Sie sind auf Seite 1von 136

THREE-DIMENSIONAL INVERSION OF MAGNETOTELLURIC DATA FROM

THE COSO GEOTHERMAL FIELD, BASED ON A FINITE DIFFERENCE,


GAUSS-NEWTON METHOD PARALLELIZED ON A
MULTICORE WORKSTATION

by
Virginie Maris


A dissertation submitted to the faculty of
The University of Utah
in partial fulfillment of the requirements for the degree of


Doctor of Philosophy
in
Geophysics


Department of Geology and Geophysics
The University of Utah
May 2011










Copyright Virginie Maris 2011
All Rights Reserved


3


The Uni ve r s i t y of Ut ah Gr aduat e Sc hool



STATEMENT OF DISSERTATION APPROVAL



The dissertation of Virginie Maris
has been approved by the following supervisory committee members:

Michael S. Zhdanov , Chair 11/09/2010

DateApproved
John M. Bartley , Member 11/11/2010

DateApproved
Susan L. Halgedahl , Member 11/17/2010

DateApproved
George R. Jiracek , Member 11/17/2010

DateApproved
Philip E. Wannamaker , Member 11/9/2010

DateApproved

and by D. Kip Solomon , Chair of
the Department of Geology and Geophysics

and by Charles A. Wight, Dean of The Graduate School.




ABSTRACT


An existing 3-D magnetotelluric (MT) inversion program written for a single
processor personal computer (PC) has been modified and parallelized using OpenMP, in
order to run the program efficiently on a multicore workstation. The program uses the
Gauss-Newton inversion algorithm based on a staggered-grid finite-difference forward
problem, requiring explicit calculation of the Frchet derivatives. The most time-
consuming tasks are calculating the derivatives and determining the model parameters at
each iteration. Forward modeling and derivative calculations are parallelized by assigning
the calculations for each frequency to separate threads, which execute concurrently.
Model parameters are obtained by factoring the Hessian using the LDL
T
method,
implemented using a block-cyclic algorithm and compact storage.
MT data from 102 tensor stations over the East Flank of the Coso Geothermal
Field, California are inverted. Less than three days are required to invert the dataset for ~
55,000 inversion parameters on a 2.66 GHz 8-CPU PC with 16 GB of RAM. Inversion
results, recovered from a halfspace rather than initial 2-D inversions, qualitatively
resemble models from massively parallel 3-D inversion by other researchers and overall,
exhibit an improved fit. A steeply west-dipping conductor under the western East Flank is
tentatively correlated with a zone of high-temperature ionic fluids based on known well
production and lost circulation intervals. Beneath the Main Field, vertical and north-
trending shallow conductors are correlated with geothermal producing intervals as well.


CONTENTS

ABSTRACT ....................................................................................................................... iii
ACKNOWLEDGMENTS ................................................................................................. vi
CHAPTERS
1 INTRODUCTION ........................................................................................................... 1
2 MT PRINCIPLES ............................................................................................................ 5
2.1 Introduction ......................................................................................................... 5
2.2 Maxwell's equations for MT theory .................................................................... 6
2.3 1-D model ........................................................................................................... 8
2.4 2-D model ......................................................................................................... 10
2.5 3-D model ......................................................................................................... 12
3 STATIC SHIFT ............................................................................................................. 15
3.1 Theory ............................................................................................................... 16
3.2 Correction techniques ....................................................................................... 20
3.3 Distribution of shifts ......................................................................................... 22
3.4 Impedance ratio frequency-dependence ........................................................... 23
3.5 Shift frequency-dependence.............................................................................. 25
3.6 Summary ........................................................................................................... 30
4 FORWARD MODELING .......................................................................................... 32
4.1 Finite difference solution .................................................................................. 33
4.2 Biconjugate gradient method ............................................................................ 35
4.3 Divergence correction ....................................................................................... 35
5 INVERSION SCHEME .............................................................................................. 37
5.1 Theory ............................................................................................................... 38
5.2 Frchet derivatives ............................................................................................ 43
5.3 Static shift ......................................................................................................... 45
5.4 Modifications from original .............................................................................. 48
5.5 Synthetic test ..................................................................................................... 49
v


6 PARALLELIZATION OF THE INVERSION CODE .............................................. 58
6.1 Introduction ....................................................................................................... 58
6.2 Frchet derivatives ............................................................................................ 60
6.3 Parameter update ............................................................................................... 63
6.4 Summary ........................................................................................................... 67
7 INVERSION OF MT DATA FROM THE COSO GEOTHERMAL FIELD ............ 68
7.1 Conceptual model for high-temperature geothermal systems .......................... 71
7.2 Coso Geothermal Field ..................................................................................... 73
7.2.1 Geologic setting ........................................................................................ 74
7.2.2 CGF description ........................................................................................ 76
7.3 Inversion ........................................................................................................... 78
7.3.1 Data ........................................................................................................... 78
7.3.2 Inversion set #1: 100 Hz - 0.63 Hz .......................................................... 83
7.3.3 Inversion set #2: 100 Hz - 1.6Hz ............................................................. 92
7.3.4 Discussion ............................................................................................... 103
7.4 Conclusions ..................................................................................................... 117
8 SUMMARY .............................................................................................................. 118
REFERENCES ............................................................................................................ 120

v
115


ACKNOWLEDGMENTS

I would like to thank Dr. P. E. Wannamaker for support and guidance of this
work, and Dr. Y. Sasaki for providing the original program on which it is based.
I am honored and deeply grateful to have been the recipient of the Society of
Exploration Geophysicists Stan and Shirley Ward, and the Charlie and J ean Smith,
graduate student scholarships.
Dr. J . Moore of the Energy and Geoscience Institute kindly spent much time
providing insight into the Coso geothermal system.
I am grateful to Dr. G. Newman and Dr. E. Gasperikova for providing Coso well
production and injection intervals, and to Dr. B. R. J ulian for earthquake epicenter data.
I would like to thank Dr. M. S. Zhdanov, chair of my supervisory committee, and
committee members Dr. J . M. Bartley, Dr. S. L. Halgedahl, Dr. G. R. J iracek, and Dr. P.
E. Wannamaker for review of this work.
Work on this inversion method was supported by U.S. Dept. of Energy contract
DE-PS36-04GO94001 to the University of Utah, Energy and Geoscience Institute (P.
Wannamaker, P. I.).
MT Data collection at the Coso Geothermal Field was supported under U.S. Dept.
of Energy contract DE-PS07-00ID13913 and Dept. of the Navy contract N68936-03-P-
0303 to the Energy and Geoscience Institute.


CHAPTER 1

1INTRODUCTION

As three-dimensional (3-D) magnetotelluric (MT) surveys investigating the
subsurface electrical structure are becoming more common, access to fast, accurate 3-D
MT inversion programs has become more important (e.g. Uchida and Sasaki, 2006). The
serial Fortran77 program developed by Sasaki (2001, 2004) for Gauss-Newton inversion
of MT data based on the finite-difference staggered grid method for forward modeling
has been restructured and parallelized under Linux using OpenMP 2.0 to allow it to run
efficiently on a multicore workstation. The modified program is used to invert MT data
from 102 soundings over the East Flank of the Coso Geothermal Field (CGF), imaging
conductive zones tentatively correlated to producing reservoirs.
MT is a geophysical technique whereby naturally occurring electromagnetic (EM)
waves are used as source fields for imaging Earths electrical resistivity structure at
depths ranging from tens of meters to hundreds of kilometers (Vozoff, 1991; Simpson
and Bahr, 2005). It is used in resource exploration and in earthquake and volcano studies.
Interpretation of MT data can be further complicated by static shifts. These shifts are the
frequency-independent, site- and source- specific responses of small-scale, near-surface
inhomogeneities, which can mask the response of deeper structures of interest. Chapters 2
and 3 provide a brief review of MT theory and static shifts.
2


Estimating the 3-D subsurface electrical conductivity structure from observed MT
data is done by nonlinear regularized parametric inversion. Several 3-D MT inversion
computer codes have been developed (Zhdanov and Golubev, 2003; Sasaki, 2004;
Mackie and Watts, 2004; Siripunvaraporn et al., 2004, 2005; Wan et al., 2006; Green et
al., 2008; Zhdanov and Gribenko, 2008). A difficult and computationally intensive task,
3-D MT inversion remains an active research area (Avdeev, 2005). One approach has
been to develop distributed computing clusters (e.g. Newman and Alumbaugh, 2000;
Hargrove et al., 2001; Wan et al., 2006; Green et al., 2008; Zhdanov and Gribenko,
2008), although these can require a substantial investment and facility footprint. An
attractive alternative is to exploit multicore designs; this presents the prospect of parallel
computing within an affordable, single-box, format.
Sasaki (2001, 2004) developed a Gauss-Newton MT inversion algorithm for serial
PCs using the staggered-grid finite-difference forward modeling method to solve
Maxwell's equations. More information on forward modeling is presented in Chapter 4.
Estimating the subsurface resistivity structure requires factoring the approximate
regularized Hessian, formed from a matrix of Frchet derivatives. The inversion
algorithm of Sasaki (2004) is formulated to solve the inverse problem simultaneously for
static shift and 3-D subsurface conductivity distribution parameters, and is discussed in
Chapter 5. Modifications made to the program other than parallelization are also
described in this Chapter.
The program has been modified to allow it to run efficiently on a multicore PC.
Parallelization is accomplished using OpenMP, an easy to use application program
interface developed for shared-memory platforms such as multicore PCs. The serial
3


Fortran77 program has been restructured and parallelized under Linux using OpenMP 2.0
directives embedded in the source code and tested on an Intel Xeon 5355, 2.66 GHz, 8-
core PC, with 16 GiB of RAM. Two key areas of the program that were parallelized are
the frequency loop containing the forward modeling and Frchet derivative calculations,
and factoring the approximate regularized Hessian, both time-consuming tasks.
Parallelization is described in Chapter 6.
The MT method has been successfully used to image subsurface electrical
resistivity in complex geothermal systems, detecting electrical resistivity variations
related to fluid flow, including due to high fluid concentrations in fractures, and to
conductive alteration minerals. The modified inversion program has been applied to
inverting MT tensor data from 102 sites collected at the East Flank of the Coso
Geothermal Field, located in southeast California (Figure 1.1). The Coso Geothermal area
is a high-temperature power-producing field in southeastern California (Monastero et al.,
2005). Previously published interpretations of the Coso data have included 3-D inversion
using a massively parallel computer, of a finely discretized model seeded with a starting
model incorporating 2-D inversion results (Newman et al., 2005a, 2005b; Newman et al.,
2008). An important structure appearing in these interpretations is a high-angle conductor
most prominent in the southwest East Flank sector correlated with fluid-filled fractures.
As is shown here, similar results can be obtained from inversion on a workstation,
starting from a halfspace. Additionally, a vertical conductor and a north-trending shallow
conductor are imaged beneath the Main Field, correlated with producing well intervals.
Further information can be found in Chapter 7.
4




Chapters 2 and 3 provide a brief review of MT theory and static shifts. More
information on forward modeling is presented in Chapter 4. The inversion algorithm
along with modifications other than parallelization are discussed in Chapter 5. Chapter 6
covers the details of parallelization. Inversion of MT data from the Coso Geothermal
Field and its geothermal implications are discussed in Chapter 7. A brief summary is
provided in Chapter 8.
Figure 1.1 Map of California showing the approximate location of the Coso Geothermal
Field. Modified from USGS Physiographic provinces map of California,
http://education.usgs.gov/california/maps/provinces_B&W1.htm, accessed J une 2010.



CHAPTER 2

2MT PRINCIPLES

2.1 Introduction
The MT method was launched by Tikhonov (1950) and Cagniard (1953). It
consists of measuring the naturally occurring, time-varying, orthogonal electric (E) and
magnetic fields (H) penetrating the earth, at a stationary point on the earth's surface. The
fields are measured at frequencies typically in the range of 0.001 Hz to 1000 Hz (Vozoff,
1991). For frequencies below 1 Hz, the fields are due to current systems in the earth's
magnetosphere; worldwide thunderstorm activity generates frequencies higher than 1 Hz
(Vozoff, 1991). At the earth's surface, these fields are considered to be downward-
propagating plane waves (Madden and Nelson, 1964). The amplitude and phase scaling
relationships between the fields depend on the subsurface conductivity structure and field
frequency, and are independent of the source strength. These relationships are expressed
through the use of the MT impedance (Z), a rank 2 tensor determined from:

2.1


The units of E are V.m
-1
and of H are A.m
-1
. The apparent resistivity and phase are

y
x
y
x
H
H
E
E
Z
6


defined from the components of the surface MT impedance measurements according to:

0
2

Z
a
=

=

) Re(
) Im(
tan
1
Z
Z


where is the angular frequency (rad.s
-1
) and
0
is the magnetic permeability of free
space, 4 x 10
-7
H.m
-1
.
The method has been thoroughly reviewed in numerous publications including,
but not limited to, Vozoff (1991), J iracek et al. (1995), Madden and Mackie (1989), and
more recently, Simpson and Bahr (2005) and Zhdanov (2009). The basis of the method,
Maxwell's equations, can be found in electromagnetics textbooks such as, but not limited
to, Stratton (1941) and Harrington (1961). Section 2.2 is intended to provide a brief
review of Maxwell's equations formulated for the MT method. The concepts of MT
impedance, apparent resistivity and phase are developed for 1-D homogeneous and
horizontally layered models in section 2.3, for 2-D models in section 2.4, and for 3-D
models in section 2.5.

2.2 Maxwell's equations for MT theory
MT theory is based on Maxwell's equations and the assumptions of a quasi-static,
monochromatic field in a source-free region. The governing equations, formulated in the
frequency domain for an
t i
e
+
time-dependence, where is the angular frequency
(rad.s
-1
), 1 = i , and t is the time (s), are:
2.2
2.3
7


E H =
H E i =

where E is the electric field intensity (V.m
-1
) and H is the magnetic field intensity (A.m
-
1
). The electric conductivity, (S.m
-1
) and the magnetic permeability, (H.m
-1
) are
material properties. At MT frequencies, contributions from dielectric displacement are
ignored. It is commonly assumed in MT theory that , and , are frequency-independent;
furthermore, that can be adequately represented as constant and equal to the free-space
value of =
0
=1.2566x10
-6
H.m
-1
(Keller, 1991). Expressing the fields as:

H E =

1

E H

=
i
1


equations 2.4 and 2.5 can be transformed into

( ) E E i =
( ) H H

i =
1

( ) 0 = E


2.4
2.5
2.6
2.7
2.8
2.9
2.10
2.11
. 0 = H
8


In the presence of a sharp conductivity boundary, the components of H and E tangential
to the boundary must be continuous across it, while E normal is discontinuous,
preserving continuity of current j where E j = (e.g. Ward and Hohmann, 1988).
In the following sections, an e(+it) time dependence is assumed together with a
right-hand cartesian coordinate system with z positive downward.

2.3 1-D model
In a homogeneous medium, is constant; 0 = E . Equations 2.8 and 2.9 reduce
to the Helmholtz equations

( )
( ) 0
0
2
2
=
=
H
E

i
i
.

For a plane wave vertically incident on a homogeneous or horizontally layered halfspace,
field vectors always lie in horizontal planes, and are instantaneously equal over entire
plane, such that derivatives with respect to x and y directions are zero (e.g. Zhdanov,
2002). The Helmholtz equations reduce to:

0
0
2
2
2
2
2
2
=

H
E
k
z
k
z


2.12
2.13
9


where the wavenumber i k = , ) 0 , , (
y x
E E = E , and ) 0 , , (
y x
H H = H . These are
homogeneous second-order ordinary differential equations with the general solution, for
equation 2.13 (Vozoff, 1991),

ikz ikz
e e E E E
+
+ =

where z is the distance from the surface, and
-
E and
+
E are the upgoing and downgoing
field amplitudes. In a uniform half-space, there is only a downgoing field,

ikz
e
+
= E E

with corresponding magnetic fields from equation 2.7,

kz
y
y
x
e E
i
z
E
i
H
+
=

1

and
kz
x
x
y
e E
i
z
E
i
H
+
=

1
.


Off-diagonal elements of the impedance are equal, and can be expressed as:

i
Z Z
yx xy
= =

2.14
2.15
2.16
2.17
2.18
10


independent of the source strength and coordinate definition. On-diagonal elements are
zero. The surface impedance is intrinsic and the apparent resistivity equals the half-space
resistivity. For an e
(+it)
time dependence, the phase will be in the first quadrant, at 45 for
the XY mode, and in the third quadrant, at 135 for the YX mode. The YX mode phase
can be represented in the first quadrant by the addition of 180. The depth of exploration
is taken to be skin depth (), given by

( )
1
2

= .

For a conductivity structure consisting of horizontal layers, the fields inside each
layer are governed by the Helmholtz equations, but must satisfy the boundary condition
that E and H fields tangential to the interface between layers must be continuous across.
Fields at the surface can be determined by assuming a starting value for the amplitude of
+
E in the basal half-space and, working upwards, calculating the fields recursively at
overlying interfaces (e.g. Ward and Hohmann, 1988). The surface impedance is not
intrinsic; the apparent resistivity will be a weighted average of the sampled conductivity
structure and the phase will digress from 45 for the XY mode, and 135 for the YX
mode.

