be observed


over a simple model, Figure 8. For this model, lets assume that the only
variation in density in the subsurface is due to the presence of a small ore
body. Let the ore body have a spherical shape with a radius of 10 meters,
Liner Env Geophysics Reader 15

buried at a depth of 25 meters below the surface, and with a density contrast
to the surrounding rocks of 0.5 g/cm3 From the table of rock densities,
Table 1, notice that the chosen density contrast is actually fairly large. The
specifics of how the gravitational acceleration was computed are not, at this
time, important.
There are several things to notice about the gravity anomaly5 produced
by this structure.
The gravity anomaly produced by a buried sphere is symmetric about
the center of the sphere. The maximum value of the anomaly is quite small.
For this example, 0.025 mgals. The magnitude of the gravity anomaly ap-
proaches zero at small ( 60 meters) horizontal distances away from the
center of the sphere.
Later, we will explore how the size and shape of the gravity anomaly
is affected by the model parameters such as the radius of the ore body, its
density contrast, and its depth of burial. At this time, simply note that
the gravity anomaly produced by this reasonably-sized ore body is small.
When compared to the gravitational acceleration produced by the earth as
a whole, 980000 mgals, the anomaly produced by the ore body represents a
change in the gravitational field of only 1 part in 40 million.
Clearly, a variation in gravity this small is going to be difficult to mea-
sure. Also, factors other than geologic structure might produce variations
in the observed gravitational acceleration that are as large, if not larger.

1.3 Measuring Gravitational Acceleration


1.3.1 How do we Measure Gravity?

How do we Measure Gravity?


As you can imagine, it is difficult to construct instruments capable of
measuring gravity anomalies as small as 1 part in 40 million. There are,
5
We will often use the term gravity anomaly to describe variations in the background
gravity field produced by local geologic structure or a model of local geologic structure.
Liner Env Geophysics Reader 16

Figure 8: Gravitational acceleration over a buried sphere.


Liner Env Geophysics Reader 17

however, a variety of ways it can be done, including:

Falling body measurements These are the type of measurements we have


described up to this point. One drops an object and directly computes
the acceleration the body undergoes by carefully measuring distance
and time as the body falls.

Pendulum measurements In this type of measurement, the gravitational


acceleration is estimated by measuring the period oscillation of a pen-
dulum.

Mass on spring measurements By suspending a mass on a spring and


observing how much the spring deforms under the force of gravity, an
estimate of the gravitational acceleration can be determined.

As will be described later, in exploration gravity surveys, the field ob-


servations usually do not yield measurements of the absolute value of grav-
itational acceleration. Rather, we can only derive estimates of variations
of gravitational acceleration. The primary reason for this is that it can be
difficult to characterize the recording instrument well enough to measure
absolute values of gravity down to 1 part in 50 million. This, however, is
not a limitation for exploration surveys since it is only the relative change
in gravity that is used to define the variation in geologic structure.

1.3.2 Falling Body Measurements

The gravitational acceleration can be measured directly by dropping an


object and measuring its time rate of change of speed (acceleration) as it
falls. By tradition, this is the method we have commonly ascribed to Galileo
Galilei. In this experiment, Galileo is supposed to have dropped objects of
varying mass from the leaning tower of Pisa and found that the gravitational
acceleration an object undergoes is independent of its mass. He is also said to
have estimated the value of the gravitational acceleration in this experiment.
Liner Env Geophysics Reader 18

While it is true that Galileo did make these observations, he didnt use a
falling body experiment to do them. Rather, he used measurements based
on pendulums.
It is easy to show that the distance a body falls is proportional to the
time it has fallen squared. The proportionality constant is the gravitational
acceleration, g. Therefore, by measuring distances and times as a body falls,
it is possible to estimate the gravitational acceleration, Figure 9.
To measure changes in the gravitational acceleration down to 1 part in
40 million using an instrument of reasonable size (say one that allows the
object to drop 1 meter), we need to be able to measure changes in distance
down to 1 part in 10 million and changes in time down to 1 part in 100
million!! As you can imagine, it is difficult to make measurements with this
level of accuracy.
It is, however, possible to design an instrument capable of measuring
accurate distances and times and computing the absolute gravity down to 1
microgal (0.001 mgals; this is a measurement accuracy of almost 1 part in 1
billion!!). Micro-g Solutions is one manufacturer of this type of instrument,
known as an Absolute Gravimeter. Unlike the instruments described next,
this class of instruments is the only field instrument designed to measure
absolute gravity. That is, this instrument measures the size of the verti-
cal component of gravitational acceleration at a given point. As described
previously, the instruments more commonly used in exploration surveys are
capable of measuring only the change in gravitational acceleration from point
to point, not the absolute value of gravity at any one point.
Although absolute gravimeters are more expensive than the traditional,
relative gravimeters and require a longer station occupation time (1/2 day to
1 day per station), the increased precision offered by them and the fact that
the looping strategies described later are not required to remove instrument
drift or tidal variations may outweigh the extra expense in operating them.
This is particularly true when survey designs require large station spacings
Liner Env Geophysics Reader 19

or for experiments needing the continuous monitoring of the gravitational


acceleration at a single location. As an example of this latter application,
it is possible to observe as little as 3 mm of crustal uplift over time by
monitoring the change in gravitational acceleration at a single location with
one of these instruments.

Figure 9: Falling body experiment to estimate gravitational acceleration.

1.3.3 Pendulum Measurements

Another method by which we can measure the acceleration due to gravity is


to observe the oscillation of a pendulum, such as that found on a grandfather
clock. Contrary to popular belief, Galileo Galilei made his famous gravity
observations using a pendulum, not by dropping objects from the Leaning
Liner Env Geophysics Reader 20

Tower of Pisa.
If we were to construct a simple pendulum by hanging a mass from a
rod and then displace the mass from vertical, the pendulum would begin
to oscillate about the vertical in a regular fashion, Figure 10. The rele-
vant parameter that describes this oscillation is known as the period6 of
oscillation.
The reason that the pendulum oscillates about the vertical is that if the
pendulum is displaced, the force of gravity pulls down on the pendulum. The
pendulum begins to move downward. When the pendulum reaches vertical
it cant stop instantaneously. The pendulum continues past the vertical and
upward in the opposite direction. The force of gravity slows it down until
it eventually stops and begins to fall again. If there is no friction where the
pendulum is attached to the ceiling and there is no wind resistance to the
motion of the pendulum, this would continue forever.
Because it is the force of gravity that produces the oscillation, one might
expect the period of oscillation to differ for differing values of gravity. In
particular, if the force of gravity is small, there is less force pulling the
pendulum downward, the pendulum moves more slowly toward vertical, and
the observed period of oscillation becomes longer. Thus, by measuring the
period of oscillation of a pendulum, we can estimate the gravitational force
or acceleration.
It can be shown that the period of oscillation of the pendulum, T, is
proportional to one over the square root of the gravitational acceleration, g.
q
T =2 (k/g) (4)

The constant of proportionality, k, depends on the physical characteristics


of the pendulum such as its length and the distribution of mass about the
pendulums pivot point.
6
The period of oscillation is the time required for the pendulum to complete one cycle
in its motion. This can be determined by measuring the time required for the pendulum
to reoccupy a given position. In the Figure 10 example, the period of oscillation of the
pendulum is approximately two seconds.
Liner Env Geophysics Reader 21

Like the falling body experiment described previously, it seems like it


should be easy to determine the gravitational acceleration by measuring the
period of oscillation. Unfortunately, to be able to measure the acceleration
to 1 part in 50 million requires a very accurate estimate of the instrument
constant k. K cannot be determined accurately enough to do this.
All is not lost, however. We could measure the period of oscillation of
a given pendulum at two different locations. Although we can not estimate
k accurately enough to allow us to determine the gravitational acceleration
at either of these locations because we have used the same pendulum at the
two locations, we can estimate the variation in gravitational acceleration at
the two locations quite accurately without knowing k.
The small variations in pendulum period that we need to observe can be
estimated by allowing the pendulum to oscillate for a long time, counting the
number of oscillations, and dividing the time of oscillation by the number
of oscillations. The longer you allow the pendulum to oscillate, the more
accurate your estimate of pendulum period will be. This is essentially a
form of averaging. The longer the pendulum oscillates, the more periods
over which you are averaging to get your estimate of pendulum period, and
the better your estimate of the average period of pendulum oscillation.
In the past, pendulum measurements were used extensively to map the
variation in gravitational acceleration around the globe. Because it can take
up to an hour to observe enough oscillations of the pendulum to accurately
determine its period, this surveying technique has been largely supplanted
by the mass on spring measurements described next.

1.3.4 Mass and Spring Measurements

The most common type of gravimeter7 used in exploration surveys is based


on a simple mass-spring system. If we hang a mass on a spring, the force
7
A gravimeter is any instrument designed to measure spatial variations in gravitational
acceleration.
Liner Env Geophysics Reader 22

Figure 10: Motion of a simple pendulum.

of gravity will stretch the spring by an amount that is proportional to the


gravitational force, Figure reftbmass. It can be shown that the proportion-
ality between the stretch of the spring and the gravitational acceleration is
the magnitude of the mass hung on the spring divided by a constant, k,
which describes the stiffness of the spring. The larger k is, the stiffer the
spring is, and the less the spring will stretch for a given value of gravitational
acceleration,
mg
x= . (5)
k
Like pendulum measurements, we can not determine k accurately enough
to estimate the absolute value of the gravitational acceleration to 1 part
in 40 million. We can, however, estimate variations in the gravitational
acceleration from place to place to within this precision. To be able to
Liner Env Geophysics Reader 23

do this, however, a sophisticated mass-spring system is used that places the


mass on a beam and employs a special type of spring known as a zero-length
spring.
Instruments of this type are produced by several manufacturers; LaCoste
and Romberg8 , Texas Instruments (Worden Gravity Meter), and Scintrex.
Modern gravimeters are capable of measuring changes in the Earths grav-
itational acceleration down to 1 part in 100 million. This translates to a
precision of about 0.01 mgal. Such a precision can be obtained only under
optimal conditions when the recommended field procedures are carefully
followed.

Figure 11: Mass and spring experiment to estimate gravitational accelera-


tion.

8
Figure from Introduction to Geophysical Prospecting, M. Dobrin and C. Savit.
Liner Env Geophysics Reader 24

Figure 12: Worden gravimeter.

1.4 Factors That Affect Gravitational Acceleration


1.4.1 Overview

Thus far we have shown how variations in the gravitational acceleration can
be measured and how these changes might relate to subsurface variations
in density. Weve also shown that the spatial variations in gravitational
acceleration expected from geologic structures can be quite small.
Because these variations are so small, we must now consider other factors
that can give rise to variations in gravitational acceleration that are as large,
if not larger, than the expected geologic signal. These complicating factors
can be subdivided into two catagories: those that give rise to temporal
variations and those that give rise to spatial variations in the gravitational
acceleration.

Temporal Based Variations These are changes in the observed acceler-


ation that are time dependent. In other words, these factors cause
Liner Env Geophysics Reader 25

Figure 13: LaCoste and Romberg gravimeter.

variations in acceleration that would be observed even if we didnt


move our gravimeter.

Instrument Drift Changes in the observed acceleration caused by


changes in the response of the gravimeter over time.
Tidal Affects Changes in the observed acceleration caused by the
gravitational attraction of the sun and moon.

Spatial Based Variations These are changes in the observed acceleration


that are space dependent. That is, these change the gravitational
acceleration from place to place, just like the geologic affects, but they
Liner Env Geophysics Reader 26

are not related to geology.

Latitude Variations Changes in the observed acceleration caused


by the ellipsoidal shape and the rotation of the earth.

Elevation Variations Changes in the observed acceleration caused


by differences in the elevations of the observation points.

Slab Effects Changes in the observed acceleration caused by the ex-


tra mass underlying observation points at higher elevations.

Topographic Effects Changes in the observed acceleration related


to topography near the observation point.

1.4.2 Temporal Based Variations

Instrument Drift Defintion of instrument drift: A gradual and uninten-


tional change in the reference value with respect to which measurements are
made9 .
Although constructed to high-precision standards and capable of mea-
suring changes in gravitational acceleration to 0.01 mgal, problems do exist
when trying to use a delicate instrument such as a gravimeter.
Even if the instrument is handled with great care (as it always should
be - new gravimeters cost about US$30,000), the properties of the materials
used to construct the spring can change with time. These variations in
spring properties with time can be due to stretching of the spring over
time or to changes in spring properties related to temperature changes. To
help minimize the later, gravimeters are either temperature controlled or
constructed out of materials that are relatively insensitive to temperature
changes. Even still, gravimeters can drift as much as 0.1 mgal per day.
Shown in Figure 14 is an example of a gravity data set10 collected at the
same site over a two day period. There are two things to notice from this
9
Definition from the Encyclopedic Dictionary of Exploration Geophysicists by R. E.
Sheriff, published by the Society of Exploration Geophysicists.
10
Data are from: Wolf, A. Tidal Force Observations, Geophysics, V, 317-320, 1940.
Liner Env Geophysics Reader 27

set of observations. First, notice the oscillatory behavior of the observed


gravitational acceleration. This is related to variations in gravitational ac-
celeration caused by the tidal attraction of the sun and the moon. Second,
notice the general increase in the gravitational acceleration with time. This
is highlighted by the green line. This line represents a least-squares, best-fit
straight line to the data. This trend is caused by instrument drift. In this
particular example, the instrument drifted approximately 0.12 mgal in 48
hours.

Figure 14: Tidal variations at a fixed location (red) and the interpreted
instrument drift (green).

Tides Definition of tidal effect11 : Variations in gravity observations re-


sulting from the attraction of the moon and sun and the distortion of the
earth so produced.
Superimposed on instrument drift is another temporally varying compo-
11
Definition from the Encyclopedic Dictionary of Exploration Geophysicists by R. E.
Sheriff, published by the Society of Exploration Geophysicists.
Liner Env Geophysics Reader 28

nent of gravity. Unlike instrument drift, which results from the temporally
varying characteristics of the gravimeter, this component represents real
changes in the gravitational acceleration. Unfortunately, these are changes
that do not relate to local geology and are hence a form of noise in our
observations.
Just as the gravitational attraction of the sun and the moon distorts the
shape of the ocean surface, it also distorts the shape of the earth. Because
rocks yield to external forces much less readily than water, the amount the
earth distorts under these external forces is far less than the amount the
oceans distort. The size of the ocean tides, the name given to the distortion
of the ocean caused by the sun and moon, is measured in terms of meters.
The size of the solid earth tide, the name given to the distortion of the earth
caused by the sun and moon, is measured in terms of centimeters.
This distortion of the solid earth produces measurable changes in the
gravitational acceleration because as the shape of the earth changes, the
distance of the gravimeter to the center of the earth changes (recall that
gravitational acceleration is proportional to one over distance squared). The
distortion of the earth varies from location to location, but it can be large
enough to produce variations in gravitational acceleration as large as 0.2
mgals. This effect would easily overwhelm the example gravity anomaly
described previously.
An example of the variation in gravitational acceleration observed at
one location (Tulsa, Oklahoma) is shown in Figure 1512 . These are raw
observations that include both instrument drift (notice how there is a general
trend in increasing gravitational acceleration with increasing time) and tides
(the cyclic variation in gravity with a period of oscillation of about 12 hours).
In this case the amplitude of the tidal variation is about 0.15 mgals, and
the amplitude of the drift appears to be about 0.12 mgals over two days
(Figure 14).
12
Data are from: Wolf, A. Tidal Force Observations, Geophysics, V, 317-320, 1940.
Liner Env Geophysics Reader 29

Figure 15: Time-variation in gravitational acceleration at one location.

A Correction Strategy for Instrument Drift and Tides The result


of the drift and the tidal portions of our gravity observations is that repeated
observations at one location yield different values for the gravitational ac-
celeration. The key to making effective corrections for these factors is to
note that both alter the observed gravity field as slowly varying functions
of time.
One possible way of accounting for the tidal component of the gravity
field would be to establish a base station* near the survey area and to
continuously monitor the gravity field at this location while other gravity
observations are being collected in the survey area. This would result in a
record of the time variation of the tidal components of the gravity field that
could be used to correct the survey observations.
Definition of base station13 : A reference station that is used to estab-
lish additional stations in relation thereto. Quantities under investigation
13
Definition from the Encyclopedic Dictionary of Exploration Geophysicists by R. E.
Sheriff, published by the Society of Exploration Geophysicists.
Liner Env Geophysics Reader 30

have values at the base station that are known (or assumed to be known)
accurately. Data from the base station may be used to normalize data from
other stations.**
This procedure is rarely used for a number of reasons.

It requires the use of two gravimeters. For many gravity surveys, this
is economically infeasable.

The use of two instruments requires the mobilization of two field crews,
again adding to the cost of the survey.

Most importantly, although this technique can be used to remove the


tidal component, it will not remove instrument drift. Because two
different instruments are being used, they will exhibit different drift
characteristics. Thus, an additional drift correction would have to be
performed. Since, as we will show below, this correction can also be
used to eliminate earth tides, there is no reason to incur the extra
costs associated with operating two instruments in the field.

Instead of continuously monitoring the gravity field at the base station,


it is more common to periodically reoccupy (return to) the base station.
This procedure has the advantage of requiring only one gravimeter to mea-
sure both the time variable component of the gravity field and the spatially
variable component. Also, because a single gravimeter is used, corrections
for tidal variations and instrument drift can be combined.
Shown in Figure 16 is an enlargement of the tidal data set shown previ-
ously. Notice that because the tidal and drift components vary slowly with
time, we can approximate these components as a series of straight lines. One
such possible approximation is shown below as the series of green lines. The
only observations needed to define each line segment are gravity observations
at each end point, four points in this case. Thus, instead of continuously
monitoring the tidal and drift components, we could intermittantly measure
Liner Env Geophysics Reader 31

them. From these intermittant observations, we could then assume that the
tidal and drift components of the field varied linearly (that is, are defined
as straight lines) between observation points, and predict the time-varying
components of the gravity field at any time.
For this method to be successful, it is vitally important that the time
interval used to intermittantly measure the tidal and drift components not
be too large. In other words, the straight-line segments used to estimate
these components must be relatively short. If they are too large, we will get
inaccurate estimates of the temporal variability of the tides and instrument
drift.
For example, assume that instead of using the green lines to estimate the
tidal and drift components we could use the longer line segments shown in
blue. Obviously, the blue line is a poor approximation to the time-varying
components of the gravity field. If we were to use it, we would incorrectly
account for the tidal and drift components of the field. Furthermore, because
we only estimate these components intermittantly (that is, at the end points
of the blue line) we would never know we had incorrectly accounted for these
components.

Tidal and Drift Corrections: A Field Procedure Lets now consider


an example of how we would apply this drift and tidal correction strategy to
the acquisition of an exploration data set. Consider the small portion of a
much larger gravity survey shown in Figure 18. To apply the corrections, we
must use the following procedure when acquiring our gravity observations:

Establish the location of one or more gravity base stations. The lo-
cation of the base station for this particular survey is shown as the
yellow circle. Because we will be making repeated gravity observa-
tions at the base station, its location should be easily accessible from
the gravity stations comprising the survey. This location is identified,
for this particular station, by station number 9625 (This number was
Liner Env Geophysics Reader 32

Figure 16: Detail plot of tidal and drift variation in gravity data from a
fixed station.

choosen simply because the base station was located at a permanent


survey marker with an elevation of 9625 feet).

Establish the locations of the gravity stations appropriate for the par-
ticular survey. In this example, the location of the gravity stations are
indicated by the blue circles. On the map, the locations are identified
by a station number, in this case 158 through 163.

Before starting to make gravity observations at the gravity stations,


the survey is initiated by recording the relative gravity at the base
station and the time at which the gravity is measured.

We now proceed to move the gravimeter to the survey stations num-


bered 158 through 163. At each location we measure the relative
gravity at the station and the time at which the reading is taken.

After some time period, usually on the order of an hour, we return to


Liner Env Geophysics Reader 33

Figure 17: The true tide/drift curve (dots) can be approximated by straight
line segments (green). But if the time interval between calibration measure-
ments is too long, we get a poor approximation (blue).

the base station and remeasure the relative gravity at this location.
Again, the time at which the observation is made is noted.

If necessary, we then go back to the survey stations and continue mak-


ing measurements, returning to the base station every hour.

After recording the gravity at the last survey station, or at the end of
the day, we return to the base station and make one final reading of
the gravity.

The procedure described above is generally referred to as a looping pro-


cedure with one loop of the survey being bounded by two occupations of
the base station. The looping procedure defined here is the simplest to im-
plement in the field. More complex looping schemes are often employed,
particularly when the survey, because of its large aerial extent, requires the
use of multiple base stations.
Liner Env Geophysics Reader 34

Figure 18: Gravity station basemap.

Tidal and Drift Corrections: Data Reduction Using observations


collected by the looping field procedure, it is relatively straight forward to
correct these observations for instrument drift and tidal effects. The basis
for these corrections will be the use of linear interpolation to generate a
prediction of what the time-varying component of the gravity field should
look like. Shown in Figure 19 is a reproduction of the spreadsheet used to
reduce the observations collected in the survey defined on the last page.
The first three columns of the spreadsheet present the raw field obser-
vations; column 1 is simply the daily reading number (that is, this is the
first, second, or fifth gravity reading of the day), column 2 lists the time
of day that the reading was made (times listed to the nearest minute are
sufficient), column 3 represents the raw instrument reading (although an in-
Liner Env Geophysics Reader 35

strument scale factor needs to be applied to convert this to relative gravity,


and we will assume this scale factor is one in this example).
A plot of the raw gravity observations versus survey station number is
shown in Figure 20. Notice that there are three readings at station 9625.
This is the base station which was occupied three times. Although the lo-
cation of the base station is fixed, the observed gravity value at the base
station each time it was reoccupied was different. Thus, there is a time vary-
ing component to the observed gravity field. To compute the time-varying
component of the gravity field, we will use linear interpolation between sub-
sequent reoccupations of the base station. For example, the value of the
temporally varying component of the gravity field at the time we occupied
station 159 (dark gray line) is computed using the expressions given in Fig-
ure 21.
After applying corrections like these to all of the stations, the temporally
corrected gravity observations are plotted in Figure 22.
There are several things to note about the corrections and the corrected
observations.
One check to make sure that the corrections have been applied correctly
is to look at the gravity observed at the base station. After application of
the corrections, all of the gravity readings at the base station should all be
zero. The uncorrected observations show a trend of increasing gravitational
acceleration toward higher station number. After correction, this trend no
longer exists. The apparent trend in the uncorrected observations is a result
of tides and instrument drift.

1.4.3 Spatial Based Variations

Latitude Dependent Changes in Gravitational Acceleration Two


features of the earths large-scale structure and dynamics affect our gravity
observations: its shape and its rotation. To examine these effects, lets
consider slicing the earth from the north to the south pole, Figure 23. Our
Liner Env Geophysics Reader 36

Figure 19: Gravity survey worksheet


Liner Env Geophysics Reader 37

Figure 20: Raw data from the gravity survey.

Figure 21: Linear interpolation equations to compute time-varying compo-


nent of the gravity field.
Liner Env Geophysics Reader 38

Figure 22: Gravity data (mgal) after tide/drift correction.

slice will be perpendicular to the equator and will follow a line of constant
longitude between the poles.
Shape To a first-order approximation, the shape of the earth through
this slice is elliptical, with the widest portion of the ellipse aligning with
the equator. This model for the earths shape was first proposed by Isaac
Newton in 1687. Newton based his assessment of the earths shape on a set
of observations provided to him by a friend, named Richer, who happened
to be a navigator on a ship. Richer observed that a pendulum clock that
ran accurately in London consistently lost 2 minutes a day near the equator.
Newton used this observation to estimate the difference in the radius of the
earth measured at the equator from that measured at one of the poles and
came remarkably close to the currently accepted values.
Although the difference in earth radii measured at the poles and at the
equator is only 22 km (this value represents a change in earth radius of
only 0.3in conjunction with the earths rotation, can produce a measurable
Liner Env Geophysics Reader 39

change in the gravitational acceleration with latitude. Because this produces


a spatially varying change in the gravitational acceleration, it is possible to
confuse this change with a change produced by local geologic structure.
Fortunately, it is a relatively simple matter to correct our gravitational ob-
servations for the change in acceleration produced by the earths elliptical
shape and rotation.
To first order14 , the elliptical shape of the earth causes the gravitational
acceleration to vary with latitude because the distance between the gravime-
ter and the earths center varies with latitude. As discussed previously, the
magnitude of the gravitational acceleration changes as one over the distance
from the center of mass of the earth to the gravimeter squared. Thus, quali-
tatively, we would expect the gravitational acceleration to be smaller at the
equator than at the poles, because the surface of the earth is farther from
the earths center at the equator than it is at the poles.
Rotation In addition to shape, the fact that the earth is rotating also
causes a change in the gravitational acceleration with latitude. This affect is
related to the fact that our gravimeter is rotating with the earth as we make
our gravity reading. Because the earth rotates on an axis passing through
the poles at a rate of once a day and our gravimeter is resting on the earth
as the reading is made, the gravity reading contains information related to
the earths rotation.
We know that if a body rotates, it experiences an outward directed
force known as a centrifugal force. The size of this force is proportional to
the distance from the axis of rotation and the rate at which the rotation
is occurring. For our gravimeter located on the surface of the earth, the
rate of rotation does not vary with position, but the distance between the
rotational axis and the gravity meter does vary. The size of the centrifugal
14
You should have noticed by now that expressions like to first order or to a first
order approximation have been used rather frequently in this discussion. But, what do
they mean? Usually, this implies that when considering a specific phenomena that could
have several root causes, we are considering only those that are the most important.
Liner Env Geophysics Reader 40

force is relatively large at the equator and goes to zero at the poles. The
direction this force acts is always away from the axis of rotation. Therefore,
this force acts to reduce the gravitational acceleration we would observe at
any point on the earth, from that which would be observed if the earth were
not rotating.

Figure 23: A vertical slice through the earth.

Correcting for Latitude Dependent Changes Correcting observa-


tions of the gravitational acceleration for latitude dependent variations aris-
ing from the earths elliptical shape and rotation is relatively straight for-
ward. By assuming the earth is elliptical with the appropriate demensions,
is rotating at the appropriate rate, and contains no lateral variations in
geologic structure (that is, contains no interesting geologic structure), we
can derive a mathematical formulation for the earths gravitational acceler-
ation that depends only on the latitude of the observation. By subtracting
the gravitational acceleration predicted by this mathematical formulation
from the observed gravitational acceleration, we can effectively remove from
the observed acceleration those portions related to the earths shape and
Liner Env Geophysics Reader 41

rotation.
The mathematical formula used to predict the components of the grav-
itational acceleration produced by the earths shape and rotation is called
the Geodetic Reference Formula of 1967. The predicted gravity is called the
normal gravity,

gn = 978.03185(1.0 + 0.005278895 sin2 0.00023462 sin4 ) , (6)

where gn is the normal gravity15 in cm/s2 and is latitude.


How large is this correction compared to our observed gravitational ac-
celeration? And, because we need to know the latitudes of our observation
points to make this correction, how accurately do we need to know loca-
tions? At a latitude of 45 degrees, the gravitational acceleration varies
approximately 0.81 mgals per kilometer. Thus, to achieve an accuracy of
0.01 mgals, we need to know the north-south location of our gravity stations
to about 12 meters.

Variation in Gravitational Acceleration Due to Changes in Eleva-


tion In Figure 24, imagine two gravity readings taken at the same location
and at the same time with two perfect (no instrument drift and the read-
ings contain no errors) gravimeters; one placed on the ground, the other
place on top of a step ladder. Would the two instruments record the same
gravitational acceleration?
No, the instrument placed on top of the step ladder would record a
smaller gravitational acceleration than the one placed on the ground. Why?
Remember that the size of the gravitational acceleration changes as the
gravimeter changes distance from the center of the earth. In particular, the
size of the gravitational acceleration varies as one over the distance squared
between the gravimeter and the center of the earth. Therefore, the gravime-
ter located on top of the step ladder will record a smaller gravitational
15
Gravitational acceleration expected for a rotating ellipsoidal earth without any geo-
logic complications and no surface features.
Liner Env Geophysics Reader 42

acceleration, because it is positioned farther from the earths center than


the gravimeter resting on the ground.
Therefore, when interpreting data from our gravity survey, we need to
make sure that we dont interpret spatial variations in gravitational accel-
eration that are related to elevation differences in our observation points as
being due to subsurface geology. Clearly, to be able to separate these two
effects, we are going to need to know the elevations at which our gravity
observations are taken.

Figure 24: Observed gravity depends on elevation of the instrument

Accounting for Elevation Variations: The Free-Air Correction To


account for variations in the observed gravitational acceleration that are re-
lated to elevation variations, we incorporate another correction to our data
known as the Free-Air Correction. In applying this correction, we math-
ematically convert our observed gravity values to ones that look like they
were all recorded at the same elevation, thus further isolating the geological
component of the gravitational field.
To a first-order approximation, the gravitational acceleration observed
Liner Env Geophysics Reader 43

on the surface of the earth varies at about -0.3086 mgal per meter in elevation
difference. The minus sign indicates that as the elevation increases, the
observed gravitational acceleration decreases. The magnitude of the number
says that if two gravity readings are made at the same location, but one is
done a meter above the other, the reading taken at the higher elevation will
be 0.3086 mgal less than the lower. Compared to size of the gravity anomaly
computed from the simple model of an ore body, 0.025 mgal, the elevation
effect is huge!
To apply an elevation correction to our observed gravity, we need to
know the elevation of every gravity station. If this is known, we can correct
all of the observed gravity readings to a common elevation16 (usually chosen
to be sea level) by adding -0.3086 times the elevation of the station in meters
to each reading. Given the relatively large size of the expected corrections,
how accurately do we actually need to know the station elevations?
If we require a precision of 0.01 mgals, then relative station elevations
need to be known to about 3 cm. To get such a precision requires very
careful location surveying to be done. In fact, one of the primary costs of a
high-precision gravity survey is in obtaining the relative elevations needed
to compute the Free-Air correction.

Variations in Gravity Due to Excess Mass The free-air correction


accounts for elevation differences between observation locations. Although
observation locations may have differing elevations, these differences usually
result from topographic changes along the earths surface. Thus, unlike
the motivation given for deriving the elevation correction, the reason the
elevations of the observation points differ is because additional mass has been
placed underneath the gravimeter in the form of topography. Therefore, in
addition to the gravity readings differing at two stations because of elevation
differences, the readings will also contain a difference because there is more
16
This common elevation to which all of the observations are corrected to is usually
referred to as the datum elevation.
Liner Env Geophysics Reader 44

mass below the reading taken at a higher elevation than there is of one taken
at a lower elevation.
As a first-order correction for this additional mass, we will assume that
the excess mass underneath the observation point at higher elevation, point
B in Figure 25, can be approximated by a slab of uniform density and thick-
ness. Obviously, this description does not accurately describe the nature of
the mass below point B. The topography is not of uniform thickness around
point B and the density of the rocks probably varies with location. At this
stage, however, we are only attempting to make a first-order correction.
More detailed corrections will be considered next.

Figure 25: Replacement of topography by a uniform slab.


Liner Env Geophysics Reader 45

Correcting for Excess Mass: The Bouguer Slab Correction Al-


though there are obvious shortcomings to the simple slab approximation to
elevation and mass differences below gravity stations, it has two distinct
advantages over more complex (realistic) models.

Because the model is so simple, it is rather easy to construct predic-


tions of the gravity produced by it and make an initial, first-order
correction to the gravity observations for elevation and excess mass.

Because gravitational acceleration varies as one over the distance to


the source of the anomaly squared and because we only measure the
vertical component of gravity, most of the contributions to the gravity
anomalies we observe on our gravimeter are directly under the meter
and rather close to the meter. Thus, the flat slab assumption can
adequately describe much of the gravity anomalies associated with
excess mass and elevation.

Corrections based on this simple slab approximation are referred to as


the Bouguer Slab Correction. It can be shown that the vertical gravi-
tational acceleration associated with a flat slab can be written simply as
0.04193 h. Where the correction is given in mgals, is the density of
the slab in gm/cm3 , and h is the elevation difference in meters between the
observation point and elevation datum. h is positive for observation points
above the datum level and negative for observation points below the datum
level.
Notice that the sign of the Bouguer Slab Correction makes sense. If an
observation point is at a higher elevation than the datum, there is excess
mass below the observation point that wouldnt be there if we were able
to make all of our observations at the datum elevation. Thus, our grav-
ity reading is larger due to the excess mass, and we would therefore have
to subtract a factor to move the observation point back down to the da-
Liner Env Geophysics Reader 46

tum. Notice that the sign of this correction is opposite to that used for the
elevation correction.
Also notice that to apply the Bouguer Slab correction we need to know
the elevations of all of the observation points and the density of the slab
used to approximate the excess mass. In choosing a density, use an average
density for the rocks in the survey area. For a density of 2.67 gm/cm3 , the
Bouguer Slab Correction is about 0.11 mgals/m.

Variations in Gravity Due to Nearby Topography Although the


slab correction described previously adequately describes the gravitational
variations caused by gentle topographic variations (those that can be approx-
imated by a slab), it does not adequately address the gravitational variations
associated with extremes in topography near an observation point. Consider
the gravitational acceleration observed at point B shown in Figure 26.
In applying the slab correction to observation point B, we remove the
effect of the mass surrounded by the blue rectangle. Note, however, that in
applying this correction in the presence of a valley to the left of point B, we
have accounted for too much mass because the valley actually contains no
material. Thus, a small adjustment must be added back into our Bouguer
corrected gravity to account for the mass that was removed as part of the
valley and, therefore, actually didnt exist.
The mass associated with the nearby mountain is not included in our
Bouguer correction. The presence of the mountain acts as an upward di-
rected gravitational acceleration. Therefore, because the mountain is near
our observation point, we observe a smaller gravitational acceleration di-
rected downward than we would if the mountain were not there. Like the
valley, we must add a small adjustment to our Bouguer corrected gravity to
account for the mass of the mountain.
These small adjustments are referred to as Terrain Corrections. As noted
above, Terrain Corrections are always positive in value. To compute these
Liner Env Geophysics Reader 47

corrections, we are going to need to be able to estimate the mass of the


mountain and the excess mass of the valley that was included in the Bouguer
Corrections. These masses can be computed if we know the volume of each
of these features and their average densities.

Figure 26: Topographic correction to observed gravity data.