2.4 2-D model
In 2-D models, conductivity varies with depth and horizontally along one axis
only. Structures strike perpendicular to this vertical plane and are assumed to extend to
infinity; vertical planes perpendicular to the strike direction are identical. Consider, for
2.19
11


example, a structure with conductivity variations only in the y-z plane. For measurement
axes coincident with the structural axes, field derivatives with respect to x, the strike
direction, will be zero and scalar Maxwell's equations yield

.
1 1 1 1
1 1
2
2
2
2

=
z
H
z y
H
y i z
E
y
E
i
H
z
E
y
E
i z
H
y
H
E
x x
y
z
x
x x
y
z
x




The MT response separates naturally into two modes which propagate independently:
transverse electric (TE), where ( ) 0 , 0 ,
x
TE
E = E , is parallel to strike, and ( )
z y
TE
H H , , 0 = H ;
and transverse magnetic (TM) where ( )
z y
TM
E E , , 0 = E , and ( ) 0 , 0 ,
x
TM
H = H , is parallel to
strike. (e.g. Vozoff, 1991; deLugao and Wannamaker, 1996; Zhdanov, 2002). At lateral
conductivity variations, E
TE
is tangential to and continuous across the interface,

TE TE
i E E =
2


while E
TM
is normal to the interface, and discontinuous from charges accumulated at the
boundary maintaining current continuity (e.g. Wannamaker et al., 1984; J iracek, 1990).
Hence, 0
TM
E and the galvanic term, ( )
TM
E is required when formulating
equation 2.8 for the TM mode,

( )
TM TM TM
i E E E =
2
.

2.20
2.21
2.22
12


Note that H
TM
, parallel to strike, remains continuous and can be determined from

( )
TM TM
i H H

=
1
,

from which auxiliary E
TM
components can be calculated (deLugao and Wannamaker,
1996). The TE mode is considered more prone to distortion from finite strike, 3-D effects
than the TM mode because boundary charge effects are absent from TE physics
(Wannamaker, 1999).
For a 2-D structure striking in the x-direction, the impedance consists of off-
diagonal elements only,

=
=
=
0
0
x
y
yx
y
x
xy
H
E
Z
H
E
Z
Z .

The YX component of the impedance will correspond to the transverse magnetic (TM)
mode and the XY to the transverse electric (TE) mode in this coordinate convention. A
rotation matrix can be applied to data collected at an arbitrary angle, to align it with the
strike direction (e.g. Vozoff, 1972; Simpson and Bahr, 2005).

2.5 3-D model
3-D conductivity structures can be represented as consisting of a domain of
anomalous conductivities, superimposed on 1-D, homogeneous or horizontally layered,
2.23
2.24
13


background. The total fields (E
t
,H
t
) can be represented as the sum of background fields
(E
p
,H
p
) generated by the idealized horizontally-layered background, and scattered fields
(E
s
,H
s
), caused by the anomalous domain (E
s
,H
s
):

s p T
s p T
H H H
E E E
+ =
+ =


Expressing equation 2.8 in terms of E
p
and E
s
, and making use of equation 2.12,
describing the behavior of E
p
in a 1-D model, the relationship between the background
and scattered fields can be expressed as:

p p s s
i i E E E ) ( = +

where
p
is the conductivity of the horizontally layered earth at the depth at which the
fields are evaluated. Analytic solutions have been published for specific geometries, but
for an arbitrary conductivity distributions, there are no analytical solutions, and scattered
fields must be numerically approximated.
To calculate the impedance in computer simulations, fields are measured for at
least two source directions, s
1
, and s
2
, preferably orthogonal to one another, where

HH
EH
Z
ij
ij
=
and

2.25
2.26
2.27
14


.
2 1 2 1
2 2 1 2
1 2 2 1
1 2 2 1
2 1 1 2
s
x
s
y
s
y
s
x
s
x
s
y
s
x
s
y yy
s
y
s
x
s
y
s
x xx
s
y
s
y
s
y
s
y yx
s
x
s
x
s
x
s
x xy
H H H H HH
H E H E EH
H E H E EH
H E H E EH
H E H E EH
=
=
=
=
=


Estimating the impedance from measured EM field time series invokes a statistical
generalization of the preceding equation where the source field polarizations are
presumed to vary over the recording interval (e.g. Gamble et al., 1979). In a regionally 3-
D environment, the on-diagonal elements of the impedance are non-zero no matter what
coordinate orientation is chosen except in special cases of symmetry.
2.28





CHAPTER 3

3STATIC SHIFT

A static shift is, by definition, the frequency-independent distortion of an apparent
resistivity sounding (see review in Wannamaker, 1999). It is caused by a near-surface
inhomogeneity of dimension significantly smaller than the shortest EM field wavelength
sampled in the sounding. The inhomogeneity is assumed to have a frequency-independent
galvanic-only response; the frequency-dependent inductive component of the response is
considered negligible (e.g. Groom and Bailey, 1989). In the case of 1-D and 2-D TE
mode regional structures, meeting this condition is sufficient in representing, and
removing, the data distortion using frequency-independent parameters. In the case of 3-D
regional structures, the data distortion may be frequency dependent, despite the near-
surface inhomogeneity generating only a galvanic response.
A theoretical discussion of shifts is provided in section 3.1, followed by a brief
review of different correction techniques in section 3.2. Shift distribution is discussed in
section 3.3. Shifts are simulated for a hypothetical near-surface inhomogeneity contained
within a regional 3-D model representing an idealized geothermal system and examined
in sections 3.4 and 3.5, followed by a discussion in section 3.6.

16


3.1 Theory
Consider a regional subsurface conductivity structure, which may be 3-D, with
corresponding regional electric (E
R
) and magnetic (H
R
) fields related by the impedance
(Z
R
), where the superscript R denotes that these quantities are defined for the regional
subsurface conductivity structure. A "distorting" small-scale, near-surface inhomogeneity
is introduced in the regional structure. The measured electric field (E
d
) can be expressed
as the sum of the regional field and of a scattered field, introduced by the presence of the
inhomogeneity:

.
S R d
E E E + =

Following Wannamaker et al. (1984), assuming that the inhomogeneity is small enough
that the regional electric field is constant over the inhomogeneitys lateral extent, the
scattered field corresponding to the inhomogeneity can be written in terms of the incident
regional field as:

[ ]
R R d
P E E E + =

assuming that the fields are linearly related. No assumptions are made regarding the
character of the entries in P at this time, allowing it to remain complex. Assuming that the
total magnetic field is unaffected by the inhomogeneity and remains essentially H
R
, the
"distorted" impedance, Z
d
, relating the fields in the presence of the inhomogeneity, can be
expressed in terms of the regional fields as:
3.1
3.2
17


[ ]
R d R R
P H Z E E = +

where the entries of P depend on the inhomogeneity and are complex. Agarwal and
Weaver (2000) demonstrate the validity of this assumption for the frequency range
commonly used in MT soundings. The distorted and regional impedances are related by:

[ ]

+
+
=
+ =
R
yy
R
yx
R
xy
R
xx
R d
Z Z
Z Z
d I c
b a I
P Z Z 1


where the entries of [ ] P , inhomogeneity parameters a, b, c and d, may be complex and
frequency-dependent. Individual modes can be expressed as:

.
R
yy
R
yy
R
xy d
yy
R
yx
R
yx
R
xx d
yx
R
xy
R
xy
R
yy d
xy
R
xx
R
xx
R
yx d
xx
Z d I
Z
Z
c Z
Z d I
Z
Z
c Z
Z
Z
Z
b a I Z
Z
Z
Z
b a I Z

+ + =

+ + =

+ + =

+ + =


If the inhomogeneity is of dimension and conductivity such that it is smaller than the skin
depth at the highest frequency of interest, then it can be assumed that the frequency-
dependent part of the scattered field can be neglected, leaving only the galvanic
3.3
3.4
3.5
18


component. The inhomogeneity parameters are then real only and frequency-
independent.
The XY mode apparent resistivity and phase in the presence of the inhomogeneity
can be expressed as:

R
xy a
R
xy
R
yy
R
xy R
xy
R
yy
d
xy a
Z
Z
b a
Z
Z
Z
b a
,
2
0
2
2
,
1
1

+ + =
+ +
=

( )
( )
R
xy
R
xy
R
xy d
xy
Z
Z
=
=
tan
Re
Im
tan


with similar expressions obtained for the remaining modes. The phase is unaffected by
the inclusion of the inhomogeneity in the regional structure. The apparent resistivity is
scaled by a multiplicative factor termed shift. Note than when displayed on a log-scale,
the shift appears additive.
For a regionally 1-D (horizontally planar) or 2-D structure with measurement axes
rotated parallel to strike, the impedances simplify to:






3.6
3.7
19


( )
( )
R
xy
d
yy
R
yx
d
yx
R
xy
d
xy
R
yx
d
xx
R
yy
R
xx
cZ Z
Z d I Z
Z a I Z
bZ Z
Z Z
=
+ =
+ =
=
= = 0


with
R
yx
R
xy
Z Z =
for the 1-D case. For these situations, the apparent resistivity shift will be
frequency-independent provided the inhomogeneity parameters are frequency-
independent.
As developed above for the 3-D case, the apparent resistivity shift may vary with
frequency because of the inclusion of complex, frequency-dependence of the regional
impedance ratios. Depending on the location of the receiver with respect to regional
structure, the ratio of on- to off-diagonal impedances may be frequency-dependent; thus,
even if the inhomogeneity generates only galvanic fields, the frequency-dependence may
be re-introduced into the static shift.
Pellerin and Hohmann (1990) and Spitzer (2001) demonstrate that the magnitude
and direction of the shift depends on the location of the electric field measurement
electrodes with respect to the location of the inhomogeneity. In practice, the electric field
in a particular direction is obtained by measuring the voltage between a pair of electrodes
aligned in such a direction, and dividing by the distance between them. Accurate
modeling of shifts in the vicinity of the inhomogeneity requires including the location of
the electric field measurement electrodes with respect to the location of the
inhomogeneity. In most cases, the electric field measurements are simulated as entirely
3.8
20


within or entirely outside the inhomogeneity. No consideration is given to the position of
the electrodes.

3.2 Correction techniques
Numerous techniques have been developed to address quantifying and possibly
removing static shift from MT data prior to inversion and interpretation. Identifying the
presence of and quantifying how regionally 1-D and 2-D data are affected by 2-D and 3-
D effects in small-scale overlying structure lead to the developing of distortion analysis
and phase tensor techniques (Bahr, 1988; Groom and Bailey, 1989; Caldwell et al.,
2004). Wannamaker (1999) groups these techniques into invariants and averages, spatial
averaging, and impedance tensor decomposition. Reviews of these techniques can also be
found in J iracek (1990). The majority of these techniques are primarily applicable to 1-D
and 2-D interpretations. Ogawa (2002) identifies three categories of techniques most
applicable to dealing with static shifts in 3-D interpretations: (1) spatial filtering, as in
electromagnetic array profiling (EMAP); (2) the use of independent information that is
free from galvanic distortion, such as obtained by time-domain (TEM) central loop
soundings; and (3) solving for static shifts as parameters in inversion.
The EMAP method, developed by Torres-Verdin and Bostick (1992a, 1992b),
consists in measuring electric fields using a continuous profile of electric dipoles placed
end-to-end. orresponding magnetic fields and, optionally, orthogonal electric field
measurements, are collected at discrete intervals. The static shift is removed from the
data by spatially low-pass filtering the electric field. It exploits the fact that the secondary
electric field due to the near-surface heterogeneity integrated along the profile path is
21


zero mean. This type of survey requires extensive field operations and continuous
sampling of the electric field (difficult) (Sternberg et al., 1988). While similar to the
traditional MT method, Pellerin and Hohmann (1990) consider the differences in field
operation and data processing required for EMAP significant enough for it to be
considered a distinct method from traditional MT.
Of the techniques most applicable for correcting MT data in the context of
regional 3-D interpretations, TEM measurements are the most commonly used. TEM
central loop soundings are relatively free from galvanic distortion and can be used at the
MT sounding location to estimate an average 1-D shallow structure (Pellerin and
Hohmann, 1990). Sternberg et al. (1988) demonstrated using TEM central loop soundings
to shift MT soundings, by matching the curves in the frequency range of overlap, and by
joint inversion. Pellerin and Hohmann (1990) transform TEM sounding data into 1-D MT
sounding by calculating the MT response at high frequencies for the 1-D shallow
subsurface structure obtained from inversion of the TEM sounding, and shift the
observed MT response using the computed MT response. Meju (1996) advocates joint
inversion of TEM and MT phase data, similar to Sternberg et al. (1988). Alternative
methods for computing the MT response used for curve-shifting are to use the shallow
subsurface structure obtained from DC resistivity soundings (Spitzer, 2001), or to use a-
priori geologic information (e.g. J ones, 1988).
An alternative to correcting the data for the presence of static shift is to consider
static shifts as parameters to be solved for in the inversion for subsurface structure.
deGroot-Hedlin (1991) introduced the concept of including static shifts as parameters to
be solved for in 2-D inversion, along with subsurface conductivity structure. To stabilize
22


the inversion, a zero-sum constraint was applied to the static shift. deGroot-Hedlin (1995)
modified the zero-sum stabilizer such that sum of the shifts can be set to any user-
specified constant, allowing for the possibility of bias. Ogawa and Uchida (1996) solve
for the static shifts and 2-D subsurface structure, imposing a Gaussian-distribution
constraint on the static shift. Sasaki (2004) applies the concept of solving for static shift
and subsurface conductivity distribution to 3-D, allowing the user the choice of whether
to use a zero-sum or gaussian-distribution, zero-mean stabilizer.

3.3 Distribution of shifts
No theoretical basis has been published as to whether the distribution of static
shifts observed for a collection of MT soundings should be Gaussian-distributed, have a
mean of zero or should sum to zero. A zero-sum constraint assumes that static shifts are
zero-mean random perturbations affecting each MT site; as the number of MT sites
increases, the sum of the static shifts should approach zero (deGroot-Hedlin, 1991). This
is argued to be the case for the electric field along a profile (EMAP analysis). The
assumption is considered to be generally reasonable (deGroot-Hedlin, 1995), but assumes
the shifts are random and that there are a large number of sites (Ogawa and Uchida,
1996). It does not take into account the possibility of bias inherent in MT site selection
and does not imply any particular distribution on the static shifts. Ogawa and Uchida
(1996) interpret histograms of static shifts estimated from collocated TEM and MT
soundings, published by Sternberg et al. (1988), to support the assumption of Gaussian
distribution. Sasaki and Meju (2006) analyzed collocated MT and TEM soundings from
various surveys and found "good support" for Gaussian distributions, skewed depending
23


on geological environment. Deeply weathered regions and sedimentary terrains with
conductive overburdens were found to have net negative shifts and regions with recent
volcanics to have net positive shifts.
Sasaki and Meju (2006) investigate the artifacts created by static shifts in MT data
on 3-D inversion results. Without including static shifts as inversion parameters, they
observe that artifacts generated by 3-D inversion are concentrated in the near-surface
structures, and conclude that large-scale conductivity models recovered by 3-D inversion
of MT data are more satisfactory than can be recovered by 1-D and 2-D inversions.
Including static shifts as parameters in the inversion creates fewer artifacts and allows for
a model with lower misfit and smoother convergence. Sasaki and Meju (2006) found that
the Gaussian and the zero-sum stabilizers worked equally well in simulating the synthetic
data and their shifts, which were approximately Gaussian and zero-mean, hence zero-
sum.

3.4 Impedance ratio frequency-dependence
To better understand the conditions under which a near-surface inhomogeneity in
a regionally 3-D structure generates a frequency-dependent shift, the response from a 3-D
geothermal system is simulated, using the idealized model developed by Pellerin et al.
(1996). The model consists of a 25 m reservoir overlain by a thin, 5 m clay cap, in a
200 m half-space host (Figure 3.1). The clay cap is 6 km by 4.8 km laterally, 375 m
thick, with its top at 300 m depth. The reservoir is 2.7 km by 2.7 km laterally, and 5.4 km
thick, immediately beneath the clay cap. Synthetic data were determined for this

24








Figure 3.1 Idealized geoelectrical conductivity structure of a high-temperature
geothermal system (Pellerin et al., 1996), with receiver locations used in forward
modeling. A small inhomogeneity is introduced for modeling shifts.
4
Plan View
..............
.... ..... .....
........ .
Receiver locations I
2
: : : : ::: : :. inhomogenei ty
-2
Reservoir
(250hm.m
Clay Cap
(50hm.m)

Easting (km)
o
2
6
Depth view
CCIft:l:liiiJlEllliiiiiji;ci- Nea rsurface inhomogeneity
Reservoir
250hm.mj

Easting (km)
25


model using the forward modeling option in the modified program version of Sasaki
(2004). The finite difference mesh consists of 77 (x) x 73(y) x 40(z) nodes, with a
minimum node spacing of 150 m horizontally and 35 m vertically. Because the regional
structure is symmetric, receivers are distributed throughout the NE quarter of the domain
only, and are concentrated near the horizontal boundaries of the clay cap and reservoir.
The response is estimated for 255 receivers, at 19 frequencies equally spaced
logarithmically from 1000 Hz to 0.03 Hz. While 1000 Hz is higher than what is typically
used in MT surveys for geothermal targets, this will allow us to see the transition from
including some inductive component to purely low-frequency galvanic-response of when
a near-surface inhomogeneity is included in the regional model.
The regional impedance ratios are constructed and evaluated. Ratios of on- to off-
diagonal impedances were calculated using the synthetic data for the regional model
corresponding to an idealized geothermal system. These ratios are used when determining
the XY and YX mode apparent resistivity data. The frequency-dependence inherent in the
regional impedance ratios at each receiver is measured as the maximum difference across
all frequencies (Figure 3.2). The most frequency-dependent impedance ratios are
observed near the edge of the clay cap parallel to the direction of the measured magnetic
field along the eastern edge for the XY mode and along the northern edge for the YX
mode. The absolute value of the maximum impedance ratio is approximately 1.5.