Terrain Corrections Like Bouguer Slab Corrections, when computing


Terrain Corrections we need to assume an average density for the rocks
exposed by the surrounding topography. Usually, the same density is used
for the Bouguer and the Terrain Corrections. Thus far, it appears as though
applying Terrain Corrections may be no more difficult than applying the
Bouguer Slab Corrections. Unfortunately, this is not the case.
To compute the gravitational attraction produced by the topography,
we need to estimate the mass of the surrounding terrain and the distance
of this mass from the observation point (recall, gravitational acceleration
is proportional to mass over the distance between the observation point
and the mass in question squared). The specifics of this computation will
vary for each observation point in the survey because the distances to the
various topographic features varies as the location of the gravity station
moves. As you are probably beginning to realize, in addition to an estimate
of the average density of the rocks within the survey area, to perform this
correction we will need a knowledge of the locations of the gravity stations
and the shape of the topography surrounding the survey area.
Liner Env Geophysics Reader 48

Estimating the distribution of topography surrounding each gravity sta-


tion is not a trivial task. One could imagine plotting the location of each
gravity station on a topographic map, estimating the variation in topo-
graphic relief about the station location at various distances, computing the
gravitational acceleration due to the topography at these various distances,
and applying the resulting correction to the observed gravitational accelera-
tion. A systematic methodology for performing this task was formalized by
Hammer17 in 1939. Using Hammers methodology by hand is tedious and
time consuming. If the elevations surrounding the survey area are avail-
able in computer readable format, computer implementations of Hammers
method are available and can greatly reduce the time required to compute
and implement these corrections.
Although digital topography databases are widely available, they are
commonly not sampled finely enough for computing what are referred to
as the near-zone Terrain Corrections in areas of extreme topographic re-
lief or where high-resolution (less than 0.5 mgals) gravity observations are
required. Near-zone corrections are terrain corrections generated by topog-
raphy located very close (closer than 558 ft) to the station. If the topography
close to the station is irregular in nature, an accurate terrain correction may
require expensive and time-consuming topographic surveying. For example,
elevation variations of as little as two feet located less than 55 ft from the
observing station can produce Terrain Corrections as large as 0.04 mgals.

1.4.4 Summary of Gravity Types

We have now described the host of corrections that must be applied to


our observations of gravitational acceleration to isolate the effects caused
by geologic structure. The wide variety of corrections applied can be a bit
intimidating at first and has led to a wide variety of names used in con-
17
Hammer, Sigmund, 1939, Terrain corrections for gravimeter stations, Geophysics, 4,
184-194.
Liner Env Geophysics Reader 49

junction with gravity observations corrected to various degrees. Lets recap


all of the corrections commonly applied to gravity observations collected for
exploration geophysical surveys, specify the order in which they are applied,
and list the names by which the resulting gravity values go.

Observed Gravity (gobs ) - Gravity readings observed at each gravity


station after corrections have been applied for instrument drift and
tides.

Latitude Correction (gn ) - Correction subtracted from gobs that ac-


counts for the earths elliptical shape and rotation. The gravity value
that would be observed if the earth were a perfect (no geologic or to-
pographic complexities) rotating ellipsoid is referred to as the normal
gravity.

Free Air Corrected Gravity (gf a ) - The Free-Air correction accounts


for gravity variations caused by elevation differences in the observation
locations. The form of the Free-Air gravity anomaly, gf a , is given by;

gf a = gobs gn + 0.3086 h (mgal) (7)

where h is the elevation at which the gravity station is above the


elevation datum chosen for the survey (this is usually sea level).

Bouguer Slab Corrected Gravity (gb ) - The Bouguer correction is a


first-order correction to account for the excess mass underlying obser-
vation points located at elevations higher than the elevation datum.
Conversely, it accounts for a mass deficiency at observations points
located below the elevation datum. The form of the Bouguer gravity
anomaly, gb , is given by;

gb = gobs gn + 0.3086 h 0.04193 h (mgal) (8)

where is the average density of the rocks underlying the survey area.
Liner Env Geophysics Reader 50

Terrain Corrected Bouguer Gravity (gt ) - The Terrain correction ac-


counts for variations in the observed gravitational acceleration caused
by variations in topography near each observation point. The terrain
correction is positive regardless of whether the local topography con-
sists of a mountain or a valley. The form of the Terrain corrected,
Bouguer gravity anomaly, gt, is given by;

gt = gobs gn + 0.3086 h 0.04193 h + T C (mgal) (9)

where T C is the value of the computed Terrain correction.

Assuming these corrections have accurately accounted for the variations


in gravitational acceleration they were intended to account for, any remain-
ing variations in the gravitational acceleration associated with the Terrain
Corrected Bouguer Gravity, gt , can now be assumed to be caused by geologic
structure.

1.5 Isolating Gravity Anomalies of Interest


1.5.1 Local and Regional Gravity Anomalies

In addition to the types of gravity anomalies defined on the amount of pro-


cessing performed to isolate geological contributions, there are also specific
gravity anomaly types defined on the nature of the geological contribution.
To define the various geologic contributions that can influence our gravity
observations, consider collecting gravity observations to determine the ex-
tent and location of a buried, spherical ore body. An example of the gravity
anomaly expected over such a geologic structure has already been shown.
Obviously, this model of the structure of an ore body and the surrounding
geology has been greatly over simplified. Lets consider a slightly more
complicated model for the geology in this problem. For the time being we
will still assume that the ore body is spherical in shape and is buried in
sedimentary rocks having a uniform density. In addition to the ore body,
Liner Env Geophysics Reader 51

lets now assume that the sedimentary rocks in which the ore body resides
are underlain by a denser Granitic basement that dips to the right. This
geologic model and the gravity profile that would be observed over it are
shown in Figure 27
Notice that the observed gravity profile is dominated by a trend indi-
cating decreasing gravitational acceleration from left to right. This trend is
the result of the dipping basement interface. Unfortunately, were not in-
terested in mapping the basement interface in this problem; rather, we have
designed the gravity survey to identify the location of the buried ore body.
The gravitational anomaly caused by the ore body is indicated by the small
hump at the center of the gravity profile.
The gravity profile produced by the basement interface only is shown
in figure 28. Clearly, if we knew what the gravitational acceleration caused
by the basement was, we could remove it from our observations and iso-
late the anomaly caused by the ore body. This could be done simply by
subtracting the gravitational acceleration caused by the basement contact
from the observed gravitational acceleration caused by the ore body and
the basement interface. For this problem, we do know the contribution to
the observed gravitational acceleration from basement, and this subraction
yields the desired gravitational anomaly due to the ore body, Figure 29.
From this simple example you can see that there are two contributions
to our observed gravitational acceleration. The first is caused by large-scale
geologic structure that is not of interest. The gravitational acceleration
produced by these large-scale features is referred to as the Regional Gravity
Anomaly. The second contribution is caused by smaller-scale structure for
which the survey was designed to detect. That portion of the observed
gravitational acceleration associated with these structures is referred to as
the Local or the Residual Gravity Anomaly.
Because the Regional Gravity Anomaly is often much larger in size than
the Local Gravity Anomaly, as in the example shown above, it is imperative
Liner Env Geophysics Reader 52

that we develop a means to effectively remove this effect from our gravity
observations before attempting to interpret the gravity observations for local
geologic structure.

1.5.2 Sources of the Local and Regional Gravity Anomalies

Notice that the Regional Gravity Anomaly is a slowly varying function of


position along the profile line. This feature is a characteristic of all large-
scale sources. That is, sources of gravity anomalies large in spatial extent
(by large we mean large with respect to the profile length) always produce
gravity anomalies that change slowly with position along the gravity profile.
Local Gravity Anomalies are defined as those that change value rapidly along
the profile line. The sources for these anomalies must be small in spatial
extent (like large, small is defined with respect to the length of the gravity
profile) and close to the surface.
As an example of the effects of burial depth on the recorded gravity
anomaly, consider three cylinders all having the same source dimensions and
density contrast with varying depths of burial, Figure 30. For this example,
the cylinders are assumed to be less dense than the surrounding rocks.
Notice that at as the cylinder is buried more deeply, the gravity anomaly
it produces decreases in amplitude and spreads out in width. Thus, the
more shallowly buried cylinder produces a large anomaly that is confined
to a region of the profile directly above the cylinder. The more deeply
buried cylinder produces a gravity anomaly of smaller amplitude that is
spread over more of the length of the profile. The broader gravity anomaly
associated with the deeper source could be considered a Regional Gravity
Contribution. The sharper anomaly associated with the more shallow source
would contribute to the Local Gravity Anomaly.
In this particular example, the size of the regional gravity contribution is
smaller than the size of the local gravity contribution. As you will find from
your work in designing a gravity survey, increasing the radius of the deeply
Liner Env Geophysics Reader 53

Figure 27: Gravity profile over a simple ore body and complicating geologic
structure. Density values shown are relative to sediment density.
Liner Env Geophysics Reader 54

Figure 28: Gravity profile due to the basement structure alone the regional
gravity anomaly.

Figure 29: By subtracting the regional gravity anomaly from the original
gravity profile, we generate the gravity anomaly associated with the ore
body.
Liner Env Geophysics Reader 55

buried cylinder will increase the size of the gravity anomaly it produces
without changing the breadth of the anomaly. Thus, regional contributions
to the observed gravity field that are large in amplitude and broad in shape
are assumed to be deep (producing the large breadth in shape) and large in
aerial extent (producing a large amplitude).

1.5.3 Separating Local and Regional Gravity Anomalies

Because Regional Anomalies vary slowly along a particular profile and Lo-
cal Anomalies vary more rapidly, any method that can identify and isolate
slowly varying portions of the gravity field can be used to separate Regional
and Local Gravity Anomalies. The methods generally fall into three broad
categories:

Direct Estimates - These are estimates of the regional gravity anomaly


determined from an independent data set. For example, if your gravity
survey is conducted within the continential US, gravity observations
collected at relatively large station spacings are available from the Na-
tional Geophyiscal Data Center
(http://julius.ngdc.noaa.gov/ngdc.html)
on CD-ROM
(http://www.ngdc.noaa.gov/seg/fliers/se-0703.html).
Using these observations, you can determine how the long-wavelength
gravity field varies around your survey and then remove its contribu-
tion from your data.

Graphical Esimates - These estimates are based on simply plotting the


observations, sketching the interpreters esimate of the regional grav-
ity anomaly, and subtracting the regional gravity anomaly estimate
from the raw observations to generate an estimate of the local gravity
anomaly.

Mathematical Estimates - This represents any of a wide variety of


Liner Env Geophysics Reader 56

Figure 30: Falling body experiment to estimate gravitational acceleration.


Liner Env Geophysics Reader 57

methods for determining the regional gravity contribution from the


collected data through the use of mathematical procedures. Examples
of how this can be done include:

Moving Averages - In this technique, an estimate of the regional


gravity anomaly at some point along a profile is determined by
averaging the recorded gravity values at several nearby points.
Averaging gravity values over several observation points enhances
the long-wavelength contributions to the recorded gravity field
while suppressing the shorter-wavelength contributions.

Function Fitting - In this technique, smoothly varying mathe-


matical functions are fit to the data and used as estimates of the
regional gravity anomaly. The simplest of any number of possible
functions that could be fit to the data is a straight line.

Filtering and Upward Continuation - These are more sophisicated


mathematical techniques for determining the long-wavelength por-
tion of a data set. Those interested in finding out more about
these types of techniques can find descriptions of them in any
introductory geophysical textbook.

1.5.4 Local/Regional Gravity Anomaly Separation Example

As an example of estimating the regional anomaly from the recorded data


and isolating the local anomaly with this estimate consider using a moving
average operator. With this technique, an estimate of the regional grav-
ity anomaly at some point along a profile is determined by averaging the
recorded gravity values at several nearby points. The number of points over
which the average is calculated is referred to as the length of the operator
and is chosen by the data processor. Averaging gravity values over sev-
eral observation points enhances the long-wavelength contributions to the
recorded gravity field while suppressing the shorter-wavelength contribu-
Liner Env Geophysics Reader 58

tions. Consider the sample gravity data shown in Figure 31.


Moving averages can be computed across this data set. To do this the
data processor chooses the length of the moving average operator. That is,
the processor decides to compute the average over 3, 5, 7, 15, or 51 adjacent
points. As you would expect, the resulting estimate of the regional gravity
anomaly, and thus the local gravity anomaly, is critically dependent on this
choice. Shown in Figure 32 are two estimates of the regional gravity anomaly
using moving average operators of lengths 15 and 35.
Depending on the features of the gravity profile the processor wishes to
extract, either of these operators may be appropriate. If we believe, for
example, the gravity peak located at a distance of about 30 on the profile
is a feature related to a local gravity anomaly, notice that the 15 length
operator is not long enough. The average using this operator length almost
tracks the raw data, thus when we subtract the averages from the raw data
to isolate the local gravity anomaly the resulting value will be near zero. The
35 length operator, on the other hand, is long enough to average out the
anomaly of interest, thus isolating it when we subtract the moving average
estimate of the regional from the raw observations.
The residual gravity estimates computed for each moving average oper-
ator are shown in Figure 33.
As expected, few interpretable anomalies exist after applying the 15
point operator. The peak at a distance of 30 has been greatly reduced in
amplitude and other short-wavelength anomalies apparent in the original
data have been effectively removed. Using the 35 length operator, the peak
at a distance of 30 has been successfully isolated and other short-wavelength
anomalies have been enhanced. Data processors and interpreters are free to
choose the operator length they wish to apply to the data. This choice is
based solely on the features they believe represent the local anomalies of
interest. Thus, separation of the regional from the local gravity field is an
interpretive process.
Liner Env Geophysics Reader 59

Although the interpretive nature of the moving average method for esti-
mating the regional gravity contribution is readily apparent, you should be
aware that all of the methods described above require interpreter input of
one form or another. Thus, no matter which method is used to estimate the
regional component of the gravity field, it should always be considered an
interpretational process.

Figure 31: Sample gravity data for seperation of local/regional anomalies.

1.6 Gravity Anomalies Over Bodies With Simple Shapes


1.6.1 Gravity Anomaly Over a Buried Point Mass

Previously we defined the gravitational acceleration due to a point mass as


Gm
gr = (10)
r2
where G is the gravitational constant, m is the mass of the point mass, and r
is the distance between the point mass and our observation point. Figure 34
shows the gravitational acceleration we would observe over a buried point
mass. Notice, the acceleration is highest directly above the point mass and
decreases as we move away from it.
Liner Env Geophysics Reader 60

Figure 32: Two moving average examples calculated from the sample gravity
data.

Figure 33: Falling body experiment to estimate gravitational acceleration.


Liner Env Geophysics Reader 61

Computing the observed acceleration based on the equation given above


is easy and instructive. First, lets derive the equation used to generate the
graph shown above. Let z be the depth of burial of the point mass and
x is the horizontal distance between the point mass and our observation
point. Notice that the gravitational acceleration caused by the point mass
is in the direction of the point mass; that is, its along the vector r. Before
taking a reading, gravity meters are leveled so that they only measure the
vertical component of gravity; that is, we only measure that portion of the
gravitational acceleration caused by the point mass acting in a direction
pointing down. The vertical component of the gravitational acceleration
caused by the point mass can be written in terms of the angle as
Gm
gr = 2 cos . (11)
r
Now, it is inconvenient to have to compute r and for various values of x
before we can compute the gravitational acceleration. Lets now rewrite the
above expression in a form that makes it easy to compute the gravitational
acceleration as a function of horizontal distance x rather than the distance
between the point mass and the observation point r and the angle .
can be written in terms of z and r using the trigonometric relationship
between the cosine of an angle and the lengths of the hypotenuse and the
adjacent side of the triangle formed by the angle,
z
cos = . (12)
r
Likewise, r can be written in terms of x and z using the relationship between
the length of the hypotenuse of a triangle and the lengths of the two other
sides known as Pythagorean Theorem,

r = (x2 + z 2 )1/2 . (13)

Substituting these into our expression for the vertical component of the
gravitational acceleration caused by a point mass, we obtain
Gmz
g = 2 . (14)
(x + z 2 )3/2
Liner Env Geophysics Reader 62

Knowing the depth of burial, z, of the point mass, its mass, m, and the
gravitational constant, G, we can compute the gravitational acceleration we
would observe over a point mass at various distances by simply varying x in
the above expression. An example of the shape of the gravity anomaly we
would observe over a single point mass is shown above.
Therefore, if we thought our observed gravity anomaly was generated by
a mass distribution within the earth that approximated a point mass, we
could use the above expression to generate predicted gravity anomalies for
given point mass depths and masses and determine the point mass depth and
mass by matching the observations with those predicted from our model.
Although a point mass doesnt appear to be a geologically plausible den-
sity distribution, as we will show next, this simple expression for the grav-
itational acceleration forms the basis by which gravity anomalies over any
more complicated density distribution within the earth can be computed.

1.6.2 Gravity Anomaly Over a Buried Sphere

It can be shown that the gravitational attraction of a spherical body of finite


size and mass m is identical to that of a point mass with the same mass m.
Therefore, the expression derived on the previous page for the gravitational
acceleration over a point mass
Gmz
g = . (15)
+ z 2 )3/2
(x2

also represents the gravitational acceleration over a buried sphere. For ap-
plication with a spherical body, it is convenient to rewrite the mass, m, in
terms of the volume and the density contrast of the sphere with the sur-
rounding earth using
4
m = v where v= R3 (16)
3
where v is the volume of the sphere, is the density contrast of the sphere
with the surrounding rock, and R is the radius of the sphere. Thus, the
Liner Env Geophysics Reader 63

Figure 34: Gravitational acceleration over a point mass.

gravitational acceleration over a buried sphere can be written as


4
3R3 G z
g = . (17)
(x2 + z 2 )3/2

Although this expression appears to be more complex than that used to


describe the gravitational acceleration over a buried sphere, the complexity
arises only because weve replaced m with a term that has more elements. In
form, this expression is still identical to the gravitational acceleration over
a buried point mass.
Liner Env Geophysics Reader 64

1.6.3 Model Indeterminancy

We have now derived the gravitational attraction associated with a simple


spherical body. The vertical component of this attraction was shown to be
equal to:
4
m = v where v= R3 (18)
3
Notice that our expression for the gravitational acceleration over a sphere
contains a term that describes the physical parameters of the spherical body;
its radius, R, and its density contrast, , in the form R3 .
R and are two of the parameters describing the sphere that we would
like to be able to determine from our gravity observations (the third is the
depth to the center of the sphere z). That is, we would like to compute
predicted gravitational accelerations given estimates of R and , compare
these to observed values, and then vary R and until the predicted accel-
eration matches the observed acceleration.
This sounds simple enough, but there is a significant problem: there
are an infinite number of combinations of R and that produce exactly
the same gravitational acceleration! For example, lets assume that we have
found values for R and that fit our observations such that R3 = 31.25
Any other combination of values for R and will also fit the observa-
tions as long as R cubed times equals 31.25. Examples of the gravity
observations produced by four of these solutions are shown in Figure 35.
Our inability to uniquely resolve parameters describing a model of the
earth from geophysical observations is not specific to the gravity method but
is present in all geophysical methods. This is referred to using a variety of
expressions: Model Interminancy, Model Equivalence, and Nonuniqueness
to name a few. No matter what it is called, it always means the same
thing; a particular geophysical method can not uniquely define the geologic
structure underlying the survey. Another way of thinking about this problem
is to realize that a model of the geologic structure can uniquely define the
Liner Env Geophysics Reader 65

gravitational field over the structure. The gravitational field, however, can
not uniquely define the geologic structure that produced it.
If this is the case, how do we determine which model is correct? To do
this we must incorporate additional observations on which to base our inter-
pretation. These additional observations presumably will limit the range of
acceptable models we should consider when interpreting our gravity observa-
tions. These observations could include geologic observations or observations
from different types of geophysical surveys.

Figure 35: Gravity anomalies over four spheres with different radius and
density. Due to non-uniqueness, all profiles are identical.
Liner Env Geophysics Reader 66

1.6.4 Gravity Calculations over Bodies with more Complex Shapes

Although it is possible to derive analytic expressions for the computation


of the gravitational acceleration over additional bodies with simple shapes
(cylinders, slabs, etc.), we already have enough information to describe a
general scheme for computing gravity anomalies over bodies with these and
more complex shapes. The basis for this computation lies in the approxi-
mation of a complex body as a distribution of point masses.
Previously, we derived the vertical component of the gravitational accel-
eration due to a point mass with mass m as
Gmz
g = . (19)
(x2 + z 2 )3/2
We can approximate the body with complex shape as a distribution of
point masses. The gravitational attraction of the body is then nothing more
than the sum of the gravitational attractions of all of the individual point
masses as illustrated in Figure 36.
In mathematical notation, this sum can be written as
G m z1 G m z2 G m z3
g = + + +
((x d1 ) + z1 )
2 2 3/2 ((x d2 ) + z2 )
2 2 3/2 ((x d3 )2 + z32 )3/2
(20)
where z represents the depth of burial of each point mass, d represents
the horizontal position of each point mass, and x represents the horizontal
position of the observation point. Only the first three terms have been
written in this equation. There is, in actuality, one term in this expression
for each point mass. If there are N point masses, this equation can be written
more compactly as
X
N
G m zi
g = . (21)
i = 1
((x di )2 + zi2 )3/2

For more detailed information on the computation of gravity anoma-


lies over complex two- and three- dimensional shapes look at the following
references.
Liner Env Geophysics Reader 67

Figure 36: The gravity profile for an arbitray body can be approximated by
considering it to be composed of many point masses.

Talwani, Worzel, and Landisman, Rapid Gravity Computations for


Two-Dimensional Bodies with Application to the Mendocino Subma-
rine Fraction Zone, Journal Geophysical Research, 64, 49-59, 1959.

Talwani, Manik, and Ewing, Rapid Computation of Gravitational At-


traction of Three-Dimensional Bodies of Arbitrary Shape, Geophysics,
25, 203-225, 1960.
68

Part II
Magnetic Surveying
2 Geophysical Surveying Using Magnetic Meth-
ods (T. Boyd)
2.1 Introduction
2.1.1 Historical Overview

Unlike the gravitational observations described in the previous section, man


has been systematically observing the earths magnetic field for almost 500
years. Sir William Gilbert (left) published the first scientific treatise on the
earths magnetic field entitled De magnete. In this work, Gilbert showed that
the reason compass needles point toward the earths north pole is because
the earth itself appears to behave as a large magnet. Gilbert also showed
that the earths magnetic field is roughly equivalent to that which would be
generated by a bar magnet located at the center of the earth and oriented
along the earths rotational axis. During the mid-nineteenth century, Karl
Frederick Gauss confirmed Gilberts observations and also showed that the
magnetic field observed on the surface of the earth could not be caused
by magnetic sources external to the earth, but rather had to be caused by
sources within the earth.
Geophysical exploration using measurements of the earths magnetic field
was employed earlier than any other geophysical technique. von Werde
located deposits of ore by mapping variations in the magnetic field in 1843.
In 1879, Thalen published the first geophysical manuscript entitled The
Examination of Iron Ore Deposits by Magnetic Measurements.
Even to this day, the magnetic methods are one of the most commonly
used geophysical tools. This stems from the fact that magnetic observa-
tions are obtained relatively easily and cheaply and few corrections must
be applied to the observations. Despite these obvious advantages, like the
Liner Env Geophysics Reader 69

gravitational methods, interpretations of magnetic observations suffer from


a lack of uniqueness.

2.1.2 Similarities Between Gravity and Magnetics

Geophysical investigations employing observations of the earths magnetic


field have much in common with those employing observations of the earths
gravitational field. Thus, you will find that your previous exposure to, and
the intuitive understanding you developed from using, gravity will greatly
assist you in understanding the use of magnetics. In particular, some of the
most striking similarities between the two methods include:

Geophysical exploration techniques that employ both gravity and mag-


netics are passive. By this, we simply mean that when using these two
methods we measure a naturally occurring field of the earth: either
the earths gravitational or magnetic fields. Collectively, the gravity
and magnetics methods are often referred to as potential methods 18 ,
and the gravitational and magnetic fields that we measure are referred
to as potential fields.

Identical physical and mathematical representations can be used to


understand magnetic and gravitational forces. For example, the fun-
damental element used to define the gravitational force is the point
mass. An equivalent representation is used to define the force derived
from the fundamental magnetic element. Instead of being called a
point mass, however, the fundamental magnetic element is called a
magnetic monopole. Mathematical representations for the point mass
and the magnetic monopole are identical.
18
The expression potential field refers to a mathematical property of these types of force
fields. Both gravitational and the magnetic forces are known as conservative forces. This
property relates to work being path independent. That is, it takes the same amount of
work to move a mass, in some external gravitational field, from one point to another re-
gardless of the path taken between the two points. Conservative forces can be represented
mathematically by simple scalar expressions known as potentials. Hence, the expression
potential field.
Liner Env Geophysics Reader 70

The acquisition, reduction, and interpretation of gravity and magnetic


observations are very similar.

2.1.3 Differences Between Gravity and Magnetics

Unfortunately, despite these similarities, there are several significant differ-


ences between gravity and magnetic exploration. By-in-large, these differ-
ences make the qualitative and quantitative assessment of magnetic anoma-
lies more difficult and less intuitive than gravity anomalies.

The fundamental parameter that controls gravity variations of interest


to us as exploration geophysicists is rock density. The densities of rocks
and soils vary little from place to place near the surface of the earth.
The highest densities we typically observe are about 3.0 gm/cm3 ,
and the lowest densities are about 1.0 gm/cm3 . The fundamental
parameter controlling the magnetic field variations of interest to us,
magnetic susceptibility, on the other hand, can vary as much as four to
five orders of magnitude19 . This variation is not only present amongst
different rock types, but wide variations in susceptibility also occur
within a given rock type. Thus, it will be extremely difficult with
magnetic prospecting to determine rock types on the basis of estimated
susceptibilities.

Unlike the gravitational force, which is always attractive, the mag-


netic force can be either attractive or repulsive. Thus, mathematically,
monopoles can assume either positive or negative values.

Unlike the gravitational case, single magnetic point sources (monopoles)


can never be found alone in the magnetic case. Rather, monopoles al-
ways occur in pairs. A pair of magnetic monopoles, referred to as
19
One order of magnitude is a factor of ten. Thus, four orders of magnitude represent
a variation of 10 000.
Liner Env Geophysics Reader 71

a dipole, always consists of one positive monopole and one negative


monopole.

A properly reduced gravitational field is always generated by subsur-


face variations in rock density. A properly reduced magnetic field,
however, can have as its origin at least two possible sources. It can be
produced via an induced magnetization, or it can be produced via a
remanent magnetization. For any given set of field observations, both
mechanisms probably contribute to the observed field. It is difficult,
however, to distinguish between these possible production mechanisms
from the field observations alone.

Unlike the gravitational field, which does not change significantly with
time20 , the magnetic field is highly time dependent.

2.1.4 Magnetic Monopoles

Recall that the gravitational force exerted between two point masses of mass
m1 and m2 separated by a distance r is given by Newtons law of gravitation,
which is written as
G m 1 m2
Fg = (22)
r2
where G is the gravitational constant. This law, in words, simply states that
the gravitational force exerted between two bodies decreases as one over the
square of the distance separating the bodies. Since mass, distance, and the
gravitational constant are always positive values, the gravitational force is
always an attractive force.
Charles Augustin de Coulomb, in 1785, showed that the force of attrac-
tion or repulsion between electrically charged bodies and between magnetic
poles also obey an inverse square law like that derived for gravity by Newton.
20
By this we are only referring to that portion of the gravity field produced by the
internal density distribution and not that produced by the tidal or drift components of
the observed field. That portion of the magnetic field relating to internal earth structure
can vary significantly with time.
Liner Env Geophysics Reader 72

To make the measurements necessary to prove this, Coulomb (independent


of John Michell) invented the torsion balance.
The mathematical expression for the magnetic force experienced between
two magnetic monopoles is given by
1 p1 p2
Fm = (23)
r2
where is a constant of proportionality known as the magnetic permeability,
p1 and p2 are the strengths of the two magnetic monopoles, and r is the
distance between the two poles. In form, this expression is identical to
the gravitational force expression written above. There are, however, two
important differences.
Unlike the gravitational constant, G, the magnetic permeability, , is a
property of the material in which the two monopoles, p1 and p2 , are located.
If they are in a vacuum, is called the magnetic permeability of free space.
Unlike M1 and M2 , p1 and p2 can be either positive or negative in sign. If p1
and p2 have the same sign, the force between the two monopoles is repulsive.
If p1 and p2 have opposite signs, the force between the two monopoles is
attractive.

2.1.5 Forces Associated with Magnetic Monopoles

Given that the magnetic force applied to one magnetic monopole by another
magnetic monopole is given by Coulombs equation, what does the force look
like? Assume that there is a negative magnetic pole, p1 < 0.0, located at
a point x = 1 and y = 0. Now, lets take a positive magnetic pole,
p2 > 0.0, and move it to some location (x, y) and measure the strength and
the direction of the magnetic force field. Well plot this force as an arrow in
the direction of the force with a length indicating the strength of the force.
Repeat this by moving the positive pole to a new location. After doing
this at many locations, you will produce a plot similar to the one shown in
Figure 37.
Liner Env Geophysics Reader 73

As described by Coulombs equation, the size of the arrows should de-


crease as one over the square of the distance between the two magnetic
poles21 and the direction of the force acting on p2 is always in the direction
toward p1 (the force is attractive)22 .
If instead p1 is a positive pole located at x = 1, the plot of the magnetic
force acting on p2 is the same as that shown above except that the force is
always directed away from p1 (the force is repulsive), Figure 38.

Figure 37: Forces associated with a negative magnetic pole.

2.1.6 Magnetic Dipoles

So far everything seems pretty simple and directly comparable to gravi-


tational forces, albeit with attractive and repulsive forces existing in the
magnetic case when only attractive forces existed in the gravitational case.
Now things start getting a bit more complicated. The magnetic monopoles
that we have been describing have never actually been observed!!
21
For plotting purposes, the arrow lengths shown in the figures above decay proportional
to one over the distance between the two poles rather than proportional to one over the
square of the distance between the two poles. If the true distance relationship were used,
the lengths of the arrows would decrease so rapidly with distance that it would be difficult
to visualize the distance-force relationship being described.
22
If we were to plot the force of gravitational attraction between two point masses, the
plot would look identical to this.
Liner Env Geophysics Reader 74

Figure 38: Forces associated with a positive magnetic pole.

Rather, the fundamental magnetic element appears to consist of two


magnetic monopoles, one positive and one negative, separated by some dis-
tance. This fundamental magnetic element consisting of two monopoles is
called a magnetic dipole.
Now lets see what the force looks like from this fundamental magnetic
element, the magnetic dipole? Fortunately, we can derive the magnetic force
produced by a dipole by considering the force produced by two magnetic
monopoles. In fact, this is why we began our discussion on magnetism by
looking at magnetic monopoles. If a dipole simply consists of two magnetic
monopoles, you might expect that the force generated by a dipole is simply
the force generated by one monopole added to the force generated by a
second monopole. This is exactly right!!
Earlier, we plotted the magnetic forces associated with two magnetic
monopoles. These are reproduced in Figure 39 as the red and purple arrows.
If we add these forces together using vector addition, we get the green
arrows. These green arrows now indicate the force associated with a mag-
netic dipole consisting of a negative monopole at x = 1, labeled S, and
a positive monopole at x = 1, labeled N. Shown in Figure 40 are the force
arrows for this same magnetic dipole without the red and purple arrows
Liner Env Geophysics Reader 75

indicating the monopole forces.


The force associated with this fundamental element of magnetism, the
magnetic dipole, now looks more complicated than the simple force asso-
ciated with gravity. Notice how the arrows describing the magnetic force
appear to come out of the monopole labeled N and into the monopole labeled
S.
You may recognize this force distribution. It is nothing more than the
magnetic force distribution observed around a simple bar magnet. In fact, a
bar magnet can be thought of as nothing more than two magnetic monopoles
separated by the length of the magnet. The magnetic force appears to
originate out of the north pole,N, of the magnet and to terminate at the
south pole, S, of the magnet.

Figure 39: Force components for two magnetic monopoles (red and purple)
a magnetic dipole (green).

2.1.7 Field Lines for a Magnetic Dipole

Another way to visualize the magnetic force field associated with a magnetic
dipole is to plot the field lines for the force. Field lines are nothing more
than a set of lines drawn such that they are everywhere parallel to the
direction of the force you are trying to describe, in this case the magnetic
Liner Env Geophysics Reader 76

Figure 40: Magnetic dipole force vectors.

force. Shown in Figure 41 is the spatial variation of the magnetic force


(green arrows)23 associated with a magnetic dipole and a set of field lines
(red lines) describing the force.
Notice that the red lines representing the field lines are always parallel
to the force directions shown by the green arrows. The number and spacing
of the red lines that we have chosen to show is arbitrary except for one
factor. The position of the red lines shown has been chosen to qualitatively
indicate the relative strength of the magnetic field. Where adjacent red lines
are closely spaced, such as near the two monopoles (blue and yellow circles)
comprising the dipole, the magnetic force is large. The greater the distance
between adjacent red lines, the smaller the magnitude of the magnetic force.

2.1.8 Units Associated with Magnetic Poles

The units associated with magnetic poles and the magnetic field are a
bit more obscure than those associated with the gravitational field. From
23
Unlike the force plots shown on the previous page, the arrows representing the force
have not been rescaled. Thus, you can now see how rapidly the size of the force decreases
with distance from the dipole. Small forces are represented only by an arrow head that is
constant in size. In addition, please note that the vertical axis in the above plot covers a
distance almost three times as large as the horizontal axis.
Liner Env Geophysics Reader 77

Figure 41: Field lines (red) and force vectors (green) for a magnetic dipole.

Coulombs expression,
1 p1 p2
Fm = (24)
r2
we know that force must be given in Newtons,N , where a Newton is a
kg m/s2 . We also know that distance has the units of meters, m. Perme-
ability, ;, is defined to be a unitless constant. The units of pole strength
are defined such that if the force, F , is 1 N and the two magnetic poles are
separated by 1 m, each of the poles has a strength of 1 Amp m (Ampere-
meters). In this case, the poles are referred to as unit poles.
The magnetic field strength, H, is defined as the force per unit pole
strength exerted by a magnetic monopole, p1 . H is nothing more than
Coulombs expression divided by p2 ,
Fm p1
H= = (25)
p2 r2
Liner Env Geophysics Reader 78

The magnetic field strength H is the magnetic analog to the gravitational


acceleration, g.
Given the units associated with force, N , and magnetic monopoles,
Amp m, the units associated with magnetic field strength are Newtons
per Ampere-meter, N/(Amp m). A N/(Amp m) is referred to as a tesla
(T ), named after the renowned inventor Nikola Tesla.
When describing the magnetic field strength of the earth, it is more
common to use units of nanoteslas (nT ), where one nanotesla is 1 billionth
of a tesla. The average strength of the Earths magnetic field is about
50 000 nT . A nanotesla is also commonly referred to as a gamma.

2.2 Magnetization of Materials


2.2.1 Induced Magnetization

When a magnetic material, say iron, is placed within a magnetic field, H,


the magnetic material will produce its own magnetization. This phenomena
is called induced magnetization.
In practice, the induced magnetic field (that is, the one produced by the
magnetic material) will look like it is being created by a series of magnetic
dipoles located within the magnetic material and oriented parallel to the
direction of the inducing field, H, as shown in Figure 42.
The strength of the magnetic field induced by the magnetic material due
to the inducing field is called the intensity of magnetization, I.

2.2.2 Magnetic Susceptibility

The intensity of magnetization, I, is related to the strength of the inducing


magnetic field, H, through a constant of proportionality,k, known as the
magnetic susceptibility
I = kH . (26)

The magnetic susceptibility is a unitless constant that is determined


by the physical properties of the magnetic material. It can take on either
Liner Env Geophysics Reader 79

Figure 42: Magnetic induction.

positive or negative values. Positive values imply that the induced magnetic
field, I, is in the same direction as the inducing field, H. Negative values
imply that the induced magnetic field is in the opposite direction as the
inducing field.
In magnetic prospecting, the susceptibility is the fundamental material
property whose spatial distribution we are attempting to determine. In this
sense, magnetic susceptibility is analogous to density in gravity surveying.

2.2.3 Mechanisms of Magnetic Induction

The nature of magnetization material is in general complex, governed by


atomic properties, and well beyond the scope of this series of notes. Suf-
fice it to say, there are three types of magnetic materials: paramagnetic,
diamagnetic, and ferromagnetic.