3.5 Shift frequency-dependence
A near-surface, small inhomogeneity is placed near the northeast corner of the
clay cap, where the regional impedance ratios are highly frequency-dependent, and the
26










Figure 3.2 Plan view of the maximum, across all frequencies modeled, of the impedance
ratios used in calculating XY and YX impedances for the geothermal model of Pellerin
et al. (1996). Boundaries of the clay cap and reservoir are shown.
27


ensuing distorted response calculated. Hypothetical shifts are simulated by including a
small, near-surface inhomogeneity in the regional geothermal system model. The
inhomogeneity consists of a 35 m thick and 150 m x 150 m wide body of contrasting
resistivity located near the boundary of the clay cap (Figure 3.1). Synthetic data were
calculated for the inhomogeneity resistivity contrast of 10 with respect to the surrounding
half-space (20 m, 2000 m). The distorted soundings are shown in Figure 3.3.
Including the inhomogeneity in the regional model causes distortion of both the phase
and apparent resistivity soundings. Apparent resistivity soundings are distorted such that
the sounding for the conductive inhomogeneity is shifted downwards with respect to the
undistorted sounding that for the resistive inhomogeneity is shifted upwards.
At frequencies where the inductive component of the small inhomogeneity
response is negligible, distorted soundings can be approximated numerically using the
regional structure and some estimate of the shift or inhomogeneity parameters. Shifts are
estimated by calculating the arithmetic average of the ratio of the modeled distorted and
regional apparent resistivity data at the three longest periods in the frequency range of
interest, where both the shift and the impedance ratio are approximately constant.
Assumptions inherent in this approximation are that the frequency-dependence of the
impedance ratios can be neglected, and that the inhomogeneity parameters are constant at
the frequencies of interest. The distorted sounding can be corrected to approximate the
regional sounding by dividing the apparent resistivity by the shift for each mode (Figure
3.4). The difference between distorted and regional soundings can be defined as
negligible when the distorted data fall within the envelope formed by the regional data
error. Arbitrarily, a data weight error of 0.01 log10 units is chosen for the apparent
28



Figure 3.3 Distorted apparent resistivity soundings obtained by introducing a shallow
inhomogeneity into the geothermal model, alternately set to be more resistive than the
surrounding near-surface layer or more conductive.
29








Figure 3.4 Regional sounding and distorted sounding after being corrected using a
constant shift, and their normalized difference in apparent resistivity, for the receiver
directly above the inhomogeneity introduced in the geothermal model. Normalized phase
differences are included to evaluate where the shifts are truly static.
Sounding distorted by Sounding distorted by
_10
3
resistive inhomogeneity
10
J
conductive inhomogeneity
E E


m 10
2
10
2
0 0
a:



10'
10'
000000000000
10' 10'
Frequency (Hz)
a:


10'
<{ 10'
10'
Regional , XY
--Regional , YX
o Corrected, XY
Corrected, YX
000000000000
10' 10'
Frequency (Hz)
Difference in loglO(Pa) ' regional and corrected sounding normalized by 0.01
10.
2
... l Ir:i ... 1
10
4
10
2
10 10.
2
10
4
10
2
10 10.
2
Frequency (Hz) Frequency (Hz)
o Normali zed residual. XY
Normalized residual . YX
Difference in t. regional and corrected sounding normalized by 0.66
0
I ;;.:;;.;;;.:;;.:;; .;;;";;.;;'.JI
10
4
10
2
10 10.
2
Frequency (Hz)
.. ,,, .. : .... ,, I
10
4
10
2
10 10.
2
Frequency (Hz)
30


resistivity and 0.66 for the phase, corresponding to good quality MT data. The distorted
phase data are within the data error for frequencies lower than 20 Hz for the XY mode
data and lower than 100 Hz for the YX data (Figure 3.4). Both soundings exhibit phase
distortions at higher frequencies, suggesting that the inhomogeneity has a nonnegligible
inductive response and should not, at these frequencies, be considered to produce a static
shift. The corrected data most differ from the regional data at high frequencies, where
the inhomogeneity has a nonnegligible inductive component also seen in the phase. The
difference between the regional and the corrected apparent resistivity data is generally
less than the data error for frequencies where the inhomogeneity generates a static shift,
despite the frequency-dependent regional impedance ratios observed for this sounding.
This suggests that the inhomogeneity parameters which scale the impedance ratios are
very small, rendering the frequency-dependent contribution of the regional impedance
ratio trivial for this test example.

3.6 Summary
Numerical modeling suggests that including a small, near-surface inhomogeneity
near the 3-D geothermal model developed by Pellerin et al. (1996) distorts neighboring
apparent resistivity soundings which would be measured in its absence. The distortion
can be simulated using inhomogeneity parameters and undistorted impedance ratios.
Provided the frequency of investigation is low enough that induction in the
inhomogeneity is negligible, phase measurements are unaffected and inhomogeneity
parameters are frequency-independent. Frequency dependence of the apparent resistivity
distortion due to varying impedances is observable, but not significant in this example.
31


Soundings corrected using approximate static shifts are within the assigned data errors.
Two approximations are considered, both based on constant shifts. In the first estimate,
the impedance ratio is assumed frequency-independent and the distorted data are
corrected by applying the mean of the difference between the regional and distorted data.
In the second estimate, corrections are calculated using the on-diagonal inhomogeneity
parameters a and d, neglecting parameters b and c and the impedance ratios altogether.
Both estimates are similar to the modeled distorted data, particularly at the longer
periods. This reinforces the validity of the assumption of static shifts implicit when
solving for the shifts along with the 3-D conductivity structure. There may be other
models that show greater frequency dependence of distortion, but these are difficult to
predict a priori.





CHAPTER 4

4FORWARD MODELING

To calculate the MT apparent resistivity and phase at the earth's surface requires
solving Maxwell's equations (ME) for a particular subsurface conductivity structure.
Analytical solutions of ME for arbitrary 3-D conductivity structures do not exist;
numerical approximation is necessary, either in differential or integral form. The most
widely used solution methods are based on finite differences (e.g. Madden and Mackie,
1989; Smith, 1996a; Newman and Alumbaugh, 1995, 2002; Siripunvaraporn et al., 2004,
2005, among others), and on integral equations (e.g. Hohmann, 1975; Weidelt, 1975;
Wannamaker et al., 1984; Zhdanov, 2009, among others). The FD method is attractive
because of the apparent simplicity of its implementation compared to that of the IE
approach and because it generates sparse regular system matrices for solution. The IE
approach is attractive because it requires discretizing only the volume containing the
arbitrary 3-D conductivity structure; the FD method requires discretizing a much larger
volume including the background and extending into the air. A more thorough
description of the methods and their relative attributes can be found in, e.g., Zhdanov
(2009).
Sasaki's program is based on using the finite difference method to solve the
differential form of Maxwell's equations, using staggered grids. The forward problem and
33


its solution using the finite-difference method are described in section 4.1. The resulting
system of equations is solved using a preconditioned biconjugate gradient method,
described in section 4.2, using the divergence-correction technique of Smith (1996b),
described in section 4.3. Sasaki (2001) verified the accuracy of the forward solution by
simluating airborne EM data and comparing them to results obtained using the finite-
difference solution of Newman and Alumbaugh (1995).

4.1 Finite difference solution
Conductivity structures can be represented as consisting of a domain of
anomalous conductivities, superimposed on a horizontally-layered background. Total
fields can be determined by summing background fields, (E
p
,H
p
) analytically calculated,
and scattered fields (E
s
,H
s
), caused by the anomalous domain (E
s
,H
s
), which must be
numerically approximated. From equation 2.26, the scattered fields in ME in 3-D
domains can be written in terms of the background field:

( )
p p s
i k E E ) (
2
= +

To solve for the scattered fields, the conductivity structure is discretized into rectangular
cells; conductivity is assumed to be homogeneous inside each cell. Derivatives can then
be approximated as finite differences. The scattered electric fields are solved for, and
magnetic fields are calculated from the solution using ME.
The finite-difference scheme is implemented using staggered grids, pioneered by
Yee (1996) for use in electromagnetic modeling. Electric and magnetic field components
4.1
34


are estimated at different locations forming grids staggered with respect to one another.
First-order central differences are used to represent differentiation operations. The appeal
of this method resides in discretized curl, divergence and gradient operators which hold
true (Smith, 1996a). For a given cell of homogeneous conductivity, electric fields are
measured at the centers of the cell edges; E
x
are measured at the centers of the cell edges
parallel to the x-direction, E
y
at the centers of edges parallel to the y direction and E
z
at
the centers of edges parallel to the z direction. Magnetic fields are measured at the centers
of the cell faces, in a similar manner. Note that it is also possible to sample electric fields
at the cell-face centers and the magnetic fields on the cell-edge centers (e.g. Smith,
1996a). Sasaki (2004) tested both configurations and found no significant difference in
forward modeling responses between the two. Where adjacent cells are of different
conductivity, a weighted average conductivity is used. Weights are proportional to the
cross-sectional area of each cell that is normal to the sampled component (Sasaki, 2001).
The resulting system of equations expressed in matrix notation is:

s K =
s
f

where K is the system matrix and s is the source vector. The diagonal entries of K are
complex; off-diagonal entries are real. The system matrix is made symmetric by scaling
each row and the source vector entry by the product of the grid spacings spanning the
sampling position (Smith, 1996a). The source vector depends on the source polarization
and the boundary conditions (Sasaki, 2004). The condition applied is that the scattered
electric field is negligible at the edges of the finite-difference mesh, such that the total
4.2
35


electric field, tangential to the edge of the mesh, is continuous across it. Implicit in this
condition is that the edge-most finite difference cells define the horizontally-layered
background structure, and that lateral conductivity variations must be sufficiently distant
from the boundary. Hence, the finite-difference method requires a large mesh that
extends far past the volume of interest (Zhdanov et al., 1997). Madden and Mackie
(1989) discuss the design of finite-difference meshes for the MT forward problem. In
order to satisfy zero boundary conditions, the distance to the boundaries of the finite-
difference domain must be greater than the largest skin depth considered.

4.2 Biconjugate gradient method
The scattered field solution is obtained by solving the system of equations using
the biconjugate gradient (BiCG) method, preconditioned with incomplete Cholesky
decomposition (Mackie et al., 1994; Smith, 1996b). The system matrix is symmetric, but
because of the complex entries on the diagonal, it is not Hermitian and the CG method
cannot be directly applied. The BiCG method is a generalization of the CG method, with
the conjugate transpose replaced by a simple transpose (Smith, 1996b). Note that the
BiCG method is different from the CG method in that the residual is not monotonically
decreasing. To account for this, the scattered field solution is updated only when the
residual decreases.

4.3 Divergence correction
Solution for the scattered fields is improved by including Smith's (1996b)
correction procedure that enforces divergence-free conditions on current density.
36


Convergence of the scattered field solution tends to degrade at low frequencies. As the
frequency falls, both the spatial derivatives in the system matrix and the frequency-
dependent source vector become small. The solution becomes computationally inexact,
introducing current sinks or sources throughout the finite-difference mesh. Smith (1996b)
developed a procedure which essentially cancels these artifacts. The procedure is applied
to the updated scattered field solution, within the outermost loop of the biconjugate
gradient scheme, prior to iterative refinement. The correction vector (E
c
) is obtained
from:

( ) ( )
s c
E E =

The correction vector and the source vector are complex; the divergence is a linear
operator. Real and imaginary corrections are computed separately using the CG method,
with incomplete Cholesky decomposition preconditioning. The corrected scattered field
can then be calculated as the sum of the correction and the scattered field solution.
4.3


CHAPTER 5

5INVERSION SCHEME

The subsurface conductivity structure sampled in MT surveying can be
represented approximately as a set of model block parameters with heterogeneously
distributed conductivities. These conductivities can be deduced by inversion of MT
measurements. The inversion process for MT measurements is an iterative process
consisting of calculating the predicted data for some model, determining the misfit
between the predicted and observed data, modifying the model to reduce the misfit, and
repeating these steps. The process is terminated when some minimum error or maximum
number of iterations is reached. The starting model can be estimated from a priori
structural information or data averages.
There have been several computer codes generated for 3-D MT inversion (e.g.
Sasaki, 2001; Zhdanov and Golubev, 2003; Mackie and Watts, 2004; Siripunvaraporn et
al., 2004, 2005). For example, Newman et al. (2005a) introduced a finite-difference-
based 3-D massively parallel inversion algorithm for MT data, which, however, requires
large computing resources and long run times. In recent years, CEMI has developed
several algorithms and computer codes for MT inversion based on fast forward-modeling
approximations and on rigorous integral equation (IE) solvers (Golubev et al., 1999;
Golubev and Zhdanov, 2000; Zhdanov, 2002; Wan, et al., 2006; Green et al., 2008;
38


Zhdanov and Gribenko, 2008). Note that interpretation of MT data can be further
complicated by static shifts. The inversion algorithm of Sasaki (2004) is formulated to
solve the inverse problem simultaneously for static shift and 3-D subsurface conductivity
distribution parameters. The program runs on a personal computer, and can handle
moderate-sized data sets. The inversion is done using the Gauss-Newton method, which
requires calculating Frchet derivatives at each iteration.
Section 5.1 introduces the theory behind the MT inversion scheme implemented
in the program. Sensitivity calculations are described in section 5.2. The inclusion of
static shifts in the inversion scheme is described in section 5.3. In section 5.4, the
program is demonstrated by inverting synthetic data for a model similar to that of the
idealized geothermal system presented by Pellerin et al. (1996).

5.1 Theory
The inverse MT problem is that of approximating the subsurface conductivity
structure corresponding to observed MT measurements. Solving this problem consists of
finding some conductivity structure which minimizes the weighted residual, r

( )
obs pre d
d d W r =

which is the difference between the observed data d
obs
, and the predicted data d
pre
,
normalized by the reciprocal of the data standard deviation, stored in diagonal matrix W
d
.
The data consist of the apparent resistivity, in log
10
-space, and the impedance phase. The
predicted data can be determined according to
5.1
39



) A(m d =
pre


whereA is a nonlinear forward operator, based on Maxwells equations, and m is a vector
of subsurface conductivity model parameters. For the Sasaki-based inversion program
here, the predicted data are calculated using the finite-difference method outlined in
Chapter 4.
Minimizing the misfit is equivalent to minimizing the size of the residual
measured using the L2 norm, known as the misfit functional , where

min ) (
2
2
= = r m .

The normalized RMS misfit (nRMS) between predicted and observed data is calculated
using

( )
N
nRMS
m
=

where N is the number of data. As the predicted data approach the observed data to
within the data standard deviation, the misfit should approach 1.
Minimizing the misfit functional is an ill-posed problem, rendered conditionally
stable by including stabilizing functionals. Popular stabilizing functionals are described
5.2
5.3
5.4
40


in Tarantola (1987, 2005) and Zhdanov (2009). The parametric functional used in the
program includes two stabilizing functionals,

( ) ( )
( )
( ) ( ) ( )
min
2
2
1
2
2
1
2
2
1
2
2
2
1
2
2
2
1
=

+ + +
+
=
c
apriori z C
c
apriori y C
c
apriori x C
m
apriori m
obs pre d
z y x
m m C W m m C W m m C W
m m W
d d W m



Terms and are regularization parameters, which control the trade-off in fitting the
residual and the stabilizers. The first stabilizer, scaled by , is of the form m , and acts
to restrict the magnitude of the conductivity parameters in the new model to adhere to
those of the a priori model. This stabilizer may function as a Marquardt factor by using
the model from the previous iteration as a priori. The second stabilizer, scaled by , is of
the form ( ) m where banded matrices C
x
, C
y
and C
z
perform first-order differencing
between adjacent parameters. This stabilizer acts on the conductivity gradient, or spatial
smoothness, to match that of the a priori model. Matrices
z y x
C C C
W W W , , and
m
W are
model weighting matrices. These stabilizers provide a maximum-smoothness type of
model.
Sasakis program uses the iterative Gauss-Newton inversion scheme.
Minimization of the residual can be expressed in the context of minimization of the
parametric functional, accomplished by calculating the first variation of the parametric
functional with respect to the change in model parameters m and setting it to zero
(Zhdanov, 2002). This yields the following equation:
5.5
41


( )
( ) ( ) ( )
( ) ( )
( ) ( )
( ) ( )
priori a z C
priori a y C
priori a x C
apriori m
obs
T
m
z
y
x
A

+
+
+

+
+ =
=
m m C W m
m m C W m
m m C W m
m m W m
d m m F m
m
m
1
2
1
2
1
2
1
2
0
1
, 2
, 2
, 2
, 2
, 2
0
0





where m
0
is the previous model, and m
1
the updated model. The forward operator is
linearized by introducing the Frchet derivative matrix F, a J acobian:

) ( ) ( ) (
0
0 0
m F m m m + +
m
A A
where
( )
0
0
0
m
m
F A
m

= .