Diamagnetism - Discovered by Michael Faraday in 1846. This form of


magnetism is a fundamental property of all materials and is caused by
the alignment of magnetic moments associated with orbital electrons
Liner Env Geophysics Reader 80

in the presence of an external magnetic field. For those elements with


no unpaired electrons in their outer electron shells, this is the only form
of magnetism observed. The susceptibilities of diamagnetic materials
are relatively small and negative. Quartz and salt are two common
diamagnetic earth materials.

Paramagnetism - This is a form of magnetism associated with elements


that have an odd number of electrons in their outer electron shells.
Paramagnetism is associated with the alignment of electron spin di-
rections in the presence of an external magnetic field. It can only
be observed at relatively low temperatures. The temperature above
which paramagnetism is no longer observed is called the Curie Tem-
perature. The susceptibilities of paramagnetic substances are small
and positive.

Ferromagnetism - This is a special case of paramagnetism in which


there is an almost perfect alignment of electron spin directions within
large portions of the material referred to as domains. Like paramag-
netism, ferromagnetism is observed only at temperatures below the
Curie temperature. There are three varieties of ferromagnetism.

Pure Ferromagnetism - The directions of electron spin alignment


within each domain are almost all parallel to the direction of
the external inducing field. Pure ferromagnetic substances have
large (approaching 1) positive susceptibilities. Ferrromagnetic
minerals do not exist, but iron, cobalt, and nickel are examples
of common ferromagnetic elements.

Antiferromagnetism - The directions of electron alignment within


adjacent domains are opposite and the relative abundance of do-
mains with each spin direction is approximately equal. The ob-
served magnetic intensity for the material is almost zero. Thus,
Liner Env Geophysics Reader 81

the susceptibilities of antiferromagnetic materials are almost zero.


Hematite is an antiferromagnetic material.
Ferromagnetism - Like antiferromagnetic materials, adjacent do-
mains produce magnetic intensities in opposite directions. The
intensities associated with domains polarized in a direction oppo-
site that of the external field, however, are weaker. The observed
magnetic intensity for the entire material is in the direction of
the inducing field but is much weaker than that observed for
pure ferromagnetic materials. Thus, the susceptibilities for ferro-
magnetic materials are small and positive. The most important
magnetic minerals are ferromagnetic and include magnetite, ti-
tanomagnetite, ilmenite, and pyrrhotite.

Figure 43: Concept of a pure ferromagnetic material, examples are the ele-
ments iron, cobalt and nickel.

2.2.4 Suseptibilities of Common Rocks and Minerals

Although the mechanisms by which induced magnetization can arise are


rather complex, the field generated by these mechanisms can be quantified
by a single, simple parameter known as the susceptibility, k. As we will
show below, the determination of a material type through a knowledge of
Liner Env Geophysics Reader 82

Figure 44: Concept of an antiferromagnetic material, an example is the


mineral hematite.

its susceptibility is an extremely difficult proposition, even more so than by


determining a material type through a knowledge of its density.
The susceptibilities of various rocks and minerals24 are shown in Table 2.
Unlike density, notice the large range of susceptibilities not only between
varying rocks and minerals but also within rocks of the same type. It is not
uncommon to see variations in susceptibility of several orders of magnitude
for different igneous rock samples. In addition, like density, there is con-
siderable overlap in the measured susceptibilities. Hence, a knowledge of
susceptibility alone will not be sufficient to determine rock type, and, al-
ternately, a knowledge of rock type is often not sufficient to estimate the
expected susceptibility.
This wide range in susceptibilities implies that spatial variations in the
observed magnetic field may be readily related to geologic structure. Be-
cause variations within any given rock type are also large, however, it will be
difficult to construct corrections to our observed magnetic field on assumed
24
Although susceptibility is unitless, its values differ depending on the unit system
used to quantify H and I. The values given here assume the use of the SI, International
System of Units (Systme International dUnits) based on the meter, kilogram, and second.
Another unit system, the cgs, centimeter, gram, and second system is also commonly used.
To convert the SI units for susceptibility given above to cgs, divide by 4.
Liner Env Geophysics Reader 83

Figure 45: Concept of a ferrimagnetic material, the most important example


is the mineral magnetite.

susceptibilities as was done in constructing some of the fundamental gravi-


tational corrections (Bouguer slab correction and Topographic corrections).

2.2.5 Remanent Magnetism

So, as weve seen, if we have a magnetic material and place it in an external


magnetic field (one that weve called the inducing field), we can make the
magnetic material produce its own magnetic field. If we were to measure
the total magnetic field near the material, that field would be the sum
of the external, or inducing field, and the induced field produced in the
material. By measuring spatial variations in the total magnetic field and by
knowing what the inducing field looks like, we can, in principle, map spatial
variations in the induced field and from this determine spatial variations in
the magnetic susceptibility of the subsurface.
Although this situation is a bit more complex than the gravitational
situation, its still manageable. There is, however, one more complication
in nature concerning material magnetism that we need to consider. In the
scenerio weve been discussing, the induced magnetic field is a direct con-
sequence of a magnetic material being surrounded by an inducing magnetic
field. If you turn off the inducing magnetic field, the induced magnetization
Liner Env Geophysics Reader 84

Material Susceptibility x103 (SI)


Air 0
Quartz 0.01
Rock Salt 0.01
Calcite 0.001 0.01
Sphalerite 0.4
Pyrite 0.05 5
Hematite 0.5 35
Illmenite 300 3500
Magnetite 1200 19 200
Limestones 03
Sandstones 0 20
Shales 0.01 15
Schist 0.3 3
Gneiss 0.1 25
Slate 0 35
Granite 0 50
Gabbro 1 90
Basalt 0.2 175
Peridotite 90 200

Table 2: Magnetic susceptibility range of typical earth materials.

disappears. Or does it?


If the magnetic material has relatively large susceptibilities, or if the
inducing field is strong, the magnetic material will retain a portion of its in-
duced magnetization even after the induced field disappears. This remaining
magnetization is called Remanent Magnetization.
Remanent Magnetization is the component of the materials magnetiza-
tion that solid-earth geophysicists use to map the motion of continents and
ocean basins resulting from plate tectonics. Rocks can acquire a remanent
magnetization through a variety of processes that we dont need to discuss
in detail. A simple example, however, will illustrate the concept. As a vol-
canic rock cools, its temperature decreases past the Curie Temperature. At
the Curie Temperature, the rock, being magnetic, begins to produce an in-
Liner Env Geophysics Reader 85

duced magnetic field. In this case, the inducing field is the Earths magnetic
field. As the Earths magnetic field changes with time, a portion of the in-
duced field in the rock does not change but remains fixed in a direction and
strength reflective of the Earths magnetic field at the time the rock cooled
through its Curie Temperature. This is the remanent magnetization of the
rockthe recorded magnetic field of the Earth at the time the rock cooled
past its Curie Temperature.
The only way you can measure the remanent magnetic component of a
rock is to take a sample of the rock back to the laboratory for analysis. This
is time consuming and expensive. As a result, in exploration geophysics,
we typically assume there is no remanent magnetic component in the ob-
served magnetic field. Clearly, however, this assumption is wrong and could
possibly bias our interpretations.

2.3 The Earths Magnetic Field


2.3.1 Magnetic Field Nomenclature

As you can see, although we started by comparing the magnetic field to the
gravitational field, the specifics of magnetism are far more complex than
gravitation. Despite this, it is still useful to start from the intuition you
have gained through your study of gravitation when trying to understand
magnetism. Before continuing, however, we need to define some of the
relevant terms we will use to describe the Earths magnetic field.
When discussing gravity, we really didnt talk much about how we de-
scribe gravitational acceleration. To some extent, this is because such a
description is almost obvious; gravitational accleration has some size (mea-
sured in geophysics with a gravimeter in mgals), and it is always acting
downward (in fact, it is how we define down). Because the magnetic field
does not act along any such easily definable direction, earth scientists have
developed a nomenclature to describe the magnetic field at any point on the
Earths surface.
Liner Env Geophysics Reader 86

At any point on the Earths surface, the magnetic field25 , F , has some
strength and points in some direction. The following terms, see Figure 46,
are used to describe the direction of the magnetic field.

Declination - The angle between north and the horizontal projection


of F . This value is measured positive through east and varies from 0
to 360 degrees.

Inclination - The angle between the surface of the earth and F . Posi-
tive inclinations indicate F is pointed downward, negative inclinations
indicate F is pointed upward. Inclination varies from -90 to 90 degrees.

Magnetic Equator - The location around the surface of the Earth where
the Earths magnetic field has an inclination of zero (the magnetic
field vector F is horizontal). This location does not correspond to the
Earths rotational equator.

Magnetic Poles - The locations on the surface of the Earth where the
Earths magnetic field has an inclination of either plus or minus 90
degrees (the magnetic field vector F is vertical). These locations do
not correspond to the Earths north and south poles.

2.3.2 The Earths Main Field

Ninety percent of the Earths magnetic field looks like a magnetic field that
would be generated from a dipolar magnetic source located at the center of
the Earth and aligned with the Earths rotational axis. This first order de-
scription of the Earths magnetic field was first given by Sir William Gilbert
25
In this context, and throughout the remainder of these notes, F includes contributions
from the Earths main magnetic field (the inducing field), induced magnetization from
crustal sources, and any contributions from sources external to the Earth. The main
magnetic field refers to that portion of the Earths magnetic field that is believed to be
generated within the Earths core. It constitutes the largest portion of the magnetic field
and is the field that acts to induce magnetization in crustal rocks that we are interested
in for exploration applications.
Liner Env Geophysics Reader 87

Figure 46: Definition and orientation of magnetic field components.

in 1600. The strength of the magnetic field at the poles is about 60 000 nT .
If this dipolar description of the field were complete, then the magnetic
equator would correspond to the Earths equator and the magnetic poles
would correspond to the geographic poles. Alas, as weve come to expect
from magnetism, such a simple description is not sufficient for analysis of
the Earths magnetic field.
The remaining 10% of the magnetic field can not be explained in terms of
simple dipolar sources. Complex models of the Earths magnetic field have
been developed and are available. Shown in Figure 47 is a sample of one of
these models generated by the USGS. The plot shows a map of declinations
for a model of the magnetic field as it appeared in the year 199526 .
If the Earths field were simply dipolar with the axis of the dipole oriented
along the Earths rotational axis, all declinations would be 0 degrees (the
field would always point toward the north). As can be seen, the observed
declinations are quite complex.
As observed on the surface of the earth, the magnetic field can be broken
into three separate components.

Main Field - This is the largest component of the magnetic field and
is believed to be caused by electrical currents in the Earths fluid outer
26
As well describe later, another potential complication in using magnetic observations
is that the Earths magnetic field changes with time!
Liner Env Geophysics Reader 88

core. For exploration work, this field acts as the inducing magnetic
field.

External Magnetic Field - This is a relatively small portion of the ob-


served magnetic field that is generated from magnetic sources external
to the earth. This field is believed to be produced by interactions
of the Earths ionosphere with the solar wind. Hence, temporal varia-
tions associated with the external magnetic field are correlated to solar
activity.

Crustal Field - This is the portion of the magnetic field associated


with the magnetism of crustal rocks. This portion of the field contains
both magnetism caused by induction from the Earths main magnetic
field and from remanent magnetization.

The figure shown above was constructed to emphasize characteristics


of the main magnetic field. Although this portion of the field is in itself
complex, it is understood quite well. Models of the main field are available
and can be used for data reduction.

2.3.3 Magnetics and Geology - A Simple Example

This is all beginning to get a bit complicated. What are we actually going to
observe, and how is this related to geology? The portion of the magnetic field
that we have described as the main magnetic field is believed to be generated
in the Earths core. There are a variety of reasons why geophysicists believe
that the main field is being generated in the Earths core, but these are not
important for our discussion. In addition to these core sources of magnetism,
rocks exist near the Earths surface that are below their Curie temperature
and as such can exhibit induced as well as remanent magnetization27 .
27
We will assume that there is no remanent magnetization throughout the remainder of
this discussion.
Liner Env Geophysics Reader 89

Figure 47: World map showing declination of the geomagnetic field in 1995.

Therefore, if we were to measure the magnetic field along the surface of


the earth, we would record magnetization due to both the main and induced
fields. The induced field is the one of interest to us because it relates to the
existence of rocks of high or low magnetic susceptibility near our instrument.
If our measurements are taken near rocks of high magnetic susceptibility, we
will, in general28 , record magnetic field strengths that are larger than if our
measurements were taken at a great distance from rocks of high magnetic
susceptibility. Hence, like gravity, we can potentially locate subsurface rocks
having high magnetic susceptibilities by mapping variations in the strength
of the magnetic field at the Earths surface.
Consider the example shown in Figure 48. Suppose we have a buried
dyke with a susceptibility of 0.001 surrounded by sedimentary rocks with
28
Unlike gravity, magnetic anomalies are rather complex in shape and making sweeping
statements like this can be very dangerous.
Liner Env Geophysics Reader 90

no magnetic susceptibility. The dyke in this example is 3 meters wide, is


buried 5 meters deep, and trends to the northeast. To find the dyke, we
could measure the strength of the magnetic field (in this case along an east-
west trending line). As we approach the dyke, we would begin to observe the
induced magnetic field associated with the dyke in addition to the Earths
main field. Thus, we could determine the location of the dyke and possibly
its dimensions by measuring the spatial variation in the strength of the
magnetic field.
There are several things to notice about the magnetic anomaly produced
by this dyke.

Like a gravitational anomaly associated with a high-density body, the


magnetic anomaly associated with the dyke is localized to the region
near the dyke. The size of the anomaly rapidly decays with distance
away from the dyke.

Unlike the gravity anomaly we would expect from a higher-density


dyke, the magnetic anomaly is not symmetric about the dykes mid-
point which is at a distance of zero in the above example. Not only
is the anomaly shaped differently to the left and to the right of the
dyke, but the maximum anomaly is not centered at the center of the
dyke. These observations are in general true for all magnetic anoma-
lies. The specifics of this generalization, however, will depend on the
shape and orientation of the magnetized body, its location (bodies of
the same shape and size will produce different anomalies when located
at different places), and the direction in which the profile is taken.

The size of the anomaly produced by this example is about 40 nT.


This is a pretty good- sized anomaly. It is not uncommon to look
for anomalies as small as a few nT. Thus, we must develop surveying
techniques to reduce systematic and random errors to smaller than a
few nT.
Liner Env Geophysics Reader 91

Figure 48: A magnetic anomaly example.

2.3.4 Temporal Variations of the Earths Main Field - Overview

Like the gravitational field, the magnetic field varies with time. When de-
scribing temporal variations of the magnetic field, it is useful to classify these
variations into one of three types depending on their rate of occurence and
source. Please note explicitly that the temporal variations in the magnetic
field that we will be discussing are those that have been observed directly
during human history. As such, the most well-known temporal variation,
magnetic polarity reversals, while important in the study of earth history,
will not be considered in this discussion. We will, however, consider the
following three temporal variations:
Liner Env Geophysics Reader 92

Secular Variations - These are long-term (changes in the field that


occur over years) variations in the main magnetic field that are pre-
sumably caused by fluid motion in the Earths Outer Core. Because
these variations occur slowly with respect to the time of completion of
a typical exploration magnetic survey, these variations will not com-
plicate data reduction efforts.

Diurnal Variations - These are variations in the magnetic field that


occur over the course of a day and are related to variations in the
Earths external magnetic field. This variation can be on the order
of 20 to 30 nT per day and should be accounted for when conducting
exploration magnetic surveys.

Magnetic Storms - Occasionally, magnetic activity in the ionosphere


will abrubtly increase. The occurrence of such storms correlates with
enhanced sunspot activity. The magnetic field observed during such
times is highly irregular and unpredictable, having amplitudes as large
as 1000 nT. Exploration magnetic surveys should not be conducted
during magnetic storms.

2.3.5 Secular Variations

The fact that the Earths magnetic field varies with time was well established
several centuries ago. In fact, this is the primary reason that permanent
magnetic observatories were established from which we have learned how
the magnetic field has changed over the past few centuries. Many sources
of historical information are available.
Shown in Figure 49 is a plot of the declination and inclination of the
magnetic field around Britain from the years 1500 through 1900.
At this one location, you can see that over the past 400 years, the dec-
lination has varied by almost 37 degrees while the inclination has varied
by as much as 13 degrees. These changes are generally assumed to be as-
Liner Env Geophysics Reader 93

sociated with the Earths main magnetic field. That is, these are changes
associated with that portion of the magnetic field believed to be generated
in the Earths core. As such, solid earth geophysicists are very interested in
studying these secular variations, because they can be used to understand
the dynamics of the Earths core.
To understand these temporal variations and to quantify the rate of vari-
ability over time, standard reference models are constructed from magnetic
observatory observations about every five years. One commonly used set
of reference models is known as the International Geomagnetic Reference
Field. Based on these models, it is possible to predict the portion of the
observed magnetic field associated with the Earths main magnetic field at
any point on the Earths surface, both now and for several decades in the
past.
Because the main magnetic field as described by these secular variations
changes slowly with respect to the time it takes us to complete our explo-
ration magnetic survey, this type of temporal variation is of little importance
to us.

Figure 49: Time variation of the declination and inclination of the earths
magnetic field.
Liner Env Geophysics Reader 94

2.3.6 Diurnal Variations

Of more importance to exploration geophysical surveys are the daily, or di-


urnal, variations of the Earths magnetic field. These variations were first
discovered in 1722 in England when it was also noted that these daily vari-
ations were larger in summer than in winter.
The plot in Figure 50 shows typical variations in the magnetic data
recorded at a single location (Boulder, Colorado) over a time period of two
days. Although there are high-frequency components to this variation, no-
tice that the dominant trend is a slowly varying component with a period
of about 24 hours. In this location, at this time, the amplitude of this daily
variation is about 20 nT.
These variations are believed to be caused by electric currents induced
in the Earth from an external source. In this case, the external source is
believed to be electric currents in the upper atmosphere, or the ionosphere.
These electric currents in the ionosphere are in turn driven by solar activity.
Given the size of these variations, the size of the magnetic anomalies we
would expect in a typical geophysical survey, and the fact that surveys could
take several days or weeks to complete, it is clear that we must account for
diurnal variations when interpreting our magnetic data.