This leads to the system of equations

[ ][ ] [ ] b A m J J
T T
=
1

where
( )

+
=

=
apr m
apr z C
apr y C
apr x C
obs pre d
m
z C
y C
x C
d
z
y
x
z
y
x
m W
m C W
m C W
m C W
d d m F W
b
W
C W
C W
C W
F W
J
m m

0 , 0
0 0
; .

5.6
5.7
5.8
5.9
5.10
42


As the approximate regularized Hessian, matrix J
T
J , is square and symmetric, the system
of equations is solved using the LDL
T
method. The updated model can then be used at the
next iteration as the previous model. Implicit in linearizing the forward operator, the
change in model parameters is limited so the Frchet derivative matrix must be re-
calculated for each iteration. Inversion schemes other than Gauss-Newton such as
steepest descent and conjugate gradients have also been used successfully.
At each iteration, the regularization parameters control how closely the updated
model adheres to the class of models imposed by the stabilizers. Various techniques for
selecting regularization parameters have been developed (e.g. Zhdanov, 2002). Rather
than using a full regularization approach in this implementation, regularization
parameters can be selected at each iteration by a line search to determine the parameter
which results in the lowest misfit (Sasaki, 2001; Sasaki and Meju, 2006). The nRMS
misfits of three models, each obtained for different regularization parameters, are
calculated. Starting from a model typically distant from the true, coupled with the
requirement for few iterations driving the choice of regularization parameters, the misfit
functional may not decrease monotonically with decreasing regularization parameter as
illustrated in Zhdanov (2002). The parameter estimated to provide the lowest nRMS
misfit is selected to be used for that iteration of the inversion. As this is not the rigorous
approach to regularization as described in Zhdanov (2009), there is no guarantee that the
parametric functional, normalized misfit and regularization parameter will decrease
monotonically (deGroot-Hedlin and Constable, 1990).
The non-zero diagonal entries of the model weighting matrix applied to the
magnitude stabilizer are calculated according to
43



( )
4 1
F F diag
T
m
= W

as recommended in Zhdanov (2002), with a floor of 0.05. The weighting matrices
z y x
C C C
W W W , , applied to the gradient stabilizer are based on averaging the weights applied
to the magnitude stabilizer for inversion cells beneath the receiver.

5.2 Frchet derivatives
The Frchet matrix consists of the derivatives of the apparent resistivity and phase
with respect to the model parameters. These derivatives can be expressed in terms of the
derivative of the impedance:

( )
( )
( )
( )
pre
i
i
b
i
i
i
i
b
pre
i
b
i c
i
b
pre
i a
Z
dm
dZ
Z
Z
Z
dm
d
dm
dZ
Z
dm
d

2
,
cos
Re
Re
Re
Im
Im
Re
2

=


The derivative of the impedance can in turn be determined from the derivative of the
fields.

HH
dHH Z dEH
dm
dZ
ij ij
b
ij

=
where

5.11
5.12
5.13
44


( )
2 1 2 1 2 1 2 1
2 1 2 1
s
x
s
y
s
x
s
y
s
y
s
x
s
y
s
x
s
x
s
y
s
y
s
x
dH H H dH dH H H dH dHH
H H H H HH
+ =
=

and

2 2 2 2 1 2 1 2
1 2 1 2 2 1 2 1
1 2 1 2 2 1 2 1
2 1 2 1 1 2 1 2
s
x
s
y
s
x
s
y
s
x
s
y
s
x
s
y yy
s
y
s
x
s
y
s
x
s
y
s
x
s
y
s
x xx
s
y
s
y
s
y
s
y
s
y
s
y
s
y
s
y yx
s
x
s
x
s
x
s
x
s
x
s
x
s
x
s
x xy
dH E H dE dH E H dE dEH
dH E H dE dH E H dE dEH
dH E H dE dH E H dE dEH
dH E H dE dH E H dE dEH
+ =
+ =
+ =
+ =


Field derivatives can be obtained using the adjoint equation approach, formally derived
by, e.g., McGillivray and Oldenburg (1990). The method makes use of the reciprocity
condition of EM data (Parasnis, 1988), and a similar approach is described by deLugao
and Wannamaker (1996). In summary, derivatives at a receiver can be obtained by
determining the electric fields within the entire domain, due to fictitious horizontal
electric and magnetic dipole sources placed at the receiver. Calculating the full derivative
matrix requires finding the predicted data for four dipole sources, for each receiver, at
each frequency. The derivatives with respect to a change in the conductivity of a model
parameter are obtained by multiplying together the dipole field and the MT field, and
integrating over the parameter. To determine the electric fields in the subsurface due to
horizontal electric and magnetic dipoles at the surface, background and scattered fields
are solved for separately. The background field is calculated for a homogeneous
halfspace, of conductivity equal to that immediately beneath the dipole source, in order to
avoid singularities due to the source. Equations for dipoles in halfspaces and their
derivations can be found in, e.g., Spies and Frischknecht (1991). The scattered field can
5.14
5.15
45


then be solved for using the same system matrix as used for the MT forward problem,
and following the same procedure. Once the scattered field solution is obtained, the total
fields are determined by summing scattered and background fields.
Calculating rigorous Frchet derivatives can be the most time-consuming portion
of the inversion process. Once rigorous Frchet derivatives have been calculated for the
first several iterations, they can be approximated at future iterations using Broyden's
method (Broyden, 2000). Developed in 1965, this method of approximation is used to
correct the J acobian when the J acobian is not expected to change significantly from one
iteration to the next. The approximation is given by

( )
( )
( ) ( ) m m
m
m F r r F F


+ =
T
T
i i i i i 1 1 1

where
i b i
i i
m
m m m
10
1
log =
=



To limit run-time, the user of the algorithm as modified herein can specify optionally
after a minimum of two model parameter updates that later iterations are calculated with
approximate J acobians determined using Broyden's method.

5.3 Static shift
Assuming that near-surface, small-scale inhomogeneities generate frequency-
independent, real-only shifts, predicted data can be expressed in log space as;
5.16
5.17
46


Gs m
Gs d d
+ =
+ =
) A(
R
pre
d
pre


where
d
pre
d is a vector of predicted data, composed of the logarithms of apparent
resistivities and phases; ) A(m d =
R
pre
are the logarithms of the predicted apparent
resistivities and phases, due to the regional structure; s is a vector of static shift
parameters; and G is a book-keeping matrix of ones and zeros, that relates corresponding
predicted apparent resistivity values and static shifts (Sasaki, 2004). For each entry in d
pre

which is an apparent resistivity, the corresponding row in G will contain a 1, with
remaining entries 0. The 1 will be in the column corresponding to the entry in s which
contains the shift applied to that apparent resistivity and to all other apparent resistivity
entries in d
pre
, which are for the same receiver and mode, at different frequencies. For
entries in d
pre
which are phases, the corresponding rows in G contain only zeros.
To solve the inverse problem for both subsurface conductivity model parameters
m and static shift s, and to constrain the solution, the following objective function is
defined:

( )
( ) ( ) ( )
( )
2
2
2
2
2
1
2
2
2
1
2
2
1
2
2
1
2
2
2
s m m W
m m C W m m C W m m C W
d d W

+ +

+ + +
=

apr m
priori a z C priori a y C priori a x C
obs
d
d
z y x
pre
.

The fourth and final stabilizer is applied to the static shifts, enforcing a Gaussian
distribution with zero mean:
5.18
5.19
47



( )
2
2
2
2

= s s .

Note that how strictly the stabilizer is enforced is determined by the regularization
parameter applied to the constraints. Increasing drives both the size of the static
shifts and their mean to zero as well as binding them more tightly to a Gaussian
distribution.
This objective function, minimized using a direct Gauss-Newton scheme, yields
the following system of equations (Sasaki, 2004):

[ ] [ ] b J
s
m
J J
T T
=

1
1

where

( )

+
=

0
;
0
0
0
0
0
1 , 1
1
1
apr m
apr z C
apr y C
apr x C
obs i pre i d
m
z C
y C
x C
d d
z
y
x
i
z
y
x
i
m W
m C W
m C W
m C W
d d m F W
b
I
W
C W
C W
C W
G W F W
J
m
m



The system of equations is solved using the LDL
T
method, as when static shifts are not
included.

5.20
5.21
5.22
48


5.4 Modifications from original
Modifications were made to the original program, other than restructuring and
parallelization. In the original program by Sasaki, the model parameter update is obtained
by solving a system of equations of the form

[ ] [ ] b
s
m
J =

1
1


using the Gram-Schmidt method, rather than the system described in equation 5.21,
solved using the LDL
T
method. One advantage of working with the system in equation.
5.23 is that the matrix-matrix multiplication to calculate the approximate Hessian is not
required. A disadvantage is that the number of rows in J is proportional to the number of
data as well as to the number of model parameters, while the Hessian is proportional to
the number of model parameters. The number of rows in both are proportional to the
number of parameters. Working with the LDL
T
in equation 5.21 allowed straightforward
application of parallelization concepts for the parameter step on a multicore workstation.
The parametric functional and stabilizers used in the modified program have been
altered from the original, which was of the form:

( ) ( ) ( ) ( ) [ ]
2
2
1
2
2
1 1
2
2

apriori m apriori obs pre
m m W m m m C d d W + + + = .

The smoothness stabilizer initially used was based on minimizing the rate of change
(Laplacian) between adjacent model parameters. This has been replaced by minimizing
5.23
5.24
49


the change in adjacent model parameters with respect to the change in adjacent
parameters of the apriori model (equation. 5.5) (cf. Rodi and Mackie, 2001). The a priori
model could be set to either remain fixed using the starting model or to be updated during
the inversion. Smoothness and adherence stabilizers are decoupled in the modified
version and may refer to either fixed or updated a priori model.
The model weighting matrices used by the original program rely on a complicated
function based on model parameter sizes and depths. In the modified program, model
weighting matrices are based on the diagonal of the approximate Hessian (equation.
5.11), as recommended in Zhdanov (2002).
A data weighting matrix based on data errors read from the input file was
introduced into the program. Data weights can then reflect the confidence interval of the
data. The user can exclude specific data points by flagging the data weight using a large
error. As well as carrying out joint XY and YX mode inversions, the program has been
modified to invert a single principal mode at a time, or any combination of apparent
resistivity and/or phase data.
Various other minor modifications have been made throughout the program, in an
effort to customize it as needs arose. For example, it has been modified to read in a 3-D
conductivity structure and carry out forward modeling only, a useful feature when
preparing synthetic models for testing.

5.5 Synthetic test
To demonstrate the inversion, a synthetic model based on the conceptual
geothermal system of Pellerin et al. (1996) was inverted. The model consists of a 5 m
50


clay cap and 25 m reservoir, embedded in a 200 m halfspace (Figure 5.1). The
dimensions of the clay cap are 2.6 x 2.2 x 0.4 km, with the top at 300 m. The reservoir is
immediately beneath, with dimensions of 1.4 x 1.0 km x 3.3 km. XY and YX mode data
were predicted using the forward modeling option of the program, for 30 stations at 14
frequencies equally spaced logarithmically from 100 to 0.25 Hz. Errors with a Gaussian
distribution and standard deviation of 0.02 log10 (m) and 1.32 were applied to the
data.




Figure 5.1 Model used to generate synthetic data. MT stations are shown as dots in the
plan view.
51


The synthetic data were calculated using a finite difference mesh with 63 x 55 x
43 nodes. A minimum horizontal spacing of 100 m was used in a 4.2 km by 3.4 km
region centered on and encompassing the domain of anomalous conductivity, the spacing
doubling as distance away from this region increases (e.g. Madden and Mackie, 1989).
The edge-most nodes are 38.7 km from the central region. Vertically, node spacing
increases from 40 m immediately beneath the surface to 460 m at 6 km depth, thereafter
doubling until a maximum depth of 34.5 km is reached. Node spacing above the air-earth
interface increases from 40 m to 20.4 km. The mesh size was selected to be large enough
to satisfy the boundary conditions of the lowest frequency data while retaining
sufficiently small node spacing in the near-surface zone of anomalous conductivity for
predicting higher-frequency data. For an average resistivity of 100 m, the depth of
penetration is approximately 500 m at 100 Hz and 10 km at 0.25 Hz.
The synthetic data were inverted using a similar finite difference mesh with 63 x
55 x 43 nodes, and an inversion domain containing 29 x 25 x 26 model parameters.
Model parameters have minimum dimensions of 200 m horizontally and 100 m
vertically. The smoothness stabilizer was fixed at 2 for the first model update, and
allowed to vary at later iterations. The adherence stabilizer was fixed at 1 for all
iterations. Rigorous Frchet derivatives were calculated for the starting model and
subsequent two iterations with Broyden approximations thereafter. The starting model
was a halfspace of 105 m, determined by the program based on the input apparent
resistivity data. An nRMS misfit of 1.03 was achieved after 6 model updates (Figure 5.2).
Misfit was observed to increase at the subsequent iteration and the inversion terminated.

52



Predicted and observed soundings from the MT station in the southwest (#1) and
from near the center (#15) of the survey are shown in Figure 5.3. The recovered
conductivity structure agrees relatively well with the true model (Figure 5.4). The clay
cap appears to be the most prominent feature and the conductivity structure
corresponding to the reservoir is more diffuse than the true model, similar to observations
made by Pellerin et al (1996).
Static shifts with a Gaussian distribution and standard deviation of 0.12 log
10
(m)
were applied to the synthetic data, along with errors. These data were inverted using a
similar finite difference mesh and inversion domain as previously used. The
regularization parameter controlling the smoothness stabilizer was fixed to a value of 2
for the first model update, and allowed to vary at later iterations; that for the adherence
stabilizer was fixed to 1; and that applied to the static shift stabilizer was fixed to 0.7.
These values were determined by trial and error. Rigorous Frchet derivatives were
calculated for the starting model and subsequent two, with Broyden approximations
Figure 5.2 Convergence observed for inversion of the synthetic data.
53





Figure 5.3 Observed and predicted soundings from MT stations 1 and 15.
Sounding for Central MT station (15)
10
3
,----------------------,
10
2
10' 10
Frequency (Hz)

Observed data, XY Mode
Observed data, YX Mode
Predicted response, XY Mode
--Predicted response, YX Mode
en
{l60 '''=.-=,
40
'"
it 20
010
2
10' 10
u
Frequency (Hz)
Sounding for SouthWest MT station (1)
10' 10
u
Frequency (Hz)
Observed data, XY Mode
Observed data, YX Mode
Predicted response, XY Mode
--Predicted response, YX Mode

Cl
<1>60
E --11
40 ,I , 0_
'"
it 20
010
2
1 0' 1 0
Frequency (Hz)
54




thereafter. The starting model was a halfspace of 107 m, determined by the program
based on the input apparent resistivity data. The target nRMS misfit of 1.0 was achieved
after 7 model parameter updates (Figure 5.5). The increase in misfit for the 4th model
parameter update is attributed to the use of a nonrigorous regularization inversion
approach. Predicted and observed soundings from the MT station in the southwest (#1)
and from near the center (#15) of the survey are shown in Figure 5.6. The recovered
conductivity structure agrees relatively well with the true model (Figure 5.7) and the
predicted model from inversion of the synthetic data with errors only (Figure 5.4). The
recovered shifts have a mean of -0.03 and standard deviation of 0.054 log
10
(m),
approximately half that applied (Figure 5.8). The near-surface inversion domain is
discretized sufficiently finely to generate the remaining amount of distortion not
accounted for in the recovered shifts (Sasaki and Meju, 2006).