2.3.7 Magnetic Storms

In addition to the relatively predictable and smoothly varying diurnal vari-


ations, there can be transient, large amplitude (up to 1000 nT) variations
in the field that are referred to as Magnetic storms. The frequency of these
storms correlates with sunspot activity. Based on this, some prediction of
magnetic storm activity is possible. The most intense storms can be ob-
served globally and may last for several days.
Figure51 shows the magnetic field recorded at a single location during
such a transient event. Although the magnetic storm associated with this
event is not particularly long-lived, notice that the size of the magnetic field
Liner Env Geophysics Reader 95

Figure 50: Time variation of the magnetic field at a fixed location.

during this event varies by almost 100 nT in a time period shorter than 10
minutes!!
Exploration magnetic surveys should not be conducted during times of
magnetic storms. This is simply because the variations in the field that they
can produce are large, rapid, and spatially varying. Therefore, it is difficult
to correct for them in acquired data.

2.4 Magnetometers
2.4.1 Instrumentation Overview

Instruments for measuring aspects of the Earths magnetic field are among
some of the oldest scientific instruments in existence. Magnetic instruments
can be classified into two types.

Mechanical Instruments - These are instruments that are mechanical in


nature that usually measure the attitude (its direction or a component
of its direction) of the magnetic field. The most common example of
this type of instrument is the simple compass. The compass consists
of nothing more than a small test magnet that is free to rotate in
Liner Env Geophysics Reader 96

Figure 51: Field variation during a magnetic storm can be rapid and dra-
matic.

the horizontal plane. Because the positive pole of the test magnet
is attracted to the Earths negative magnetic pole and the negative
pole of the test magnet is attracted to the Earths positive magnetic
pole, the test magnet will align itself along the horizontal direction
of the Earths magnetic field. Thus, it provides measurements of the
declination of the magnetic field. The earliest known compass was
invented by the Chinese no later than the first century A.D., and
more likely as early as the second century B.C.

Although compasses are the most common type of mechanical device


used to measure the horizontal attitude of the magnetic field, other de-
vices have been devised to measure other components of the magnetic
field. Most common among these are the dip needle and the torsion
magnetometer. The dip needle, as its name implies, is used to measure
the inclination of the magnetic field. The torsion magnetometer is a
devise that can measure, through mechanical means, the strength of
the vertical component of the magnetic field.

Magnetometers - Magnetometers are instruments, usually operating


Liner Env Geophysics Reader 97

non-mechanically, that are capable of measuring the strength, or a


component of the strength, of the magnetic field. The first advances
in designing these instruments were made during WWII when Fluxgate
Magnetometers were developed for use in submarine detection.

Since that time, several other magnetometer designs have been de-
veloped that include the Proton Precession and Alkali-Vapor magne-
tometers.

In the following discussion, we will describe only the fluxgate and the
proton precession magnetometers, because they are the most commonly used
magnetometers in exploration surveys.

2.4.2 Fluxgate Magnetometers

The fluxgate magnetometer was originally designed and developed during


World War II. It was built for use as a submarine detection device for low-
flying aircraft. Today it is used for conducting magnetic surveys from air-
craft and for making borehole measurements. A schematic of the fluxgate
magnetometer is shown in Figure 52.
The fluxgate magnetometer is based on what is referred to as the mag-
netic saturation circuit. Two parallel bars of a ferromagnetic material are
placed closely together. The susceptibility of the two bars is large enough so
that even the Earths relatively weak magnetic field can produce magnetic
saturation29 in the bars.
Each bar is wound with a primary coil, but the direction in which the
coil is wrapped around the bars is reversed. An alternating current (AC)
is passed through the primary coils causing a large, inducing magnetic field
29
Magnetic saturation refers to the induced magnetic field produced in the bars. In
general, as the magnitude of the inducing field increases, the magnitude of the induced
field increases in the same proportion as given by our mathematical expression relating the
external to the induced magnetic fields. For large external field strengths, however, this
simple relationship between the inducing and the induced field no longer holds. Saturation
occurs when increases in the strength of the inducing field no longer produce larger induced
fields.
Liner Env Geophysics Reader 98

that produces induced magnetic fields in the two cores that have the same
strengths but opposite orientations.
A secondary coil surrounds the two ferromagnetic cores and the primary
coil. The magnetic fields induced in the cores by the primary coil produce
a voltage potential in the secondary coil. In the absence of an external
field (i.e., if the earth had no magnetic field), the voltage detected in the
secondary coil would be zero because the magnetic fields generated in the
two cores have the same strength but are in opposite directions (their affects
on the secondary coil exactly cancel).
If the cores are aligned parallel to a component of a weak, external
magnetic field, one core will produce a magnetic field in the same direction
as the external field and reinforce it. The other will be in opposition to the
field and produce an induced field that is smaller. This difference is sufficient
to induce a measureable voltage in the secondary coil that is proportional
to the strength of the magnetic field in the direction of the cores.
Thus, the fluxgate magnetometer is capable of measuring the strength
of any component of the Earths magnetic field by simply re-orienting the
instrument so that the cores are parallel to the desired component. Flux-
gate magnetometers are capable of measuring the strength of the magnetic
field to about 0.5 to 1.0 nT. These are relatively simple instruments to con-
struct, hence they are relatively inexpensive ($5,000 - $10,000). Unlike the
commonly used gravimeters, fluxgate magnetometers show no appreciable
instrument drift with time.

2.4.3 Proton Precession Magnetomenters

For land-based magnetic surveys, the most commonly used magnetometer is


the proton precession magnetometer. Unlike the fluxgate magnetometer, the
proton precession magnetometer only measures the total size of the Earths
magnetic field. These types of measurements are usually referred to as total
field measurements. A schematic of the proton precession magnetometer is
Liner Env Geophysics Reader 99

Figure 52: Fluxgate magnetometer.

shown in Figure ??.


The sensor component of the proton precession magnetometer is a cylin-
drical container filled with a liquid rich in hydrogen atoms surrounded by
a coil. Commonly used liquids include water, kerosene, and alcohol. The
sensor is connected by a cable to a small unit in which is housed a power
supply, an electronic switch, an amplifier, and a frequency counter.
When the switch is closed, a DC current delivered by a battery is directed
through the coil, producing a relatively strong magnetic field in the fluid-
filled cylinder. The hydrogen nuclei (protons), which behave like minute
Liner Env Geophysics Reader 100

spinning dipole magnets, become aligned along the direction of the applied
field (i.e., along the axis of the cylinder). Power is then cut to the coil by
opening the switch. Because the Earths magnetic field generates a torque
on the aligned, spinning hydrogen nuclei, they begin to precess30 around
the direction of the Earths total field. This precession induces a small
alternating current in the coil. The frequency of the AC current is equal to
the frequency of precession of the nuclei. Because the frequency of precession
is proportional to the strength of the total field and because the constant
of proportionality is well known, the total field strength can be determined
quite accurately.
Like the fluxgate magnetometer, the proton precession magnetometer is
relatively easy to construct. Thus, it is also relatively inexpensive ($5,000 -
$10,000). The strength of the total field can be measured down to about 0.1
nT. Like fluxgate magnetometers, proton precession magnetometers show
no appreciable instrument drift with time.
One of the important advantages of the proton precession magnetometer
is its ease of use and reliability. Sensor orientation need only be set to a
high angle with respect to the Earths magnetic field. No precise leveling or
orientation is needed. If, however, the magnetic field changes rapidly from
place to place (larger than about 600 nT/m), different portions of the cylin-
drical sensor will be influenced by magnetic fields of various magnitudes, and
readings will be seriously degraded. Finally, because the signal generated by
precession is small, this instrument can not be used near AC power sources.

2.4.4 Total Field Measurements

Given the ease of use of the proton precession magnetometer, most explo-
ration geophysical surveys employ this instrument and thus measure only
30
Precession is motion like that experienced by a top as it spins. Because of the Earths
gravitational field, a spinning top not only spins about its axis of rotation, but the axis
of rotation rotates about vertical. This rotation of the tops spin axis is referred to as
precession.
Liner Env Geophysics Reader 101

Figure 53: Proton precession magnetometer.

the magnitude of the total magnetic field as a function of position. Surveys


conducted using the proton precession magnetometer do not have the ability
to determine the direction of the total field as a function of location.
Ignoring for the moment the temporally varying contribution to the
recorded magnetic field caused by the external magnetic field, the magnetic
field we record with our proton precession magnetometer has two compo-
nents:

The main magnetic field, or that part of the Earths magnetic field
generated by deep (outer core) sources. The direction and size of this
component of the magnetic field at some point on the Earths surface
is represented by the vector labeled Fe in the figure.

The anomalous magnetic field, or that part of the Earths magnetic


field caused by magnetic induction of crustal rocks or remanent mag-
Liner Env Geophysics Reader 102

netization of crustal rocks. The direction and size of this component


of the magnetic field is represented by the vector labeled Fa in the
figure.

The total magnetic field we record, labeled Ft in Figure 54, is nothing


more than the sum of Fe and Fa .
Typically, Fe is much larger than Fa , as is shown in the figure (50,000
nT versus 100 nT). If Fe is much larger than Fa , then Ft will point almost
in the same direction as Fe regardless of the direction of Fa . That is because
the anomalous field, Fa , is so much smaller than the main field, Fe , that the
total field, Ft , will be almost parallel to the main field.

Figure 54: Contributing elements for the total magnetic field.

2.5 Field Procedures


2.5.1 Modes of Acquiring Magnetic Observations

Magnetic observations are routinely collected using any one of three different
field operational strategies.

Airborne - Both fluxgate and proton precession magnetometers can be


Liner Env Geophysics Reader 103

mounted within or towed behind aircraft, including helicopters. These


so-called aeromagnetic surveys are rapid and cost effective. When
relatively large areas are involved, the cost of acquiring 1 km of data
from an aeromagnetic survey is about 40addition, data can be obtained
from areas that are otherwise inaccessible. Among the most difficult
problems associated with aeromagnetic surveys is fixing the position of
the aircraft at any time. With the development of realtime, differential
GPS systems, however, this difficulty is rapidly disappearing.

Shipborne - Magnetic surveys can also be completed over water by


towing a magnetometer behind a ship. Obviously, marine magnetic
surveying is slower than airborne surveying. When other geophysical
methods are being conducted by ship, however, it may make sense to
acquire magnetic data simultaneously.

Ground Based - Like gravity surveys, magnetic surveys are also com-
monly conducted on foot or with a vehicle. Ground-based surveys
may be necessary when the target of interest requires more closely-
spaced readings than are possible to acquire from the air. In the next
discussion we will concentrate on ground-based surveys. All of this
discussion, however, could be applied to air- and shipborne surveys
also.

Because magnetic surveying is generally far cheaper than other geophys-


ical methods, magnetic observations are commonly used for reconnaissance.
These surveys can cover large areas and are used to identify the locations of
targets for more detailed investigations. Because of their cost effectiveness,
magnetic surveys usually consist of areal distributions of data instead of sin-
gle lines of data. We will refer to the collection of geophysical observations
over a geographic area as two-dimensional surveys. Data that is collected
along a single line of observations will be referred to as one-dimensional
surveys.
Liner Env Geophysics Reader 104

2.5.2 Assuring High-Quality Observations - Magnetic Cleanli-


ness

When making total field measurements from which estimates of the sub-
surface distribution of magnetic susceptibility or the presence of subsur-
face magnetized bodies are made, it is imperative that factors affecting the
recorded field other than these be eliminated or isolated so that they can be
removed. We have already discussed several of these added complications,
including spatial variations of the Earths main magnetic field and temporal
variations mostly associated with the external magnetic field. In addition
to these factors which we can not control, there are other sources of noise
that we can control.
Because any ferromagnetic substance can produce an induced magnetic
field in the presence of the Earths main field and because modern magne-
tometers are very sensitive (0.1 nT), the field crew running the magnetic
survey must divest itself of all ferrous objects. This includes, but is not lim-
ited to, belt buckles, knives, wire-rimmed glasses, etc. As a result of this,
proton precession magnetometers are typically placed on two to three meter
poles to remove them from potential noise sources worn by the operators,
Figure 55
In addition to noise sources carried by the operators, many sources of
magnetic noise may be found in the environment. These can include any
ferrous objects such as houses, fences, railroad rails, cars, rebar in concrete
foundations, etc. Finally, when using a proton precession magnetometer,
reliable readings will be difficult to obtain near sources of AC power such as
utility lines and transformers.

2.5.3 Strategies for Dealing with Temporal Variations

Like our gravity observations, magnetic readings taken at the same loca-
tion at different times will not yield the same results. There are temporal
variations in both the Earths magnetic and gravitational fields.
Liner Env Geophysics Reader 105

Figure 55: Operating a proton precession magnetometer. (Figure from In-


troduction to Geophysical Prospecting, M. Dobrin and C. Savit.)

In acquiring gravity observations, we accounted for this temporal vari-


ability by periodically reoccupying a base station and using the variations in
this reading to account for instrument drift and temporal variations of the
field. We could use the same strategy in acquiring magnetic observations
but is not routinely done for the following reasons:

Field variations can be more erratic - Unlike the gravitational field,


Figure 15, the magnetic field can vary quite erratically with time,
as shown in and Figure 50. What this means is that to adequately
approximate the temporal variation in the magnetic field by linearly
interpolating between base station reoccupations, a very short reoccu-
pation time interval may be required. The shorter the reoccupation
interval, the more time is spent at the base station and the longer the
Liner Env Geophysics Reader 106

survey will take to complete.

Cheap Instruments - Unlike gravimeters that can cost more than $25,000,
magnetometers are relatively cheap ($7,500).

Instrument Drift - Unlike gravimeters, magnetometers show no appre-


ciable instrument drift.

With these points in mind, most investigators conduct magnetic surveys


using two magnetometers. One is used to monitor temporal variations of
the magnetic field continuously at a chosen base station, and the other is
used to collect observations related to the survey proper.
By recording the times at which each magnetic station readings are made
and subtracting the magnetic field strength at the base station recorded at
that same time, temporal variations in the magnetic field can be eliminated.
The resulting field then represents relative values of the variation in total
field strength with respect to the magnetic base station.

2.5.4 Spatially Varying Corrections?

When reducing gravity observations, there were a host of spatially varying


corrections that were applied to the data. These included latitude correc-
tions, elevation corrections, slab corrections, and topography corrections.
In principle, all of these corrections could be applied to magnetic observa-
tions also. In practice, the only corrections routinely made for are spatial
variations in the Earths main magnetic field, which would be equivalent to
latitude corrections applied to gravity observations. Why arent the other
corrections applied?
Variations in total field strength as a function of elevation are less than
0.015 nT per meter. This variation is generally considered small enough to
ignore. Variations in total field strength caused by excess magnetic material
(i.e., a slab correction) and topography could, on the other hand, be quite
Liner Env Geophysics Reader 107

significant. The problem is the large variation in susceptibilities associated


with earth materials even when those materials are of the same rock type.
Recall that in applying the slab and elevation corrections to our grav-
itational observations, we had to assume an average density for the rocks
making up the corrections. Rock densities do not vary much from rock type
to rock type. Density variations of 0.5 gm/cm3 are large. Variations among
different samples of the same rock type vary by even less. Therefore, we can
assume an average density for the correction and feel fairly confident that
our assumption is reasonable.
Magnetic susceptibilities vary be orders of magnitude even among sam-
ples of the same rock type. So, how can we choose an average susceptibility
on which to base our correction? The answer is we cant. Therefore, instead
of applying a set of corrections that we know will be wrong, we apply no
correction at all to attempt to account for excess material and topography.

2.5.5 Correcting for the Main-Field Contributions

Corrections for spatial variations in the strength of the Earths main mag-
netic field are referred to as geomagnetic corrections. One commonly used
method of accounting for these variations is to use one of the many mod-
els of the Earths main magnetic field that are available. One such set of
commonly used models of the main field is referred to as the International
Geomagnetic Reference Field (IGRF).
The IGRF models are regularly updated to account for secular variations.
Given the latitude and longitude of some point on the Earths surface, the
total field strength of the Earths main magnetic field can be calculated.
Consider a small two-dimensional survey. A plan view of such a survey is
shown in Figure 56. One commonly used method of applying the main field
correction is to linearly interpolate the computed values of the main field at
the corners of the survey throughout the survey region. These interpolated
values can then be subtracted from the field observations. After applying
Liner Env Geophysics Reader 108

this correction, you are left with that portion of the magnetic field that can
not be attributed to the Earths main magnetic field.
This two-dimensional application of linear interpolation is only slightly
more complex than the one-dimensional linear interpolation used to reduce
our gravity observations. Values of the Earths main magnetic field are first
determined from the IGRF for each corner point of the survey (c1 , c2 , c3 , c4 ).
To determine the strength of the main field at the point p, we first perform
two linear interpolations up the edges of the survey in the y direction to
determine the values of the field at the points t1 and t2 . That is, first
determine the value of the Earths main magnetic field at the point t1 by
linearly interpolating between the points c1 and c4 . Then determine the
value of the main field at the point t2 by linearly interpolating between the
points c2 and c3 . Now, linearly interpolate in the x direction between t1 and
t2 . The result is the two-dimensionally interpolated value of the field at the
point p31 .