Figure 5.4 Conductivity structure recovered by inversion of the synthetic data. The
boundaries of the clay cap and reservoir from the true model are shown. Slices are at
0.15, 0.55, 1.0 2.0 2.75 and 4.0 km.
55















Figure 5.5 Convergence observed for inversion of the synthetic data with shifts.
16
14
12
10
en
::2:
8
a:
c:
6
4
2

0
2 3 4 5 6 7 8
Model
56








Figure 5.6 Observed and predicted soundings from MT stations 1 and 15.
Sounding for SouthWest MT station (1)
10
4
.---------------__ -----.
I Observed data, XY Mode
I Observed data, YX Mode
--Predicted response, XY Mode
~ 10
3
--Predicted response, YX Mode
'"
Q)
a::
c.
c.
~ 8
Ol
10
2
~ 6
~ 4 ~ ~ : I
10
2
1 0' 1 OU
Frequency (Hz)
'"
co
5: 20
010
2
10' 10
Frequency (Hz)
Sounding for Central MT station (15)
10
3
.----------------------.
10
2
10' 10
u
Frequency (Hz)
I Observed data, XY Mode
I Observed data, YX Mode
--Predicted response, XY Mode
--Predicted response, YX Mode
10' 10
u
Frequency (Hz)
57






Figure 5.7 Conductivity structure recovered by inversion of the synthetic data distorted
by shifts. The boundaries of the clay cap and reservoir from the true model are shown.
Slices are at 0.15, 0.55, 1.0 2.0 2.75 and 4.0 km.
Figure 5.8 Histogram of the original shifts applied and those recovered from inversion.


CHAPTER 6

6PARALLELIZATION

6.1 Introduction
Expanding geophysical datasets and the drive to full 3-D interpretations are
increasingly demanding of computing resources. To meet this demand, 3-D MT inversion
programs were developed for distributed computing clusters (Newman and Alumbaugh,
2000; Wan et al., 2006; Green et al., 2008). Newman and Alumbaugh (2000) developed a
finite-difference-based, conjugate-gradient inversion implemented on a parallel
computing platform with tens to thousands of processors. Several programs have been
developed by CEMI, primarily based on the integral equation method and generally
implemented on powerful computing clusters (e.g. Wan et al., 2006; Green et al., 2008).
An alternative to computing clusters are multicore computers. While not as powerful,
they may be sufficient for routine 3-D inversion of modest datasets and are easily
available and affordable. Zhdanov and Gribenko (2008) developed an integral-equations,
conjugate-gradient-based 3-D inversion program for MT and magnetovariation data using
a multicore computer.
As the performance of single-processor PCs has seemed to plateau, multicore PCs
have evolved as a way to increase the computing performance available in a desktop
package. Unlike the traditional PC consisting of a single CPU on a chip, multicore
59


computers rely on multiple CPUs ("cores") sharing the same chip. It is anticipated that
the number of cores will double with each processor generation (Buttari et al., 2007a).
The multicore design is not equivalent to symmetric multiprocessing systems in that,
while both support parallelization, the CPUs present on a multicore chip may share on-
chip resources and the system bus. The architecture of the multicore computer is designed
to allow a program to make use of the available resources. Significant modifications to
software may be required for a serial implementation to effectively exploit the improved
performance available from parallelization for these multicore machines (Buttari et al.,
2007a). Auto-parallelization, while attractive, is unlikely to yield improvements as large
as can be obtained by restructuring and explicit parallelization.
Sasaki (2004) developed a program to invert MT data, using the Gauss-Newton
method, which runs on a personal computer and can handle moderate-sized data sets. The
program, originally developed for serial implementation, has been modified to be able to
use it effectively on a multicore computer. The inversion process can be accelerated by
implementing parallel regions to be executed concurrently on several threads. The
program, written in Fortran77, has been modified to allow fork-join parallelization in
several areas, using the OpenMP 2.0 application programming interface developed for
shared-memory platforms such as multicore PCs. The OpenMP directives are embedded
in the fortran code. When a construct identifying a program region to be executed in
parallel is encountered, it creates a team of parallel threads (fork), no larger than the
number of cores, which execute concurrently and which, upon completion of the region,
terminate (join). Two key areas of the program that were parallelized are the frequency
loop containing the forward modeling and sensitivity calculations, and factorization of
60


the approximate regularized Hessian, required for determining the new model parameters.
The program was tested on a Intel Xeon 5355, 2.66 GHz, double quad-core PC, with 16
GiB of RAM. The modifications do not impair serial performance of the program, which
can be run with and without the parallelization enabled.

6.2 Frchet derivatives
The frequency loop containing the forward modeling and sensitivity calculations
is readily amenable to coarse-grained parallelization using OpenMP, as there are no
dependencies between frequencies. In the serial implementation of the program, the
predicted data and the sensitivities are both calculated for the same frequency, before
proceeding to the next frequency. The modified program assigns to a separate thread the
forward modeling and sensitivity calculations pertaining to a single frequency, until all
frequencies have been processed. Threads execute simultaneously and asynchronously.
The subroutines controlling the forward modeling and sensitivity calculations are thread
safe, such that all shared variables are either frequency indexed or left unmodified. The
sensitivity and residual arrays are held privately. A key feature in the simplicity of this
implementation is that the full sensitivity matrix is never formed. The full sensitivity
matrix can be considered as a column of submatrices, each submatrix containing the
sensitivity terms for a particular frequency. The subroutine in which the sensitivity terms
are formed stores only the submatrix entries for a single frequency; once complete, these
entries are written to a scratch file for retrieval when generating the Hessian, and the
array cleared. The residual array is handled similarly, where subarrays of residual entries
pertaining to a single frequency are calculated, and written to a temporary scratch file.
61


Once all frequencies have been processed, the program exits the parallel section and the
misfit is calculated. The F
T
F term of the Hessian is generated in double precision in a
separate parallel section after the frequency loop. The diagonal and triangular entries of
the Hessian matrix are stored in single precision in a 1-D array, of length NP(NP-1)/2,
where NP is the number of parameters.
The time savings achieved by parallelization of the frequency loop is
demonstrated by applying the program to calculate the 3-D forward model and
derivatives for a synthetic dataset generated for 100 receivers, using a 35 x 35 x 29 node
finite-difference mesh. These synthetic data correspond to the test model used by Sasaki
(2004), consisting of a 5 Ohm-m brick, embedded in a 100 Ohm-m vertical quarter space
bound by an adjacent 10 Ohm-m quarter space. Tests were carried out by varying the
number of CPUs active. The time required to calculate the predicted data and derivatives
will vary depending on the frequency considered; as such, tests were carried out with all
frequencies assigned identical values, and repeated for different values.
Increasing the number of CPUs decreases the time required to process the
frequency loop (Figure 6.1). However, speedup appears limited. Simulated field
calculations are dominated by sparse-matrix-vector computations implemented
conventionally using compressed storage. Performance in such tasks is limited largely by
the processor-to-memory performance gap (Buttari et al., 2007); these computations on a
single processor usually run at 10% or less of the peak floating point performance (Vuduc
et al., 2002). With several processes sharing resources, data congestion further degrades
performance.

62













Figure 6.1 Time required (top panel) and speedup (bottom panel) observed when
calculating the predicted data, residual and Frchet derivatives at 8 frequencies for a
synthetic dataset. Speedup is calculated as the ratio of execution times required, when
using one CPU and when using several.
63


Determining the derivatives for a particular receiver, at each frequency, requires
calculating the scattered fields for four different dipole sources:

[ ] [ ]
i i
s E K = .

The system matrix, K, is the same for all receivers at that frequency, while E
i
and s
i

denote the scattered field and source vector for dipole source i. In the original
implementation, the scattered field solution was solved for each source, one after the
other, using the biconjugate gradient (BiCG) and the divergence correction method of
Smith (1996a). Since matrix-matrix operations are more memory efficient than matrix-
vector multiplications, the processes of solving for the scattered fields due to each dipole
source are amalgamated, as:

[ ] [ ]
4 3 2 1 4 3 2 1
s s s s E E E E K = 6.2

furthermore storing field and source matrices compactly as 1-D arrays. Similar
modifications were implemented in solving for the predicted data and in the divergence
correction procedure. These modifications result in a reduction of the time required to
process the frequency loop, and a modest improvement in speedup (Figure 6.1).

6.3 Parameter update
Parallelization has been applied in factoring the approximate regularized Hessian
for the model parameter update using the LDL
T
method, a variant of the LU method
6.1
64


specialized for symmetric matrices. The algorithm implemented is based on a block-
cyclic column layout similar to that used by Baboulin et al. (2005), within a simple
construct consisting of OpenMP locks (Figure 6.2). The matrix is subdivided into smaller
blocks; work propagates from block to block both horizontally and vertically. The
algorithm is structured to maximize reuse of the data in a block once it is loaded from
slow memory into faster, higher level memory, thus reducing data movement and
maximizing cache use (Buttari et al., 2007b). Each thread is assigned one column of
blocks, starting from the leftmost and progressing sequentially through to the rightmost
column. Once the leading triangle and adjacent block of a row have been operated on, the
algorithm assigns work to proceed vertically down to the leading triangle beneath, or
horizontally to adjacent blocks. A column of blocks must be completely processed,
before releasing the CPU to a queued column. Numerical accuracy in factoring is
maintained through double-precision operations and rolling sums. OpenMP locks ensure
that blocks are not operated on prematurely by requiring that each thread acquires
ownership of the leading triangle prior to working on a block of the same row, stalling
until the lock becomes available. The simplicity of this implementation is possible
because pivoting is not required. The time required to factor matrices of different sizes
was tested using 1,2, 4 and 8 processors (Figure 6.3). Doubling the number of processors
used to factor the system matrix approximately halves the time required. The NP
3

dependence of runtime, consistent with O(NP
3
) operations required by LDL
T
factoring, is
preserved (Figure 6.4).


65




Figure 6.2 Schematic of workflow for the LDL
T
factorization, for 4 CPUs. Work
proceeds from panel a through panel d. For each panel, blocks which are unavailable as
yet to process are colored white; those in which processing has been completed are
colored black.
66






Figure 6.3 Speedup observed when factoring the Hessian matrix using the LDL
T
method.
Figure 6.4 Time required to factor a matrix using the LDL
T
method, as a function of the
number of parameters and the number of CPUs active.
67


6.4 Summary
Substantial increases in efficiency have been achieved for 3-D MT inversion by
parallelizing an original serial algorithm to run on multicore PCs. Parallelization has been
achieved by assigning individual frequencies to separate threads, each executing the
forward modeling, residual and Frchet derivative calculations for the assigned
frequency. Increasing the number of CPUs increases the number of frequencies which
can be processed at the same time and overall, decreases the time required for the
inversion. However, speedup is limited due to lags in data transfers; newer processor
architectures are expected to improve this considerably (Barker et al., 2008).
Parallelization of the factorization of the approximate regularized Hessian is
accomplished using a block-cyclic algorithm and compact storage. Doubling the number
of CPUs used in factoring approximated halves the time required, at least for the 8-cpu
platform considered. These modifications have significantly reduced the time required to
complete a 3-D MT inversion on a single-box workstation format.


CHAPTER 7

7INVERSION OF MT DATA FROM THE COSO
GEOTHERMAL FIELD

The MT method has been successfully used since the 1970s to image subsurface
electrical resistivity in complex geothermal systems, detecting electrical resistivity
variations related to the presence of fluids and conductive alteration minerals (Berktold,
1983; Meju, 2002). Geothermal energy is typically converted to electrical energy by
drilling into a high-temperature hydrothermal convective system, heated by recently
emplaced magma at less than 10 km depth, and using the fluids to drive turbine
generators (Duffield and Sass, 2003). High-temperature geothermal systems are, in many
cases, characterized by geoelectric structures with discernible anomalous zones, which
can be caused by the high-temperature fluids and by alteration of the country rock.
Electrical and electromagnetic methods are found useful in delineating conductivity
contrasts associated with the system, particularly the clay cap, fluid distributions, and the
deeper magma reservoir, as well as structural controls. Review papers by Berktold
(1983), Martinez-Garcia (1992), and Meju (2002) discuss the use of electrical and
electromagnetic methods for high-temperature geothermal systems in more detail.
Characteristics of conceptual models for high-temperature geothermal systems are
discussed in section 7.1.
69


The Coso geothermal field (CGF) is a high-temperature hydrothermal convective
system, in one of the most active seismic regions of southern California (Lees, 2002).
The regional tectonic setting and geology are described in section 7.2. Energy has been
produced from the CGF since 1985. In 2003, 102 MT soundings and a 2-D dense array
profile were collected on the east flank of the field outside of the main production area,
by Wannamaker et al. (2004), as part of the U.S. Dept of Energys Enhanced Geothermal
Systems (EGS) program research (Figure 7.1) (Sheridan et al., 2003). Previously
published interpretations of the Coso data have included 2-D stitched vertical slices
(Wannamaker, 2004; Newman et al., 2005a), and 3-D inversion using a massively
parallel computer of a finely discretized model seeded with a starting guess incorporating
2-D inversion results (Newman et al., 2008). An important structure appearing in these
interpretations is a steeply-dipping conductor most prominent in the southwest East Flank
sector, which is interpreted to reflect fluid-filled fractures. The 3-D MT inversion
program with modifications described in Chapter 6 has been applied to inverting the data
from the 102 MT soundings. XY and YX mode apparent resistivity and phase data were
inverted separately and jointly to test for compatibility. In addition, inversions were
started from a halfspace to determine what the 3-D algorithm is capable of imaging
independently of a starting guess. An additional joint inversion including static shifts as
inversion parameters was carried out to see how well this procedure ameliorates effects
of shallow structural complications. The MT data and inversion results are discussed in
section 7.3. A brief summary of findings is presented in section 7.4.
70



Figure 7.1 Map showing geology and locations of 102 MT soundings collected over the
East Flank of the CGF by Wannamaker et al. (2004). Geology based on Whitmarsh
(2002a, 2002b), modern faults from Davatzes and Hickman (2006), and reservoir outlines
from Adams et al. (2000). Well production and injection intervals provided by F.
Monastero, courtesy of G. Newman and E. Gasperikova.
E
:.
3991


Coso Wash





g' 3988
:;:
t
o
Z
3986 .
427 428 429 430 431 432 433 434
Map Symbols
Easting (km)
MT Station .. \ ... Fault contact. Ball on down-dropped side
./'- Intrusive or depositional contact
I Surface projection of well production (blue) and
__ Modern faults (offset Holocene sediments)
Cover Units
D Vounger Alluvium
i5 D Rhyolite flows and domes ( _ )
E
2
'" ::>
a
Basalt southeast of Devil 's Kitchen
D Jlder Alluvium
southwest of SUQarloaf Mountain
northwest of Petroglyph Canyon
injection (red) intervals, line drawn through top (0)
middle and bottom.
o Approximate outline of reservoir
DK Devil's Kitchen
CHS Coso Hot Springs
WP Wheeler Prospect
Basement Units
_ Calcareous Breccia
'E

_ Felsite Intrusions
D Coso Formation, fanglomerate and air-fall pumice
D Oacite east of Coso Hot Springs
000<>00oo. Spring View Breccia
--- Independence Dike Swarm
u
'iii
D Spring Hill Leucogranite
'E
Basalt of Petroglyph Canyon
Mafic rocks
D Dacite
::>
...,
_ Mixed complex, Intrusives
N '
D Metamorphic rocks of uncertain age
N'
0..
71


7.1 Conceptual model for high-temperature geothermal systems
Geothermal electrical power is typically produced from high-temperature
geothermal systems, where fluid temperatures are greater than 150C (Wright et al.,
1985). Geothermal resources require a heat source, fluids for heat transfer, permeable and
porous reservoirs consisting of fractured host rock and the fluids contained, and a cap
rock restricting both fluid flow away from the reservoir and mixing with cooler near-
surface water (Wright et al., 1985; Spichak and Manzella, 2009). Usually, the heat source
for these systems is recently emplaced magma at less than 10 km depth (Duffield and
Sass, 2005; Spichak and Manzella, 2009). Loss of heat by the intrusion to the
surrounding rock and deep meteoric water results in the development of fluid convection
cells. Interaction between upwelling fluids and the host rock results in the precipitation,
generally strongly zoned, of secondary mineral assemblages. Precipitation of secondary
minerals occludes pores and reduces permeability. An active tectonic environment
contributes to maintaining permeability with the reservoir.
Hydrothermal alteration assemblages are strongly controlled by temperature
(Wright et al., 1985). A commonly used conceptual model of the alteration zones,
appropriate for a permeable host with pervasive alteration, is shown in Figure 7.2
(Pellerin et al., 1996; Oskooi et al., 2005). Three main zones are identified: a clay cap
consisting of smectite layer underlain by mixed smectite-illite layer, and a reservoir with
propylitic alteration. Low temperature clay minerals such as smectite are formed at
temperatures above 70C, disappearing at 220-240C. Chlorite and illite may occur
interlayered with the low-temperature alteration minerals, their proportion increasing
with temperature, especially above 180C. Pure chlorite and or illite remain at
72


Figure 7.2 Conceptual model of high-temperature geothermal system (after Pellerin et al.,
1996).





Reservoir, with
propylitic alteration

Estimated resistivity:
10 - 60 m
Smectite
Estimated resistivity: <10 m


Mixed layer with smectite & illite
Estimated resistivity: <10 m
Clay Cap
Ground surface
Background resistivity:
Unspecified
73


temperatures above 240C, along with other high-temperature propylitic alteration
minerals such as epidote (Spichak and Manzella, 2009).
Smectites are more electrically conductive than illite due to their high cation
exchange capacity; hence, in the geoelectric structure that corresponds to the conceptual
model, the clay cap typically is the most conductive feature (Wright et al. 1985, Pellerin
et al., 1996). The reservoir zone is more resistive than the clay cap, and may be of
intermediate or higher resistivity compared to the host rock. Pellerin et al. (1996)
proposed a geoelectric structure with less than 10 m for the clay cap and 10 to 60 m
for the reservoir. This resistivity structure, strongly controlled by alteration mineralogy,
has been observed, generally in geothermal areas with high permeability and pervasive
alteration (Spichak and Manzella, 2009). Topography or substantial hydrologic gradients
may distort this structure, offsetting the maximum conductivity zone indicative of the
clay cap with respect to the more resistive reservoir. In some contrast to the above model,
geothermal systems in which the reservoir zone is a conductive anomaly have been
observed for systems developed in impermeable rock with hydrothermal fluids restricted
to a few main fractures and faults (Volpi et al., 2003). Very low resistivities at depths
greater than 6 km observed in field studies are speculated to be due to the geothermal
heat source (Meju, 2002).

7.2 Coso Geothermal Field
A geologic map of the CGF is presented in Figure 7.1. This map is a compendium
of geological mapping extracted primarily from Whitmarsh (2002a, 2002b), with more
recently compiled modern faults from Davatzes and Hickman (2006) superimposed; of
74


well production and injection intervals provided by the U.S. Navy (F. Monastero, pers.
comm.) courtesy of G. Newman and E. Gasperikova (Newman et al., 2008); of
approximate geothermal reservoir boundaries based on Adams et al., (2000); and of MT
station locations.

7.2.1 Geologic setting
The CGF covers ~30 km
2
(Adams et al., 2000; Monastero et al., 2005) in the
Coso Volcanic Field in the central Coso Range, close to the south-eastern edge of the
Sierra Nevada Microplate. The CGF is located in the Walker Lane / Eastern California
Shear Zone (WLSZ), between extensional Basin and Range tectonics to the east and
strike-slip tectonics to the west (Roquemore, 1980; McClusky et al., 2001). In this region,
the WLSZ accommodates approximately 11 mm/yr of north northwest-south southeast
trending, right-lateral strike slip motion between stable North America and the Sierra
Nevada. The major eastern tectonic boundary of the Sierran microplate is the right-lateral
strike-slip Owens Valley (OV) fault (Unruh et al., 2002). Dextral shear is believed to
propagate from the OV fault, whose surface expression disappears north of Coso, to the
Airport Lake fault zone, south of Coso. Active deformation has been evidenced by
seismic activity and geodetic measurements. In the vicinity of the CGF, Unruh et al.
(2002) determined from focal mechanism studies a dominantly ESEWNW to nearly E
W extensional direction. Deformation mechanisms observed in the Coso Range include
block rotation, and partitioning of dip and strike-slip faulting (Bartley et al., 2004; Pluhar
et al., 2006; Frankel et al., 2008).
75


Basement consists of Mesozoic granitic and dioritic to gabbroic plutons similar to
those found in southern Sierra Nevada, with minor metamorphic rocks (Duffield et al.,
1980). The CGF is located in the Coso Volcanic Field, an area ~400 km
2
with ~35 km
3

lava flows and 38 rhyolite domes formed in late Cenozoic (Duffield et al., 1980). Two
separate periods of volcanism were identified. The first, lasting from 4 to 2.5 Ma,
produced calc-alkaline lava flows and pyroclastic rocks, principally in Wild Horse Mesa,
east of the CGF. The more recent, lasting from 1 Ma to 40 ka, produced rhyolitic domes
in the central and southern Coso Range. Domes were formed from an increasingly
shallow magma source, either a single rhyolitic reservoir moving upwards or a series of
reservoirs that were progressively more shallow (Duffield et al., 1980). The youngest
dome is Sugarloaf Mountain (Adams et al., 2000), immediately west of the CGF.
Duffield et al. (1980) attributed both the rhyolite domes and the CGF to a body of partial
melt at 8 km depth or more beneath the field.
The brittle-ductile transition beneath the CGF is believed to be elevated by the
emplacement of magma and/or elevated heat flow (Lees, 2002). Mapping of earthquake
hypocenters reveals that the base of seismicity, interpreted to approximate the brittle-
ductile transition zone, is approximately 11 km regionally and shallows beneath the field
to 5 km (Hauksson and Unruh, 2007). The presence of magma beneath the CGF at depths
of less than 10 km has been inferred by various authors (Manley and Bacon, 2000;
Wilson et al., 2003; Hauksson and Unruh, 2007). In addition, a deep self-sealing reservoir
containing a few percent brines, possibly of exsolved magmatic fluids, has been proposed
to exist near the brittle-ductile transition.

76


7.2.2 CGF description
The CGF is subdivided into 3 main regions: the Main Field, East Flank, and Coso
Wash (Davatzes and Hickman, 2006) (Figure 7.1). The Main Field is characterized by
high seismicity rates, high temperatures, and Quaternary rhyolite domes with the
youngest, Sugarloaf Mountain, on the westernmost edge. The East Flank, adjacent to the
Main Field, has a central spine of exposed or shallow bedrock, with low permeability.
Coso Wash is a series of subbasins, separated from the East Flank by en-echelon east-
southeast-dipping, north-northeast striking normal faults. The CGF contains 2 strongly
developed groups of faults. The first group strikes north-northeast and dips to the west
and east; the second group strikes west-northwest and is interpreted to be strike-slip faults
(Davatzes and Hickman, 2006). The first group has been successfully targeted for
geothermal energy production. Surface hydrothermal features are observed at Coso Hot
Springs, Devil's Kitchen, and Wheelers Prospect (Adams et al., 2000). Unruh et al.
(2001) have interpreted seismic reflection data to indicate that the Coso Wash fault,
among others, terminates or soles into a subhorizontal reflector at 4 - 5 km depth,
interpreted to lie at the brittle-ductile transition zone.
Analysis of production drilling indicates the presence of a north-trending zone of
concentrated permeability in the Main Field, separated from the East Flank by low-
permeability rock (Lutz et al., 1996). The Main Field, the primary producing area,
contains several weakly connected reservoirs, with distinct production fluid compositions
and temperatures. Geothermal production wells are concentrated toward the southwest
corner of the field (SW), with significant production from 2 to 3 km depth and high
temperatures (Newman et al., 2008), and at Devil's Kitchen (DK). An upflow zone exists
77


in the southwest corner of the Main Field, with fluids migrating northwards along the
high permeability zone (Adams et al., 2000, Kovac et al., 2005). In the East Flank, wells
are concentrated near the Wheeler prospect and have been drilled westward to intersect
production. J ulian et al. (2010) deduced, from an induced seismicity experiment
corroborated by surface geological observations, that an East Flank north-northeast
striking fault, expected to dip east, actually dips 75 to the WNW. The high-yield well
producing from this fault is located in the hanging wall west of the fault.
Well cuttings have revealed smectite alteration, the base of which lies near 325 m
depth in the northwest and central-west portions of the Main Field, deepening to
approximately 1500 m in the southwest part (Lutz et al., 1996). Kovac et al. (2005)
examined well cuttings from two adjacent East Flank boreholes, and located the base of
smectite alteration at depths of 3000 feet (910 m) and 5500 feet (1,670 m), considerably
greater than the Main Field area. Low-temperature alteration minerals are observed in
present-day high temperature zones of wells in both the East Flank (Kovac et al., 2005)
and Main Field (Lutz et al., 1996). Based on fluid inclusion studies and the distribution of
alteration, Kovac et al. (2005) conclude that the modern CGF is superimposed on and
controlled by the structure of an older, lower temperature system. Preservation of low-
temperature alteration is due to the low permeability of the host rock. According to
Adams et al. (2000), the modern geothermal system was predated by at least two earlier
episodes in the last 300,000 years, also likely to have been driven by intrusions related to
the development of the Coso Volcanic Field.
The preceding description of the CGF indicates that this high-temperature
geothermal system is developed in impermeable plutonic rock with hydrothermal fluids
78


restricted to a few main fractures and faults. As such, the anticipated geoelectric structure
may depart from the conceptual model described in section 7.1, typically observed in
geothermal systems developed in permeable, highly-altered rock. Rather, it may exhibit
features similar to those described by Volpi et al. (2003), in which the reservoir zone is
the most conductive feature.

7.3 Inversion
7.3.1 Data
In 2003, 102 MT soundings and a 2-D dense array profile were collected on the
East Flank of the field, outside of the main production area, by Wannamaker et al. (2004)
as part of the U.S. Dept of Energys Enhanced Geothermal Systems (EGS) program
research (Sheridan et al., 2003) (Figure 7.1). The MT survey was part of an effort to
evaluate the potential of hydro-fracturing in the East Flank as a means to increase
permeability for enhanced production (Sheridan et al., 2003).
The MT survey is located slightly to the east of the Main Field, roughly centered
on the East Flank. MT stations are on average approximately 500 m apart, along profiles
oriented approximately east-west. The survey geometry was chosen such that profiles are
roughly perpendicular to the large-scale north-south oriented horst-and-graben structural
geology and to the anticipated north-south strike of the East Flank reservoir. XY mode
data can then be considered roughly equivalent to TE mode, and YX data to TM mode
for a 2-D analogy. Data were recorded over a nominal frequency range spanning from
250 Hz to 0.01 Hz (Wannamaker et al., 2004). The presence of a DC powerline running
north-south along the west side of the survey area strongly affected the data, ultimately
79


forcing the use of a remote reference in New Mexico to mitigate noise, far more distant
than is normally required. Wannamaker et al. (2004) and Newman et al. (2008) discuss
the measures taken to collect and process the data in more detail.
Typical XY and YX mode apparent resistivity and phase data are shown in Figure
7.3 for two sites. Station E34 is located in the southwest corner of the survey area, near
the eastern edge of the Main Field. Station N32 is located in the northeast corner of the
survey, on Coso Wash. Data error floors of 0.02 log10 (m) for apparent resistivity data
and 1.32 for phase data were applied as representative of the scatter in the responses.
Inversions are restricted to data from 12 frequencies equally spaced logarithmically
ranging from 100 Hz to 0.63 Hz. Assuming a representative resistivity of 40 m, this
corresponds to an approximate depth of investigation of 300 m at 100 Hz and 4 km at
0.63 Hz, sufficiently deep to investigate the geoelectric structure pertaining to the CGF.
A prominent feature of the dataset, exhibited by the majority of the soundings, is a
pronounced apparent anisotropy in the XY and YX mode data toward lower frequencies.
Strongly increasing phase and decreasing apparent resistivity responses are observed in
the YX mode for frequencies at and below 1 Hz and 0.1 Hz, respectively, while the XY
mode response remains relatively resistive.
In view of the apparent anisotropy, XY and YX mode data are inverted both
independently and jointly using the modified version of the code by Sasaki (2004),
described in Chapter 6. Two sets of inversions are carried out. The first set (#1) of
inversions uses the frequency range from 100 Hz to 0.63 Hz, and the second set (#2) is
reduced with 1.6 Hz as the lowest frequency. Assuming a representative resistivity of 40
80



Figure 7.3 Observed apparent resistivity and phase data for MT stations E34 (top) and
N32 (bottom). Data are shown by plotting the error assigned to the data at the measured
value.
10
3
COSO-E34
XY Mode data
YX Mode data
E
9-
~ 10
2
a::
a.
a.

10'
10' 10' 10
10-' 10-'
Frequency (Hz)
80
~
g'60
:s
~ 40
'"
.c
"- 20
0
10' 10' 10
10-' 10-
2
Frequency (Hz)
10'
COSO-N32
E
9-
~ 10
1
a::
a.
a.

I
XY Mode d t ~ I
YX Mode data
10
10' 10' 10
10-' 10-'
Frequency (Hz)
80
g'60 ,
,
:s
~ 40
'"
.c
"- 20
0
10' 10' 10 1 0-' 10-'
Frequency (Hz)
81


m, this corresponds to an approximate depth of investigation of 2.5 km at 1.6 Hz. In
both cases, phase data from the lowest frequency are omitted from the inversion. As
detailed below, fitting the apparent anisotropy toward low frequency in the data with a
single isotropic 3-D resistivity model was not achieved; Newman et al. (2008)
experienced the same result.
Similar finite-difference (FD) mesh and inversion parameters were used for both
sets of inversions. Inversions from set #1 utilized a FD mesh discretized with 89 x 77
nodes in the x (north) and y (east) directions and 46 nodes in the z-direction, of which 10
nodes were in the air, and an inversion domain with 46 x 40 x 30 parameters. FD nodes
within a 6.8 km by 5.6 km zone centered on and encompassing the survey area are evenly
spaced 100 m apart; outer nodes are increasingly further apart. Vertical node spacing
increases from 35 m immediately beneath the air-surface interface to 550 m at 7.9 km
depth, thereafter doubling until a maximum depth of 24 km is reached. Within the survey
area, the inversion parameters are 200 m by 200 m wide, with a minimum thickness of 85
m immediately beneath the surface, and maximum depth of 7.9 km. The FD and
inversion domain dimensions were chosen such that boundary conditions are respected
for the lowest frequencies. For inversions from set #2, the FD mesh consisted of 89 x 77
x 43 nodes and inversion domain of 46 x 40 x 27 parameters. Inversions from set #2 do
not include the lowermost frequencies used in set #1; hence, the FD and inversion
domains do not need to extend as deep in order to meet the boundary conditions. At 1.6
Hz, assuming a resistivity of 40 m, the penetration depth is approximately 2.5 km. The
FD mesh was reduced to 43 nodes in the vertical direction, extending to a maximum
depth of 20km. The bottom of the inversion domain is reduced to 6 km depth. The FD
82


mesh and inversion parameters are otherwise identical to that used in the inversions from
set #1.
All inversions used a 20 m halfspace as the starting model, near the average of
all observed apparent resistivities. Rigorous Frchet derivatives as described in Chapter 5
were calculated for the starting halfspace and two subsequent models, with Broydens
approximation thereafter. A maximum of 9 model updates was specified. The inversion
was terminated if the misfit increased and the model from the previous iteration was
retained. Regularization parameters were held constant except for the parameter
controlling the smoothness stabilizer. The regularization parameter applied to the a-priori
stabilizer was fixed at 1 for all model updates; the parameter applied to the smoothness
stabilizer was fixed at 2 for the first model update and allowed to vary thereafter. These
values were chosen by trial and error. For subsequent model updates, the program carried
out a line search by testing three values for the smoothness parameter and selecting a
value estimated to yield the lowest misfit (Sasaki, 2001; Lee et al., 2003, Sasaki and
Meju, 2006). As this is not the rigorous approach to regularization as described in
Zhdanov (2009), and in view of the use of approximate derivatives, there is no guarantee
that the parametric functional, normalized misfit, and regularization parameter will
decrease monotonically. This approach to choosing the stabilizer value is similar to the
Occam-2 algorithm (deGroot-Hedlin and Constable, 1990). Computational times required
to complete the first four model updates are provided in Table 7.1.



83


Table 7.1. Approximate time required to complete the first four model updates for each
inversion.
Computer Inversion ApproximateTime
Intel Xeon 5355, 2.66 GHz,
double quad-core PC, 16 GiB
of RAM.
YX mode, 0.63-100 Hz 56 hours 52 min
YX mode, 1.6-100 Hz 41 hours 20 min
J oint +shifts, 1.6-100 Hz 45 hours 55 min
Intel Xeon 5355, 2.5 GHz,
double quad-core PC, 32 GiB
of RAM
XY mode, 0.63-100 Hz 63 hours 43 min
XY mode, 1.6-100 Hz 47hours 20min
J oint, 0.63-100 Hz 64 hours 31 min
J oint, 1.6-100 Hz 42 hours 58 min

7.3.2 Inversion set #1: 100 Hz - 0.63 Hz
The first set of inversions utilized data from 100 Hz to 0.63 Hz. XY and YX mode
apparent resistivity and phase data were inverted independently and jointly. The
convergence of the normalized misfit (nRMS) for the starting and subsequent updated
models is shown in Figure 7.4. The final models presented were obtained after four
model updates; misfits at further iterations were found to increase. Final misfits were
1.94, 2.25, and 4.03 for the YX, XY and joint inversions, respectively.
Misfit is exemplified by plotting, for stations E34 and N32, observed and
predicted data from the models recovered by inversion of YX mode data (Figure 7.5), XY
mode data (Figure 7.6), and by joint inversion (Figure 7.7). For the YX-mode inversion, a
rapid increase misfit at lower frequencies suggests that structures at the bottom of the
inversion domain that marginally influence the data are not resolved as well as the mid-
level structures. The misfit observed at 100 Hz could be due to inadequate representation
of small near-surface variations. Although XY mode data were not included in the
inversion, the predicted XY phase data fit reasonably well for frequencies above 40 Hz,
with rapidly increasing misfits at lower frequencies.
84




Figure 7.4 Convergence of the nRMS value at each iteration for separate YX and XY and
for joint inversions of data from 0.63 to 100 Hz.
85







Figure 7.5 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by inverting YX mode data from
100 Hz to 0.63 Hz. XY mode data and the YX mode phase data at 0.63 Hz were not
included in the inversion.
E
9-
'"
'"
10'
c:: 10'
a.
a.