2.6 Magnetic Anomalies Over Simple Shapes


2.6.1 Comparison Between Gravity and Magnetic Anomalies

Unlike the gravitational anomalies caused by subsurface density variations,


the magnetic anomalies caused by subsurface variations in magnetic suscep-
tibility are difficult to intuitively construct. This is because there are more
factors involved in controlling the shape of a magnetic anomaly than there
are in controlling the shape of a gravity anomaly.
In the case of a given subsurface density distribution, the shape of the
resulting gravity anomaly is a function of the subsurface density distribution
only. In fact, knowing the gravitational anomaly produced by a simple shape
such as a point mass is often enough to guess what the shape of the gravity
anomaly would be over a much more complicated density distribution. Once
31
There is no reason why I have chosen to first interpolate in the y direction and then
interpolate in the x direction. I could have first interpolated in the x direction between
c1 and c2 and then between c3 and c4.
Liner Env Geophysics Reader 109

Figure 56: Mapview of a 2-D magnetic survey area.

youve determined the shape of the gravity anomaly that the density distri-
bution will produce, then you can make reasonable guesses about how the
anomaly will change as the density constrast is varied or as the depth to the
density contrast is varied. In addition, the anomaly will not change shape
if the density distribution is moved to a different location on the Earth, say
from the equator to the north pole. The gravity anomaly is a function of
density only.
Magnetic anomalies, on the other hand, are a function of two indepen-
dent parameters: the subsurface distribution of susceptibility and the orien-
tation of the Earths main magnetic field. Change one of these parameters
Liner Env Geophysics Reader 110

and you change the resulting magnetic anomaly. What this means in prac-
tice is that magnetic anomalies over the same susceptibility distribution will
be different if the distribution is in a different location, say one located be-
neath the equator versus one located beneath the north pole. Additionally,
the magnetic anomaly over a two-dimensional body such as a tunnel will
look different depending on the orientation of the tunnel, say east-west or
north-south, even if the magnetic profile is always taken perpendicular to
the trend of the tunnel.
With these complexities in mind, we will not spend a great deal of time
analyzing the shapes of magnetic anomalies over simple structures; there
are many computer programs available that do this quite well. Rather, we
will look at several simple examples and qualitatively construct the magnetic
anomalies over them so that you can get a better feeling for the complexities
involved and for how it might be done in the computer.

2.6.2 Magnetic Anomaly: Magnetized Sphere at the North Pole

Lets now qualitatively construct what the magnetic anomaly of a metallic


sphere located beneath the north pole would look like. The geometry of the
sphere, the Earths main magnetic field, the field lines associated with the
anomalous field, the direction and magnitude of the anomalous field, and
a plot of the intensity of the anomalous field that would be recorded are
shown in Figure 57.
At the north (magnetic) pole, the Earths main magnetic field, Fe, points
straight down. Because the buried sphere is composed of a material with a
non-zero susceptibility, the Earths main magnetic field causes the sphere to
produce an induced magnetic field. Field lines associated with this induced
field are shown by black lines, and the magnitude and direction of the in-
duced, anomalous field, Fai, at the surface of the earth are shown by the
blue arrows.
The total field, whose strength will be recorded on a proton precession
Liner Env Geophysics Reader 111

magnetometer, will be sum of the main field, Fe, and the induced, anomalous
field, Fa. Notice that to either side of the sphere, the anomalous field points
in the opposite direction as the main field. Thus, when the main field
is removed from our observations we will observe negative values for the
anomalous field. Near the sphere, the anomalous field points in the same
direction as the main field. Therefore, when the main field is removed, we
will observe positive values for the anomalous field.
In this case, the anomalous magnetic field is symmetric about the center
of the buried sphere, is dominated by a central positive anomaly, and is
surrounded on both sides by smaller negative anomalies.

2.6.3 Magnetic Anomaly: Magnetized Sphere at the Equator

Now, lets examine the shape of the anomalous magnetic field for the exact
same metallic sphere buried at the equator.
At the equator (magnetic), the direction of the Earths main magnetic
field is now horizontal. It still induces an anomalous magnetic field in the
metallic sphere, but the orientation of field lines describing the magnetic
field are now rotated 90 degrees. As in the previous case, these field lines
are indicated by the black lines, and the strength and direction of the anoma-
lous field at the surface of the earth are shown by the blue arrows. Above
the sphere, the anomalous magnetic field, Fa , now points in the opposite
direction as the Earths main magnetic field, Fe . Therefore, the total field
measured will be less than the Earths main field, and so upon removal of
the main field, the resulting anomalous field will be negative. On either
side of the sphere, the anomalous field points in the general direction of
the main field and thus reinforces it resulting in total field measurements
that are larger than the Earths main field. Upon removal of the main field
contribution, these areas will show positive magnetic anomalies.
As with the previous case, the resulting anomaly is again symmetrically
distributed about the center of the sphere. In this case, however, the promi-
Liner Env Geophysics Reader 112

Figure 57: Magnetic anomaly associated with a metallic sphere at the (mag-
netic) north pole.
Liner Env Geophysics Reader 113

nent central anomaly is negative and is surrounded by two smaller positive


anomalies.

2.6.4 Magnetic Anomaly: Magnetized Sphere in the Northern


Hemisphere

Finally, lets examine the shape of the anomalous magnetic field for a metal-
lic sphere buried somewhere in the northern hemisphere, say near Denver,
as seen in Figure 59.
As in the previous examples, the Earths main magnetic field induces
an anomalous field in surrounding the sphere. The anomalous field is now
oriented at some angle, in this case 45 degrees, from the horizontal. By
looking at the direction of the anomalous field, Fa , in comparison with
the Earths main field, Fe , you can see that there will be a small negative
anomaly far to the south of the sphere, a large postive anomaly just south
of the sphere, and a small, broad, negative anomaly north of the sphere.
Notice that the magnetic anomaly produced is no longer symmetric about
the sphere. Unless you are working in one of those special places, like at the
magnetic poles or equator, this will always be true.
From this simple set of examples, you now see that it is indeed more diffi-
cult to visually interpret magnetic anomalies than gravity anomalies. These
visual problems, however, present no problem for the computer modeling
alogrithms used to model magnetic anomalies. You simply need to incorpo-
rate the location of your survey into the modeling algorithm to generate an
appropriate magnetic model.
Liner Env Geophysics Reader 114

Figure 58: Magnetic anomaly associated with a metallic sphere at the (mag-
netic) equator.
Liner Env Geophysics Reader 115

Figure 59: Magnetic anomaly associated with a metallic sphere at mid-north


latitude.
116

Part III
Seismic Refraction Surveying
3 My section
3.1 My subsection
3.1.1 My subsubsection

Part IV
Electrical Conductivity Surveying
4 Electromagnetic techniques

There are many electrical and electromagnetic geophysical techniques. A


major distinction is whether the method involves a controlled source or
measures some ambient field generated by a natural source, Table 3. These
are also the most prominent in shallow geophysical applications.

Natural source Controlled source


magnetotelluric (MT) Induced polarization (IP)
Tellurics-MT (TMT) Time-domain E-M (TEM)
Audio frequency MT (AMT) Electrical resistivity (DC)
E-M array profiling (EMAP) Electrical conductivity
Self-potential (SP) Ground penetrating radar (GPR)

Table 3: Prominent electrical and E-M techniques.

Electromagnetic induction32 in geophysics rests on the classical theory of


low frequency electromagnetism. At frequencies below about 10 kHz, length
scales above 1 meter, and through most earth materials, energy transmis-
sion is characterized as the slow diffusion of electromagnetic fields into rel-
atively good conductors. Radiation, scattering and other high frequency
32
http://beerfrdg.tamu.edu/
Liner Env Geophysics Reader 117

phenomena often associated with electromagnetics are neglected. As such,


geophysical EM induction has much more in common with heat conduction
in solids, or fluid flow in porous media, than wave propagation or scattering
theory. Complications arise since the electrical conductivity of geological
formations is always to some degree heterogeneous, anisotropic, dispersive,
and/or length-scale dependent.
Electrical Conductivity. The Earth can be regarded as a complex, het-
erogeneous electrical conductor. Its electrical conductivity, like its other
major transport property, permeability, spans many orders of magnitude
and depends on the length scale over which it is measured. For example, a
stable craton is regarded as a resistive feature when viewed on the global
scale, yet on a much finer scale a craton contains pockets of conductive fluids
and may be covered by highly conductive soils or zones of highly resistive
permafrost. Less dramatically, a sedimentary sequence may contain resistive
sand layers commingled with more conductive clay and shale units.
In general, electrical conductivity contains information about the Earths
interior on different scales:

10 - 100 m (environmental, hydrological, geotechnical)


100 m - 10 km (mineral exploration, crustal studies)
10 - 200 km (tectonics, fault zones, volcanism)
200 - 1000 km (global sounding, mandtle dynamics)

5 Conductivity overview

Conductivity33 surveys introduce an alternating magnetic field into the soil


which induce currents to flow and in turn generate secondary magnetic fields.
The secondary magnetic fields are measured and used to estimate electrical
conductivity of the subsurface. Conductivity surveys respond to a combina-
tion of soil moisture, soluble ion concentration and physical soil type. Moist
33
Modified from http://www.archaeo-physics.com/method.htm
Liner Env Geophysics Reader 118

soils have higher conductivity than dry soils. Fine soils (clay) have higher
conductivity than coarse soils (sands or gravels), and high salinity soils have
high conductivity.
At a more fundamental level, the conductivity is governed by the num-
ber and mobility of free charge carriers available in the soil. The principal
sources of these free charge carriers are soluble ions. Thus, the simultane-
ous availability of soil moisture and soluble salts determines the free charge
carrier concentration in the soil. The mobility of these carriers is also an
important parameter in soil conductivity. The mobility of the soluble ions
is governed by soil moisture content, soil grain size, temperature, soil com-
paction as well as the surface chemistry of the soil grains. These processes
govern soil conductivity at the low frequency used in these surveys. At
higher frequencies soil conductivity becomes a more complex issue.

6 Terrain conductivity

Conductivity34 is given in units of mmho (millimhos) since terrain materials


fall in the range 1-1000 mmho, Table 4.

Material Conductivity (mmho)


air 0
distilled water 0.01
fresh water 0.5
sea water 5000-6000
dry sand 0.01
limestone 0.5 - 2
shales 1 - 100
clays 2 - 1000
granite 0.01 - 1
dry salt 0.01 - 1
ice 0.01

Table 4: Electrical conductivities of some materials.


34
Content by C. Liner, closely following McNeil (Geophysics ref?)
Liner Env Geophysics Reader 119

Major factors affecting electric conductivity of soil or rock include

porosity

pore fluid conductivity

particle shape of soil/rock particles

water saturation

clay conductivity (if clay is present)

All of the items except temperature dependence can be summarized in


the generic equation

d = A w m S n + c , (27)

where d is the terrain conductivity, A is a global constant, w is conduc-


tivity of the pore fluid, is the porosity, S is the water saturation, m is the
particle shape factor,n is an empirical constant, and c is conductivity of the
clay component of the soil. This equation is an extended form of Archies
law and is valid for unconsolidated or lithified material. Table 5 defines the
symbols and units used in this chapter.

Symbol Name Range Units Comment


A empirical constant none use 1
m particle shape factor 1-2 none spherical=1.2; platey=1.9
n fluid saturation exponent none use 2
S fluid saturation 0-1 none
T0 temperature of original survey C

T temperature of later survey C

porosity 0-1 none


c clay surface conductivity mmho
d terrain conductivity mmho
w pore fluid conductivity mmho

Table 5: Symbols and definitions for the conductivity method.


Liner Env Geophysics Reader 120

We will take the constants to have values of A = 1 and n = 2, leading to

d = w m S 2 + c . (28)

The particle shape factor, m, is usually given one of two values: 1.2 for
spheres (sandy soil) or 1.9 for platey fragments (shaley soil).

6.1 Temperature effect

Temperature dependence on soil conductivity is determined by the temper-


ature dependence of ionic mobility. This is a approximately 2% per C for
common ions. Since pore fluid conductivity is proportional to ionic mobility,
w will show the same 2% per C change. However, any clay conductivity
will not be affected by temperature. The idea is that a measurement is
taken at a certain time and temperature (T0 ), then the same measurement
is repeated at a later time and different temperature (T ). Even if all other
factors are constant, the conductivity will change because of temperature
change. These considerations lead to an extended equation for d which
accounts for temperature variation

d (T ) = w m S 2 e(T T0 )/50 + c . (29)

6.2 Example

1. At a clay-free site the terrain conductivity, d , measured in


November (12 C) and found to be 125 mmho. At the same site
in July (40 C), all other things being constant, we should find

d (T ) = 125 e28/50 = 218.8 mmho . (30)

2. Taking the July reading of 218.8 mmho as the first reading,


show this is consistent with a November reading of 125 mmho.

3. Imagine the same situation as example 1, except that the


original d of 125 mmho included 40 mmho of clay conductivity.
Liner Env Geophysics Reader 121

In this case the November measurement should yield

d (T ) = 85 e28/50 + 40 = 148.8 + 40 = 188.8 mmho . (31)

This example points to the possibility of multi-temperature con-


ductivity measurements being used to estimate clay conductivity
for a given site.

7 Linearity and resistivity

One nice feature of the terrain conductivity equation

d = w m S 2 + c . (32)

is that d is a linear function of pore fluid conductivity. Specifically, a


change in pore fluid conductivity, w w + w , has the following effect
on terrain conductivity:

d + d = (w + w ) m S 2 + c , (33)

for a terrain conductivity change

d = w m S 2 (34)

independent of the background conductivity d . This simplifies interpreta-


tion and decouples the effects of clay and pore fluid.
Resistivity values, on the other hand, are not as simple to work with.
The relationship between conductivity (mmho) and resistivity ( m) is

= 1000/ . (35)

Since the factor of 1000 will cancel, for every conductivity, there is a corre-
sponding resistivity

d = 1/d (36)
w = 1/w (37)
w = 1/w . (38)
Liner Env Geophysics Reader 122

Substituting these relationships into the terrain conductivity equation shows


that the terrain resistivity is given by
c w
d = . (39)
w + m S 2 c
The effect of clay and pore fluid resistivity changes are thus coupled and
more difficult to isolate.

8 Conductivity and water quality

A major use of conductivity is to map water quality35 . In this section, we


will derive an approximate relationship between terrain conductivity and
total dissolved solids in the pore fluid.

1 Assume average porosity for unconsolidated materials is = 0.4. See


Table 6.

2 From d = w m S 2 + c , assume full saturation (S = 1) and clay-free


soil (c = 0) and solve for the formation factor m . From the porosity
assumption and the range of m values (See Figure 60) it follows that
d 0.25 w . This relationship is relatively independent of the value
of m.

3 Conductivity of 1 gram N aCl in 1m3 of water (1 ppm by weight) is 0.22


mmho. Observation of natural waters shows a concentration of 1 ppm
total dissolved solids (TDS) increases conductivity by 0.16 mmho.

4 Using the 0.16 mmho value gives w 0.16 T DS. Combining this with
d 0.25 w gives the result
T DS
d T DS in ppm (40)
25

35
This section from McNeil ??
Liner Env Geophysics
1
Reader 123

0.8

0.6 m = 1.2
sd
f = m

sw
0.4
m = 1.9

0.2

0
0 0.2 0.4 0.6 0.8 1
Porosity

Figure 60: Relationship between formation factor and porosity for fully
saturated, clay-free soils. (Source: Redrawn from McNeil?)

This result means that the addition of 25 ppm of TDS to soil water will
increase saturated bulk conductivity by 1 mmho. Regulation place limits
on TDS in potable (drinkable) water at 500 ppm. This gives a pore fluid
conductivity of 80 mmho, and a terrain conductivity of 20 mmho (12.5 m).
This terrain conductivity is a value typical of clay-free soils, but can quickly
be overshadowed by clay conductivity.

9 EM31 conductivity meter


9.1 Theory

The36 simplest electromagnetic (EM) methods rely on the fact that an im-
posed, alternating primary magnetic field Hp will give rise to a secondary
36
modified from http:
eas4420.em.html
Liner Env Geophysics Reader 124

Material Porosity
soils .5-.6
clay .45-.55
silt .4-.5
med to coarse mixed sand .35-.45
uniform sant .3-.4
fine to med mixed sand .3-.35
gravel .3-.4
gravel and sand .2-.35
sandstone .1-.2
shale .01-.1
limestone .01-.1

Table 6: Porosities of some materials.

magnetic field Hs whose strength is directly proportional to the electrical


conductivity of a material:
4 Hs
a = 2
(41)
0 s Hp

where a is the apparent conductivity in mmho (equivalent to milliseimens


per meter), is frequency, 0 is the permeability of free space, Hp , Hs are
primary and secondary magnetic fields at the receiver, and s is the coil
spacing (m)
Electrical conductivity is the inverse of the better-known quantity re-
sistivity and refers to how easily current moves through a material. For a
constant value of the input (primary) magnetic field, the magnitude of the
induced secondary magnetic field will be much larger for a high conductivity
material than for a low conductivity material.
The electrical conductivity of various soil and rock types depends on
many factors. Some of these factors are the degree of saturation; the degree
of compaction; the salinity of pore waters; and soil composition. Like most
geophysical methods, electromagnetic techniques cannot uniquely determine
the composition of the subsurface. The instruments instead measure a phys-
Liner Env Geophysics Reader 125

ical parameter from which the composition and structure can be deduced.
The EM31 and EM34 measure electrical conductivity in units of mil-
liSiemens per meter (mS/m). The EM31 is a single operator instrument
that has a fixed spacing of 3.7 m between the coils. Two operators are re-
quired for the EM34, which has coils that can be spaced at 10 m, 20 m, or
40 m.

9.2 Basic Interpretation

The response function R for the EM-31 or EM-34 operated in vertical mode
is given by
1
R= (42)
4 L2 /s2 + 1
where L is the depth (m).
For a two-layer model (one layer over a half-space as shown in Figure 61)
the apparent conductivity is given by

a = 1 (1 R(L, s)) + 2 R(L, s) (43)

where R(L, s) is the response function evaluated at thickness L of layer 1


and intercoil spacing, s.