....----- ,

I
, ,
,

10' 10' 10'
Frequency (Hz)


10' 10' 10'
Frequency (Hz)
80
COSO-E34
Observed data, XY Mode
I Observed data, YX Mode
- Predicted response, XY Mode
- Predicted response, YX Mode

10' 10' 10'
80
Frequency (Hz)
COSO-N32
Observed data, XY Mode
Observed data, YX Mode
Predicted response, XY Mode
- Predicted response, YX Mode

10' 10' 10'
Frequency (Hz)
86






Figure 7.6 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by inverting XY mode data from
100 Hz to 0.63 Hz. YX mode data and the XY mode phase data at 0.63 Hz were not
included in the inversion.
10'
,
,
/'
1 0 ~ ~ ~ ~ - - - - - - ~ ~ ~ ~ - - ~ - - ~ ~
1 ~ 10' 1 ~
Frequency (Hz)
1 0 0 ~ - - ~ - - ~ - - - - ~ - - - - - - - - - - - - ~
1 ~ 10' 1 ~
Frequency (Hz)
80
3l 40
'"
.<::
D-
20
COSO-E34
Observed data, XY Mode
Observed data, YX Mode
- Predicted response, XY Mode
- Predicted response, YX Mode
'I
O ~ - - - - - - - - - - ~ - - - - - - - - - - ~ ~
1 ~ 10' 1 ~
80
Frequency (Hz)
COSO-N32
Observed data, XY Mode
Observed data, YX Mode
- Predicted response, XY Mode
- Predicted response, YX Mode
O ~ - - - - - - - - - - ~ - - - - - - - - - - ~ ~
10' 10' 10
0
Frequency (Hz)
87







Figure 7.7 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by joint inversion of XY and YX
mode data from 100 Hz to 0.63 Hz. Phase data at 0.63 Hz were not included in the
inversion.
E
9-
V>
'"


Frequency (Hz)
0: 10'
a.
a.

80
COSO-E34
Observed data, XY Mode
Observed data, YX Mode
- Predicted response, XY Mode
- Predicted response, YX Mode
0,60
8

5: I : I I
20
80
COSO-N32
Observed data, XY Mode
Observed data, YX Mode
- Predicted response, XY Mode
- Predicted response, YX Mode

10' 10' 10
Frequency (Hz)
88


A roughly cylindrical, high-conductivity (<10 m) zone extending to
approximately 3 km depth is observed in the three models, centered slightly west of the
East Flank wells (Figures 7.8, 7.9, 7.10). A highly resistive zone separates the East Flank
conductor from the conductive sediments beneath Coso Wash. The East Flank conductor
is markedly wider in the XY inversion than the YX and joint inversion models. An
additional high-conductivity (<10 m) zone is observed in the southwest corner of the
survey, separated from the East Flank conductor by the highly resistive zone.
The inversions recover different structures at depths larger than 3 km. The YX
inversion recovers a broad near-horizontal conductor of less than 5 m at a depth of 3
km, extending north-south beneath the Main Flank, and east towards the East Flank
(Figure 7.8). This lateral conductor is not imaged by the XY inversion (Figure 7.9). In the
joint inversion, the lateral conductor appears, but is slightly deeper and smaller than in
the YX inversion (Figure 7.10). This discrepancy between YX and XY models may
reflect the lower frequency anisotropy.
All three inversions recover models which exhibit similar structures in the top
kilometer. Near-surface, three zones are identifiable, where conductivity is slightly
elevated (4 - 10 m), within a background of approximately 70 m (Figures 7.8 7.9,
7.10). These zones roughly coincide with the locations of Coso Hot Springs, Devil's
Kitchen, and Wheeler's Prospect.




89





Figure 7.8 Stacked horizontal slices through the model obtained by inverting YX mode
data from 0.63 to 100 Hz. MT stations are shown at 0.0 km. The colorbar is capped at
1000 m. Slices are at 0.04, 0.33, 0.7, 1.2, 1.7, 2.2, 2.7,and 3.2 km depths.
90





Figure 7.9 Stacked horizontal slices through the model obtained by inverting XY mode
data from 0.63 to 100 Hz. MT stations are shown at 0.0 km. The colorbar is capped at
1000 m.
91






Figure 7.10 Stacked horizontal slices through the model obtained by joint inversion of
XY and YX mode data from 0.63 to 100 Hz. The colorbar is capped at 1000 m.
92


7.3.3 Inversion set #2: 100 Hz - 1.6Hz
In an effort to reduce the possible influence of attempting to fit the deeper
anisotropy on the recovered overlying conductivity structure, inversions are repeated
using a further reduced data set, from 100 Hz to 1.6 Hz. Phase data at 1.6 Hz are not
included in the inversion. While deeper structures will not be resolved, structures within
upper 3 km may be better resolved than is possible when including lower frequency data.
XY and YX mode data are inverted independently and jointly, as well as jointly with
static shifts included as inversion parameters.
The convergence of the normalized misfit (nRMS) for the starting and subsequent
updated models is shown in Figure 7.11. Misfits are illustrated as for the previous
inversions by comparing, for stations E34 and N32, observed and predicted data for the
model recovered from inversion of YX data (Figure 7.12 ), YX data (Figure 7.13 ), XY
and YX data jointly (Figure 7.14), and jointly with static shifts included as inversion
parameters (Figure 7. 15). Misfit is reduced to 1.59 for the YX inversion, 1.48 for the XY
inversion, 2.62 for the joint inversion, and 2.57 when relatively small static shifts are
included in the joint inversion. These are reduced from the misfits obtained when
inverting to lower frequencies although the joint misfit still is somewhat elevated.
Examining the plots of observed and predicted data for stations E34 and N32,
similar trends are observed as when inverting the larger frequency set to 0.63 Hz (Figures
7.5, 7.6, 7.7). The largest misfits are observed for the phase data at the lowest
frequencies. Observed data at 100 Hz are generally less well fit by the joint inversion
results than by single mode inversion results. J oint inversion with static shifts as
inversion parameters was carried out, yielding a small reduction in misfit. Thus, small-
93





Figure 7.11 Convergence of the nRMS value at each iteration for YX, XY and joint
inversions, and for joint inversion with static shift included as inversion parameters, for
data from 1.6 to 100 Hz.
18
16
14
8
6
4
2
18
16
14
(f) 12
::;
10
8
6
4
2
YX Mode Inversion
23456
Model
-0- XY Mode (not inverted)
--0-- YX mode
Joint Inversion
2 3 4 5 6
Model
-o- XY Mode
----0-- YX Mode
----- Joint
18
16
14
12
(f)
::;
'E 10
8
6
4
2
20
18
16
14
12
(f)
::;
0:: 10
<=
8
6
4
2
XY Mode Inversion
23456
Model
-o- XY Mode
-0- YX Mode (not inverted)
Joint Inversion with Shifts

23456
Model
-o- XY Mode
-0- YX Mode
---Joint
94







Figure 7.12 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by inverting YX mode data from
100 Hz to 1.6 Hz. XY mode data and the YX mode phase data at 1.6 Hz were not
included in the inversion
95







Figure 7.13 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by inverting XY mode data from
100 Hz to 1.6 Hz. YX mode data and the XY mode phase data at 1.6 Hz were not
included in the inversion.
96







Figure 7.14 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by joint inversion of XY and YX
mode data from 100 Hz to 1.6 Hz. Phase data at 1.6 Hz were not included in the
inversion.
97






Figure 7.15 Observed and predicted apparent resistivity and phase data for stations E34
(top) and N32 (bottom), with predicted data obtained by joint inversion of XY and YX
mode data from 100 Hz to 1.6 Hz., with shifts included as inversion parameters. Phase
data at 1.6 Hz were not included in the inversion.
98


scale shallow complexity is not considered the primary cause of the moderate high-
frequency misfit.
Models are displayed as stacked horizontal slices for the YX inversion (Figure
7.16), XY inversion (Figure 7.17), joint inversion (Figures 7.18), and joint inversion with
shifts (Figures 7.19). Models from YX and joint inversions are generally similar within
the upper 3 km to those recovered when inverting the larger data set to 0.63 Hz (Figures
7.8, 7.10). There are some variations in the geometry of the East Flank conductor, and the
appearance of a subtle conductive anomaly, at a depth of approximately 1.5 km in the
northern part of the survey area. The subtle anomaly in the northern part of the survey
area is recovered by joint inversion, as is the East Flank conductor. Most striking in the
XY inversion of the reduced data set is the absence of the East Flank conductor beneath a
depth of 1.5 km (Figure 7.17), which apparently is not needed to fit these data. A broad,
subhorizontal conductor is imaged in all three inversions, most conductive and shallowest
in the YX inversion and least conductive and deepest in the XY inversion. This lateral
conductor could be a smoothed out version of the one previously imaged beneath the
west side of the survey. The larger lateral extent may also be a consequence of the
absence of data sensitive to these depths together with model smoothing. The model
obtained by joint inversion with static shifts is largely indistinguishable from that
obtained without (Figures 7.18, 7.19). This data set does not appear to have large static
splits or shifts.



99



Figure 7.16 Stacked horizontal slices through the model obtained by inverting YX mode
data from 1.6 Hz to 100 Hz. The colorbar is capped at 1000 m. Slices are at 0.04, 0.33,
0.7, 1.2, 1.7, 2.2, 2.7, and 3.2 km depths.
100






Figure 7.17 Stacked horizontal slices through the model obtained by inverting XY mode
data from 1.6 to 100 Hz. The colorbar is capped at 1000 m.
Easting (km)
430
428
3990 Northing(km)
3988
3986
o
2.5
3
3
101





Figure 7.18 Stacked horizontal slices through the model obtained by joint inversion of
XY and YX mode data from 1.6 to 100 Hz. The colorbar is capped at 1000 m.
Easting (km)
428
3990 Northing(km)
. - . 3988
3986
o
0.5
1.5
2
0
E
S.

2.5
';;;

'iii
Ql
2
0::
0
Ol
0
3
...J
3
102






Figure 7.19 Stacked horizontal slices through the model obtained by joint inversion of
XY and YX mode data from 1.6 to 100 Hz with shifts included as inversion parameters.
The colorbar is capped at 1000 m.
Easting (km) 432
o
0.5
1.5
2
2.5
3
E
""
.<:
a.
Q)
o
0
E
9-
Z.
;;;
~
iii
Q)
2
0::
0
Cl
0
...J
3
103


7.3.4 Discussion
In order to interpret the inversion results, plan and cross-section slices are
extracted from the model obtained by joint inversion using data from 100 to 1.6 Hz.
These are overlain with the approximate locations of fault traces selected from the
geological map by Whitmarsh (2002a, 2002b), and from Davatzes and Hickman (2006),
and with the reservoir locations as shown in Figure 7.1. Producing and injection well
intervals provided by the U.S. Navy (F. Monastero, pers. Comm.) from within a swath
approximately 600 m wide have been included on vertical cross-sections. Temperature
contours by Adams et al. (2000), obtained from fluid inclusion analysis, have been
reproduced on cross-sections through the Main Field. In addition to these data, Bruce
J ulian of the USGS has provided earthquake locations recorded from 1997 to 2008 using
the Navy Permanant Array at Coso Field (J ulian et al., 2010). These have been
reproduced on plan and cross-section views.
The near-surface conductivity structure displays enhanced conductivity zones
roughly coincident with Coso Hot Springs, Devil's Kitchen, and Wheeler's Prospect,
areas with surface hydrothermal deposits (Figure 7.20). These zones are attributed to the
presence of hydrothermal fluids and resultant low-temperature alteration. An elongate
zone of mildly elevated conductivities between Devil's Kitchen and Coso Hot Springs
appears to follow fault traces connecting them, suggesting near-surface preferential flow
along them at some point. There is little discernible resistivity expression to the
southwest (SW) portion of the field consistent with lack of shallow thermal
manifestations there (Lutz et al., 1999; Adams et al., 2000) even though it has the highest
temperatures at depth.
104





Figure 7.20 Slice through middle of topmost inversion layer, overlain with location of
reservoirs and fault traces as described in Fig. 7.1.
8 '
i p 4


3991



3990









C'
Fig, 7,25


3986 A





3985

428 429 432 433 434
Map Symbols
OE
_______ Boundary of Quaternary rhyolite (Shown in white)
g.
>-
__ , __ Fault contact. Bail on down-dropped side
1'"5
.. ,"
(Shown in white)
2' w
Q)
Modern faults (offset Holocene sediments)
0::
3
0
+MT Station r:
0
o Approximate outli ne of reservoir
-'
DK: Devil 's Kitchen
CHS: Coso Hot Springs
WP: Wheeler Prospect
105


The wedge conductor imaged on the east side of the survey, extending to
approximately 1.5 km depth, is interpreted to reflect sediments of the Coso Wash graben
(Figure 7.21, 7.22). In the north, the boundary between the wedge and adjacent resistor
seen in plan view clearly mimics the Coso Wash Fault, and appears dipping towards the
east.
A plan view, with a nominal depth of 1.9 km, shows the East Flank conductor and
the Main Field southwest conductor (Figure 7.23). These conductors broadly coincide
with the locations of the interpreted reservoirs. The East Flank conductor appears slightly
offset to the west of the interpreted reservoir. On the east side of the survey, a remnant of
Coso Wash can be seen. Other anomalies, less conductive and more restricted in size, are
present at this depth, located north and south of the East Flank Reservoir. The southern
anomaly is poorly imaged due to lack of data coverage in this area.
A south-north cross-section through the Main Field passes north through the
southwest reservoir and Devil's Kitchen (Figure 7.24). The southwest reservoir, with
resistivity less than 5 m, coincides with a 300C temperature contour and projected
production intervals. This area has been described as an upflow zone, with fluids
traveling upwards and northwards. A zone of intermediate resistivities on the order of
tens of m, also containing producing intervals, extends from the reservoir northwards.
The broad, approximately 10 m, near-surface conductor overlying the intermediate
resistivity zone is consistent with low-temperature alteration. Its absence near-surface
above the southwest reservoir is consistent with the lack of surface hydrothermal activity
noted previously. This cross-section is on the west edge of the survey area, so finer-scale
features are not well-constrained.
106






Figure 7.21 Slice through the middle of the inversion layer between 0.5 and 0.65 km
overlain with location of reservoirs and fault traces as described in Fig. 7.1, and
earthquake hypocenters occurring at depths approximately from 0.4 km to 0.7 km.
107











Figure 7.22 East-oriented cross-section along northing 3985.9 km.
108





Figure 7.23 Plan view at 1.9 km depth, showing the Main Field and East Flank
conductivity anomalies.
109







Figure 7.24 North-south vertical slice through the Devil's Kitchen area, at an Easting of
428 km. Temperature contours are reproduced from Adams et al. (2000).
110


A west-east cross-section is constructed through the East Flank reservoir (Figure
7.25). Production intervals appear to coincide with the edge of the conductive anomaly,
believed conductive due to the presence of fluids and alteration. This East Flank
conductor is separated from the Coso Wash wedge and the Main Field by highly resistive
zones, believed to represent low permeability, plutonic host rock. Both the East Flank and
the southwest Main Field reservoirs are more conductive than the clay cap.
Earthquake epicenters appear concentrated near the margins of the conductive
anomalies. Much of this seismicity, especially among the smaller events, may be
production-related. However, similar behavior of earthquake epicenters concentrated at
the margins of conductors has been observed by Mitsuhata et al. (2001), and Ogawa and
Honkura (2004), among others. Conductive zones are attributed to the rock containing
pressurized pore fluid, less likely to deform seismically. These zones may supply fluids
to neighboring stressed (more resistive) rock promoting brittle failure (Ogawa and
Honkura, 2004).
An approximate fluid porosity can be estimated for the interpreted reservoir zones
assuming the fluids lie predominantly in interconnected fractures. From Grant and West
(1965), the formula for bulk conductivity
b
(inverse of resistivity
b
) for a fracture
network is
b
=1/
b
=2
f
X/3, where X is porosity (0 to 1 range) and
f
=1/
f
is fluid
conductivity. The deeper brines of the Southwest Field approach 3 wt % NaCl (Adams et
al., 2000). Following Nesbitt (1993), such brines for T >200 C will have resistivities of
order 0.05 m. Thus, fracture porosity of 3-5% by volume are consistent with bulk
resistivities near 2 m. Shallower production fluids have salinities more in the 1-2 wt %
111









Figure 7.25 East-west vertical slice through the East Flank Reservoir, along Northing
3987.2 km.
112


range. Bulk resistivities are higher and temperatures are lower there, however, which
have complementary effects on porosity estimates.
A 3.5- 4 km deep broad-scale conductive zone is present and interpreted as fluid
related, but poorly constrained due to the limited depth penetration of and higher misfits
observed for the lowest frequencies included in the inversion. This zone could lie just
below the brittle-ductile rheological transition and, if so, likely reflects fluids
interconnected along grain edge boundaries (e.g., J iracek et al., 1995).
To further refine the geometry of the East Flank and Main Field reservoirs, one
additional inversion was carried out. This inversion used data from only 39 receivers,
concentrated in the south (Figure 7.26). The minimum horizontal spacing between finite
difference nodes was reduced to 60 m, and 77 x 81 x 43 nodes were included in the mesh.
The minimum horizontal dimension for inversion parameters was reduced to 120 m, and
42 x 44 x 27 parameters were included in the inversion. Vertically, inversion bricks
increase in size from 60 m at the surface to 500 m at the base of the inversion domain,
approximately 5 km deep. This joint inversion used data from 100 to 1.6 Hz, neglecting
the phase at 1.6 Hz, and recovered a model with an nRMS misfit of 2.4. These results are
more detailed but otherwise similar to those previously obtained and verify that the main
features are not sensitive to inversion discretization.
East Flank and southwest Main Field reservoirs are shown in Figure 7.26, at a
depth of approximately 1.1 km. A west-dipping fault may form the eastern margin of the
East Flank reservoir, similar to and possibly connected with the west-dipping fault of
J ulian et al. (2010) (Figure 7.27). The damage zone of the fault controlling the field may
extend more than a few hundred meters from the main slip plane and may exhibit
113










Figure 7.26 Plan view at 1.1 km depth through finely-discretized inversion.
114









Figure 7.27 East-west vertical slice through the East Flank Reservoir, along Northing
3987.1 km, for the finely discretized inversion.
115


substantial variation along different sections (Cochran et al., 2009). The width of the East
Flank conductor may reflect a broad, persistent damage zone in the hanging wall. The
conductor appears to extend to 3 km, consistent with well observations (Kovac, 2005).
A north-south cross-section through the Main Field is shown in Figure 7.28. In
addition to detecting the southwest reservoir and the shallow high-permeability zone, a
smaller conductive anomaly is imaged in the vicinity of Devil's Kitchen.