Figure 61: Single layer conductivity model. (Source: eas4420)


Liner Env Geophysics Reader 126

Apparent Conductivity v. Spacing


HEM-3134 vertical config,
s1=20 mmho, s2=100 mmho, L=5 mL

90
80
70

sa HmmhoL 60
50

40

30
20
0 5 10 15 20 25 30
Coil Spacing HmL

Figure 62: Conductivity and coil spacing.

If the conductivities (1 , 2 ) and the coil spacing, s, are known, the


thickness of layer 1 can then be found by solving for L
s
s (2 a )(a 21 + 2 )
L= (44)
2 (a 1 )2

9.3 Features

This information is modified from the manufacturers web site37

Maps geological variations, groundwater contaminants, or any subsur-


face feature associated with changes in ground conductivity.

Surveys are readily carried out in all regions, including those of high
surface resistivity such as sand, gravel and asphalt.

Effective depth of exploration is approximately six meters.


37
http://www.terraplus.com/em31.htm
Liner Env Geophysics Reader 127

Apparent Conductivity v. Thickness


HEM-31 vertical config,
s1=20 mmho , s2=100 mmho, s=2 mL

100

90

80

70
sa HmmhoL
60

50

40

0 1 2 3 4 5
Layer ThicknessHmL

Figure 63: Conductivity and layer thickness.

Two digital meters display both the quadrature phase and inphase
components that can be recorded simultaneously on the digital data
recorder.

Inphase component is extremely useful for for detecting shallow ore


bodies and, in waste site surveys, for searching for buried metal drums,
pipes, and other ferrous and non-ferrous metallic debris.

9.4 General

The simplest38 electromagnetic (EM) methods rely on the fact that an im-
posed, alternating magnetic field will give rise to a secondary magnetic field
whose strength is directly proportional to the electrical conductivity of a
material. Electrical conductivity is the inverse of the better-known quantity
resistivity and refers to how easily current moves through a material. For a
38
This section is modified from http://www.eas.gatech.edu/eas4420.em.html
Liner Env Geophysics Reader 128

constant value of the input (primary) magnetic field, the magnitude of the
induced secondary magnetic field will be much larger for a high conductivity
material than for a low conductivity material. The electrical conductivity
of various soil and rock types depends on many factors. Some of these fac-
tors are the degree of saturation; the degree of compaction; the salinity of
pore waters; and soil composition. Like most geophysical methods, elec-
tromagnetic techniques cannot uniquely determine the composition of the
subsurface. The instruments instead measure a physical parameter from
which the composition and structure can be deduced.
The EM-31 measures electrical conductivity in units of milliSiemens per
meter (mS/m) or millimhos (mmho). The EM-31 terrain conductivity meter
is a commonly used instrument for environmental applications.
The EM-31 is a 3.7-m long instrument consisting of a transmitter coil
and a receiver coil with a control unit in the middle. The instrument is
normally carried above the ground at hip-level, but it can also be operated
directly on the ground. In normal operation, no part of the instrument ever
needs to come in contact with the ground. Usually the instrument is carried
alongside the operator, so that it is oriented parallel to the operators steps.
The transmitter coil transmits the primary magnetic field. The receiver
coil measures the induced secondary magnetic field and displays the constant
of proportionality between the two magnetic fields (the electrical conductiv-
ity) on the control unit. The EM31 nominally senses to a depth of about
5-6 m. Figure 64 shows the instrument in operating position.
Figure 65 demonstrates how the EM-31 is assembled. The long white
pieces are usually stored in the cradles beneath the blue control unit. During
assembly, these long white pieces (which contain the coils) are detached from
the cradles and the connections between the control unit and the transmitter
and receiver are completed. We have cheat sheets available to guide you
in assembly and calibration of the instrument.
Once the control unit (shown upside down in Figure 65) is flipped over,
Liner Env Geophysics Reader 129

Figure 64: Operating position of EM-31 with features labelled. (Source:


Georgia Tech)

then the top pops off and instrument set-up and calibration begins. Fig-
ure 66 shows the dials and digital readouts on the face of the control unit.
The scale dial tells the instrument whether the expected terrain conduc-
tivity lies in the range of 0-200 mS/m (set dial on 100) or 200 mS/m (set
dial on 1000). If the scale dial is not set correctly, the DL720 data logger
(not shown here) will still log the measured value correctly.
The mode dial is primarily used during instrument calibration and setup.
During normal operation, the dial is set to oper.
The coarse-fine dials are used for adjustment during calibration and
setup.
Terrain conductivity values are displayed in the data readout window.
The orange reading button is located on the instruments transmitter.
When the operator reaches the spot to take a reading, the button is pressed
and the data are collected. When used with a data logger (not shown), it
Liner Env Geophysics Reader 130

Figure 65: Assembly of EM-31. (Source: Georgia Tech)

is also possible to run the EM-31 in continuous mode. In this case, it is not
necessary to press the reading button at each survey point.
Figure 67 is a fully labelled picture of the EM-31 in operation along a
survey line at the Georgia Tech Research Facility (GTRF) geophysics test
bed facility in Cobb County, Georgia. Note that the EM31 can be operated
either with or without a datalogger. When no logger is available, data are
recorded by hand directly off the digital readout screen.
The pictures shown here have all demonstrated operation of the EM-31
in vertical mode. In this mode, the digital readout faces upward during the
measurement. The designation of vertical refers to the fact that the coils
inside the long white tubes (transmitter and receiver) are oriented parallel
to the ground during the measurement, resulting in a vertical EM field. The
EM-31 can also be operated in horizontal mode, which involves rotating
the entire instrument 90 degrees about an axis aligned with the transmitter-
Liner Env Geophysics Reader 131

Figure 66: EM-31 controls. (Source: Georgia Tech)

receiver tubes. In this configuration, the digital readout faces toward or away
from the operator. In horizontal mode, the coils in the transmitter-receiver
tubes are aligned perpendicular to the ground during the measurement.
EM-31 data are also often collected in different directions at the same
survey location. For example, at a single spot, the operator might take one
reading in the vertical mode facing east along a survey line and then turn 90
degrees and take another reading in the vertical mode facing south, with the
instrument perpendicular to the survey line. Such surveys are particularly
useful for detecting any directional components to the subsurface feature.
For example, over a pipe oriented parallel to the survey line, one would
expect different terrain conductivity values when the measurement is taken
parallel to the survey line instead of perpendicular to the survey line.
Finally, the EM-31 records both the inphase and quadrature phase part
of the response signal. The quadrature component is the one most often used
Liner Env Geophysics Reader 132

Figure 67: EM-31 during data acquisition. (Source: Georgia Tech)

for general surveying and that the inphase component is useful for locating
metallic objects.
Figure 68 shows sample quadrature phase data from the EM-31. The
high values on the left side of the diagram were obtained close to the Atlantic
Ocean and reflect the presence of saltwater (very high conductivity; about
5000-6000 mS/m) in the near-surface. The values drop off quickly toward
the interior of the island. The feature at 1150 ft coincides with a recently
dug shallow irrigation well that pulls water out of the surficial aquifer. The
local increase in conductivity near this well indicates that a cone of saltwa-
ter may be developing below the pumping site. Using simple formula and
adopting assumptions about the conductivities of the freshwater-saturated
and underlying saltwater-saturated sand layers, it is possible to invert the
EM-31 data for the depth to the freshwater-saltwater interface beneath this
barrier island.
Liner Env Geophysics Reader 133

Figure 68: EM-31 quadrature phase data. (Source: Georgia Tech)

Since the EM-31 can only sense to 5-6 m below the surface, the low
background values obtained in the interior of this island likely indicate that
the freshwater-saltwater interface lies at depths greater than this nominal
sensing depth. The EM-34, an instrument related to the EM-31 but more
cumbersome to use, can sense to depths as great as 60 m when used in
vertical mode with maximum coil spacing (40 m).
Liner Env Geophysics Reader 134

9.5 Specifications

Measured Quantities:
1.Apparent conductivity of the ground in millisiemens per meter (mS/m, mmho)
2.Inphase ratio of the secondary to primary magnetic field in parts per thousand

Measuring Ranges:
Conductivity: 10,100,1000 mS/m (mmho)
Inphase: 12ppt

Dimensions:
Boom: 4.0m extended, 1.4m stored
Console: 24 x20 x18cm

Weight:
Instrument: 11kg
Shipping: 26kg

Noise Levels:
Conductivity: 0.1 mS/m (mmho)
Inphase: 0.03ppt

Primary Field Source: Self-contained dipole transmitter


Sensor: Self-contained receiver
Intercoil Spacing: 3.66 meters
Operating Frequency: 9.8kHz
Power Supply: 8 disposable alkaline "C" cells (approx. 20 continuous hr
Measurement Precision: 0.1% of full scale deflection
Measurement Accuracy: 5% at 20 mS/m (mmho)
Liner Env Geophysics Reader 135

9.6 Frequently asked questions

Some FAQ39

What are the EM31 units?40 The EM31 reads in millisiemens/m (mmho),
which can be converted to ohm-m by dividing the value into 1000.

What do EM31 measurements mean? The EM31 is basically mea-


suring fields induced in the Earth by its own field generator. In general,
it is providing a measure of the conductivity of the material beneath
the unit (down to about 6 m), but in the presence of a high anisotropic
medium (e.g., a pipe), things can get odd. A profile across a pipe, with
the unit parallel to the profile, will produce a large negative anomaly
at the pipe with 2 flanking highs. With the instrument perpendicular
to the profile, a large positive anomaly will be centered above the pipe.

9.7 Field procedure

Conductivity field procedures41 are similar to those for gravity surveying:

1. Arrive on-site with all necessary equipment in good working order.

2. Establish grid for subsequent profiling by using appropriate flagging


and a tape measure. Grid density should be commensurate with detail
desired from the profiling survey.

3. Calibrate EM instrument in accordance with manufacturers direc-


tions. Background electrical values should be measured outside known
or suspected contaminant areas.

4. Using an EM31-D, or its equivalent, begin profiling transects by con-


tinuously monitoring instrument readout. Stop at first flagged station,
39
http://cires.colorado.edu/people/jones.craig/GEOL4740GPR.html
41
http://www.otrain.com/sop/X0030 SOP30.html
Liner Env Geophysics Reader 136

note and record instrument reading and station number on standard-


ized field forms or in field book. Proceed with the transect repeating
the process at other flagged stations. Note any anomalous instrument
readings between flagged stations on field forms.

5. At each flagged station, obtain sounding measurement by rotating the


coil apparatus 90 degrees. Note and record these values on the field
form.

6. If utilities are present at the project site (e.g. power lines, buried
linear features), make certain to note the location of these features
and any anomalous EM readings resulting from the structures on the
field form.

9.8 Case history 1: Soil and rock characterization

Site map in Figure 69. Gridded conductivities and control points in Fig-
ure 70. Surface plot of conductivities showing 20 mmho threshold in Fig-
ure 71. The low conductivity areas (i.e., not red) would be good candidates
for ground penetrating radar which works best in terrain with conductivity
at or below 20 mmho.

9.9 Case history 2: Industrial waste

Purpose Historical42 aerial photographs and old records indicated the pres-
ence of buried waste materials scattered over an area of 30 acres.
In planning the excavation of these industrial wastes, knowledge of
their exact extent and distribution was required. EM conductivity
and metal detector surveys were performed with EM31 and EM61
instruments to delineate suspected burial trenches, pits, tanks, and
scattered cylinders. See Figure 72.
42
http://www.geosphereinc.com/cases-0.htm
Liner Env Geophysics Reader 137

Results Many burial features were accurately mapped by the investigation.


These trenches and pits contained large concentrations of metal such
as drums and industrial debris. Several features were found that were
not identified by early reconnaissance work.

Discussion Excavation and removal of buried wastes could not be planned


or executed without better definition of waste limits than provided
by historical photographs. A high density coverage of the site was
undertaken with 10 foot line spacings and data collected at 2.5 and
1 foot intervals for the EM31 and EM61. This density of coverage
permitted the recognition of underground utilities and culverts which
often disrupt the interpretation of studies with lower density coverage.
Note in the figure above that the water pipeline produced an EM31
anomaly 50 feet wide, but nearby waste trenches are still identifiable.
The EM61 results provide even higher resolution and minimize the
width of the pipeline. Both methods were used at this site because
the EM31 is capable of responding to deeper targets and filters out
small scattered metal. As seen above, both methods detected burial
trenches. In addition, the EM61 detected many smaller objects sur-
rounding the major burial zones, indicating that the trenches were
emplaced in larger landfilled areas.

9.10 Case history 3: Contaminate plume

Abstract Brownfields43 are abandoned, inactive, or underused industrial


sites, usually located in urban areas, that are affected by either real
or perceived environmental contamination. Economic revitalization of
these areas requires that these sites be environmentally assessed and
made safe for future development. The Poinciana Industrial Center in
43
This material is a partial extraction from http://www.technos-
inc.com/BRNFIELD.HTM
Liner Env Geophysics Reader 138

Miami, Florida is a 21-acre brownfield site that was studied in Phase


I and II environmental site assessments. Extensive review of histor-
ical documents, analysis of aerial photographs from the 1940s to the
present, and surface observations did not indicate the presence of any
subsurface structures (although septic tanks were suspected). Initial
soil and groundwater assessment conducted as part of the Phase II
investigation did not reveal any significant amounts of contamination.
A geophysical survey was then carried out to provide adequate spatial
sampling of the 21 acres in order to detect suspected septic tanks. The
geophysical survey revealed a number of underground storage tanks,
septic tanks, a contaminant plume and other buried debris that were
not evident during the Phase I and initial Phase II investigations.
Most of these features (55 total including eight 1000-3500 gallon tanks)
would have gone undetected if historical searches and conventional
random sampling procedures were the only methods used to assess
the property. The results of the geophysical survey guided subsequent
excavations and soil sampling which provided a thorough assessment
of the property for remediation and redevelopment. The Phase II geo-
physical survey provided the regulators and stakeholders appropriate,
adequate and accurate information in order to assess environmental
compliance and minimize future financial and environmental risks.

EM-31 conductivity survey An electromagnetic (EM) survey using a


Geonics Ltd. EM31 instrument was chosen as the primary means
to locate the suspected septic tanks and other buried metal. This in-
strument is operated on the surface, and measures both the presence
of buried metal and the electrical conductivity of the subsurface. The
EM31 responds to the presence of buried metal (both ferrous and non-
ferrous) such as pipes, drums, tanks, and metallic debris. Electrical
conductivity is a function of the soil and rock type, porosity, perme-
ability and composition of pore fluids. An increase in total dissolved
Liner Env Geophysics Reader 139

solids (TDS) levels in pore fluids will result in higher electrical conduc-
tivity values, which can be used to locate changes in natural geologic
conditions and inorganic contaminants. EM31 measurements have a
response to materials from the surface down to approximately 20 ft,
with the bulk of the response centered at 4.8 ft, providing a sampling
volume of more than 1,000 cubic ft.

A survey grid was established and the EM31 data were collected within
the survey grid along parallel survey lines spaced 10 feet apart. The
10-foot line spacing was chosen to provide the most complete spa-
tial coverage of the site given the time and budget constraints of the
project. A total of 98,300 linear feet (18.6 miles) of EM31 data were
acquired. The EM31 data were contoured and evaluated in the field,
and anomalies due to subsurface metal or higher electrical conductivity
were identified.

Contaminant plume In one portion of the site (Tract BE), a large high
conductivity area was observed in the EM31 data that is not entirely
associated with subsurface metal (Figure 73). Based on historical data
from Phase I and II research, the high conductivity area lies where a
pesticide plant was once located. The high conductivity area likely rep-
resents a plume which may be the result of an inorganic component
of the pesticides, possible biodegradation products from the pesticide,
or other inorganic components associated with the manufacturing of
pesticides. Chemical assessment and remediation within this high con-
ductivity area is currently being planned.

Conclusions The geophysical investigation of the Poinciana Industrial Cen-


ter revealed a number of subsurface features that were not apparent
from the Phases I and II historical searches, aerial photographs, or
chemical sampling alone. A total of eight verified USTs were found
on this site, even though no existing documentation could be found
Liner Env Geophysics Reader 140

for them. Also, a large plume possibly associated with pesticide con-
tamination was discovered. Excavation and sampling efforts were ef-
fectively guided by the geophysical survey so that an accurate and
efficient environmental assessment could be carried out.
Liner Env Geophysics Reader 141

25
W26 W27 W28
24

23

22

21
N
20
W25 W24 W23 W22
19

18

17

16

15
W18 W19 W20 W21
14

13

12

11

10
W17 W16 W15 W14
9

7 W7 W8 W9

5
W6 W5 W4 W3
4

1
-4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

OU / CONOCO / TU Site Map


Gypsy OCAST Project Detail Plan of Gypsy Study Site
Approximately 15 mi west of
OU: Dr. R. Young Tulsa, OK
Conoco: Dr. S. Danbom scale
Date: 16 August 1995
TU: Dr. C. Liner 80 0 80
Feet By: rlw, cll

Figure 69: Conductivity case history 1. Site map for Gypsy project showing
grid and well locations.
Liner Env Geophysics Reader 142

Figure 70: Conductivity case history 2. Gridded conductivity values and


control points.
Liner Env Geophysics Reader 143

Figure 71: Conductivity case history 2. Surface plot of gridded conductivi-


ties showing 20 mmho threshold (red = above 20 mmho).
Liner Env Geophysics Reader 144

Figure 72: Conductivity case history 2. (Source: Geosphere, Inc.)

Figure 73: EM-31 data over a contaminate plume. (Source: Technos, Inc.)
145

Part V
Ground Penetrating Radar
Surveying
10 My section
10.1 My subsection
10.1.1 My subsubsection

You might also like