7.4 Conclusions
A 3-D resistivity inversion model of the Coso geothermal system was produced
using a workstation-based algorithm whose information content is similar to that from
massively parallel inversion platforms. Previously published interpretations of the Coso
data have included 2-D stitched vertical slices (Wannamaker, 2004; Newman et al.,
2005a), and 3-D inversion (Newman et al., 2008). The 3-D inversion was carried out
using a massively parallel computer, using a finely discretized model seeded with an
initial guess incorporating 2-D inversion results. An important structure appearing in
these interpretations is a steeply-dipping conductor most prominent in the southwest East
Flank sector interpreted to reflect fluid-filled fractures.
The inversion results I report were obtained using a desktop multicore computer,
starting from a halfspace, and obtained good fits to the data (cf. Newman et al., 2008). A
nearly vertical zone of anomalously high conductivity is imaged in a similar location as
observed by Newman et al. (2008), west of and overlapping with the East Flank
production zone. Beneath the Devils Kitchen area of the Main Field, a shallow north-
trending zone of intermediate conductivity was imaged, coincident with production well
116









Figure 7.28 North-south vertical slice through the Devil's Kitchen area, at an Easting of
428 km, for the finely discretized inversion.
117


intervals. The results differ from those of Newman et al. by detecting a deep,
anomalously conductive zone in the southwest Main Flank, also coincident with
production well intervals.




CHAPTER 8

8SUMMARY

The program developed by Sasaki (2004) for 3-D inversion of MT data on single-
core desktop PC has been modified to run on the newer multicore computers. This has
been accomplished by restructuring and incorporating OpenMP statements into the
Fortran program, to enable parallelization of several key sections. The key sections
parallelized include the calculation of the Frchet derivative matrix and factoring of the
approximate regularized Hessian required to solve for the updated model parameters.
Coarse-grained parallelization of the forward modeling and Frchet derivative
matrix was implemented by distributing these calculations for each frequency to separate
threads. Threads are processed concurrently and asynchronously. While speed-up is
improved, it appears to be limited by memory-intensive scattered field solutions.
Factoring the Hessian was accelerated by implementing the LDL
T
method using a block-
cyclic algorithm. Doubling the number of processors used to factor the Hessian
approximately halves the time required. Carrying out these modifications within the
context of a desktop multicore computer results in a program that does not require
specialized resources such as a cluster, potentially broadening the user base. Trends in
capability of multicore platforms are encouraging for this type of parallelization.
119


The modified program has been used to invert MT data from 102 tensor stations
over the East Flank of the Coso geothermal field. Separate and joint inversions of XY and
YX mode data were carried out with frequencies from 100 Hz to 0.63 Hz, and repeated
for 100 Hz to 1.6 Hz. The largest finite difference modeling domain used contained
approximately 315,000 nodes; the largest inversion domain contained approximately
55,000 model parameters.
Inversions were done using various Linux-based 2.5 and 2.66 GHz Intel quad-
core computers with 8 processors, and either 16 or 32 GB of RAM. Inversions generally
required between 2 and 2 days to reach an acceptable misfit. Prior to parallelization,
similar inversions with approximately 134,000 finite difference nodes and 20,000 model
parameters required 2 to 3 days using a 3.4 GHz desktop PC. In addition to reducing the
time required to invert a dataset, the modifications have enabled the use of more finite
difference nodes and model parameters.
Several prominent features evident in the models include conductive anomalies
attributed to the East Flank and Main Field geothermal reservoirs, and to the Coso Wash
sediments. Highly conductive reservoirs of less than 10 m are imaged within a resistive
host, consistent with observations from other geothermal reservoirs developed in fracture
zones within impermeable rocks. The resistivity structure correlates well with seismicity
and production intervals, while resistivity values can be explained fracture porosity of 5%
containing fluids near 300C and 3.5 wt % salinity. The recovered conductivity structure
is qualitatively similar to that recovered by Newman et al. (2005b, 2008) using a
massively parallel algorithm on a computing cluster.


REFERENCES
Adams, M. C., J . N. Moore, S. Bjornstad, and D. I. Norman, 2000, Geologic history of
the Coso geothermal system: Geothermal Resources Council Transactions, 24,
205-209.
Agarwal, A. K., and J . T. Weaver, 2000, Magnetic distortion of the magnetotelluric
tensor: a numerical study: Earth, Planets and Space, 52, 347-353.
Avdeev, D. B., 2005, Three-dimensional electromagnetic modelling and inversion from
theory to application: Surveys in Geophysics, 26, 767-799.
Baboulin, M., L. Giraud, and S. Gratton, 2005, A parallel distributed solver for large
dense symmetric systems: applications to geodesy and electromagnetism
problems: International J ournal of High Performance Computing Applications,
19, 353-363.
Bahr, K., 1988, Interpretation of the magnetotelluric impedance tensor: regional
induction and local telluric distortion: J ournal of Geophysics, 62, 119-127.
Barker, K. J ., K. Davis, A. Hoisie, D. J . Kerbyson, M. Lang, S. Pakin, and J . C. Sancho,
2008, A performance evaluation of the Nehalem quad-core processor for
scientific computing: Parallel Processing Letters, 18, 453-469.
Bartley, J . M., D. S. Coleman, A. F. Glazner, and J . D. Walker, 2004, Late Cenozoic
deformation in the Coso Range, eastern California, and relation to the Coso
magmatic and geothermal systems: Geothermal Resource Council, Transactions,
28, 633-636.
Berktold, A., 1983, Electromagnetic studies in geothermal regions: Geophysical Surveys,
6, 173-200.
Broyden, C. G., 2000, On the discovery of the "good Broyden" method: Mathematical
Programming, 87, 209-213.
Buttari, A., J . Dongarra, J . Kurzak, J . Langou, P. Luszczek, and S. Tomov, 2007a, The
impact of multicore on math software, in B. Kagstrom, E. Elmroth, J . Dongarra,
and J . Wasniewski, eds, PARA 2006: Lecture Notes in Computer Science, 4699,
1-10.
Buttari, A., J . Langou, J . Kurzak, and J . Dongarra, 2007b, A class of parallel tiled linear
algebra algorithms for multicore architectures: Lapack working note #191, 19 pp.
121


Cagniard, L., 1953, Basic theory of the magneto-telluric methods of geophysical
prospecting: Geophysics, 18, 605-635.
Caldwell, T. G., H. M. Bibby, and C. Brown, 2004, The magnetotelluric phase tensor:
Geophysical J ournal International, 158, 457-469.
Cochran, E.S., Y-G. Li, P.M. Shearer, S. Barbot, Y. Fialko, and H.E. Vidale, 2009,
Seismic and geodetic evidence for extensive, long-lived fault damage zones:
Geology, 37, 315-318.
Cox, S. F., 1999, Deformational controls on the dynamics of fluid flow in mesothermal
gold systems, in K. J . W. McCaffrey, L. Lonergan, and J . J . Wilkinson, eds.,
Fractures, Fluid Flow and Mineralization: Geological Society Special
Publications, 155, 123 140.
Davatzes, N. E., and S. H. Hickman, 2006, Stress and faulting in the Coso geothermal
field: update and results from the East Flank and Coso Wash: Proceedings of the
Thirty-First Workshop on Geothermal Reservoir Engineering Stanford University,
12 pp.
deGroot-Hedlin, C. D., 1991, Removal of static shift in two dimensions by regularized
inversion: Geophysics, 56, 21022106.
deGroot-Hedlin, C. D., 1995, Inversion for regional 2-D resistivity structure in the
presence of galvanic scatterers: Geophysical J ournal International, 122, 877-888.
deGroot-Hedlin, C. D., and S. C. Constable, 1990, Occam's inversion to generate smooth,
two-dimensional models from magnetotelluric data: Geophysics, 55, 1613-1624.
deLugao, P. P., and P. E. Wannamaker, 1996, Calculating the two-dimensional
magnetotelluric J acobian in finite elements using reciprocity: Geophysical J ournal
International, 127, 806-810.
Duffield, W. A., C. R. Bacon, and G. B. Dalrymple, 1980, Late Cenozoic volcanism,
geochronology, and structure of the Coso Range, Inyo County, California: J ournal
of Geophysical Research, 85, 23812404.
Duffield, W. A., and J . H. Sass, 2003, Geothermal energy - clean power from the Earth's
heat: USGS Circular 1249, 43 pp.
Frankel, K. L., A. F. Glazner, E. Kirby, F. C. Monastero, M. D. Strane, M. E. Oskin, , J .
R. Unruh, J . D. Walker, S. Anandakrishnan, J . M. Bartley, D. S. Coleman, J . F.
Dolan, R. C. Finkel, D. Greene, A. Kylander-Clark, S. Marrero, L. A. Owen, and
F. Phillips, 2008, Active tectonics of the eastern California shear zone, in E. M.
Duebendorfer, and E. I. Smith, eds., Field guide to plutons, volcanoes, faults,
reefs, dinosaurs, and possible glaciation in selected areas of Arizona, California,
and Nevada: Geological Society of America Field Guide 11, 43-81.
122


Gamble, T., W. Goubau, and J . Clarke, 1979, Magnetotellurics with a remote reference,
Geophysics, 44, 53-68.
Golubev, N., M. Zhdanov, and K. Matsuo,, 1999, Three-dimensional inversion of
magnetotelluric data collected for hydrocarbon exploration in the overthrust area
in J apan: SEG Annual Meeting Expanded Technical Program Abstracts with
Biographies, 69, 275-278.
Golubev, N. G., and M. S. Zhdanov, 2000, Three-dimensional interpretation of
magnetotelluric data collected for mineral exploration in Voisey's Bay, Labrador,
Canada: Proceedings of the CEMI Annual Meeting, 559-594.
Grant, F. S., and G. F. West, 1965, Interpretation Theory in Applied Geophysics:
McGraw-Hill, New York, 584 pp.
Green, A., A. Gribenko, M. Cuma, and M. S. Zhdanov, 2008, Preliminary
results of 3D inversion of the EarthScope Oregon MT data using the integral
equation method: EOS, Transactions, American Geophysical Union, GP51C-08.
Groom, R. W., and R. C. Bailey, 1989, Decomposition of magnetotelluric impedance
tensors in the presence of local three-dimensional galvanic distortion: J ournal of
Geophysical Research, 94, 1913-1925.
Hargrove, W. W., F. M. Hoffman, and T. Sterling, 2001, The do-it-yourself
supercomputer: Scientific American, 256, 72-79.
Harrington, R. F., 1961, Time-harmonic Electromagnetic Fields: McGraw-Hill Electrical
and Electronic Engineering Series, McGraw-Hill Book Co., New York, 480 pp.
Hauksson, E., and J . Unruh, 2007, Regional tectonics of the Coso geothermal area along
the intracontinental plate boundary in central eastern California: three-
dimensional Vp and Vp/Vs models, spatial-temporal seismicity patterns, and
seismogenic deformation: J ournal of Geophysical Research, 112, 24 pp.
Hohmann, G. W., 1975, Three-dimensional induced polarization and electromagnetic
modeling: Geophysics, 40, 309-324.
J iracek, G. R., 1990, Near-surface and topographic distortions in electromagnetic
induction: Surveys in Geophysics, 11, 163-203.
J iracek, G. R., V. Haak, and K. H. Olsen, 1995, Practical magnetotellurics in a
continental rift environment: Developments in Geotectonics, 25,103-129.
J ones, A. G., 1988, Static shift of magnetotelluric data and its removal in a sedimentary
basin environment: Geophysics, 53, 967-978.
123


J ulian, B.R., G.R. Foulger, F.C. Monastero, and Steven Bjornstad, 2010, Imaging
hydraulic fractures in a geothermal reservoir: Geophysical Research Letters, 37, 5
pp.
Keller, G. V., 1988, Rock and mineral properties, in M. N. Nabighian, ed.,
Electromagnetic Methods In Applied Geophysics: Society of Exploration
Geophysics, 13-51.
Kovac, K. M., J . N. Moore, and S. J . Lutz, 2005, Geological framework of the East
Flank, Coso geothermal field: implications for EGS development: Proceedings of
the 30th Workshop on Geothermal Reservoir Engineering, Stanford, CA, 7 pp.
Lee, T. J ., T. Uchida, Y. Sasaki, and Y. Song, 2003, Characteristics of static shift in 3-D
MT inversion, Mulli-Tamsa (Geophysical Exploration), 6, 199-206.
Lees, J . M., 2002, Three-dimensional anatomy of a geothermal field, Coso, southeast-
central California, in A. F. Glazner, J . D. Walker, and J . M. Bartley, eds.,
Geologic Evolution of the Mojave Desert and Southwestern Basin and Range:
Geological Society of America Memoir 195, 259276.
Lutz, S. J ., J . N. Moore, and J . F. Copp, 1996, Integrated mineralogical and fluid
inclusion study of the Coso geothermal system, California: Proceedings of the
Twenty-first Workshop on Geothermal Reservoir Engineering, Stanford
University, 187-194.
Lutz, S. J ., J . N. Moore, M. C. Adams, and D. I. Norman, 1999, Tracing fluid sources in
the Coso geothermal system using fluid-inclusion gas chemistry: Twenty-Fourth
Workshop on Geothermal Reservoir Engineering, Stanford University, 188-195.
Mackie, R. L., J . T. Smith, and T. R. Madden, 1994, Three-dimensional electromagnetic
modeling using finite difference equations: the magnetotelluric example: Radio
Science, 29, 923-935.
Mackie, R.L., and M.D. Watts, 2004, The use of 3D magnetotelluric inversion for
exploration in complex geologic environments; potential pitfalls and real world
examples: Eos, Transactions, American Geophysical Union, 85, GP14A-01.
Madden, T. R., and R. L. Mackie, 1989, Three-dimensional magnetotelluric modeling
and inversion: Proc. IEEE, 77, 318-333.
Madden, T., and P. Nelson, 1964, A defense of Cagniard's magnetotelluric method: ONR
Rept, NR-371-401, Geophysics Lab, MIT
Manley, C. R., and C. R. Bacon, 2000, Rhyolite thermobarometry and the shallowing of
the magma reservoir, Coso Volcanic Field, California: J ournal of Petrology 41,
149174.
124


Martinez-Garcia, M., 1992, Electromagnetic induction in geothermal fields and volcanic
belts: Surveys in Geophysics, 13, 409-434.
McClusky, S. C., S. C. Bjornstad, B. H. Hager, R. W. King, B. J . Meade, M. M. Miller,
F. C. Monastero, and B. J . Souter, 2001, Present day kinematics of the Eastern
California Shear Zone from a geodetically constrained block model: Geophysical
Research Letters, 28, 3369-3372.
McGillivray, P. R., and D. W. Oldenburg, 1990, Methods for calculating Frchet
derivatives and sensitivities for the nonlinear inverse problem: a comparative
study: Geophysical Prospecting, 38, 499-524.
Meju, M. A., 1996, J oint inversion of TEM and distorted MT soundings: some effective
practical considerations: Geophysics, 61, 56-65.
Meju, M. A., 2002, Geoelectromagnetic exploration for natural resources: models, case
studies and challenges: Surveys in Geophysics, 23, 133-205.
Mitsuhita, Y., Y. Ogawa, M. Mishina, T. Kono, T, Yokokura, and T. Uchida, 2001,
Electromagnetic heterogeneity of the seismogenic region of 1962 M6.5 Northern
Miyagi Earthquake, northeastern J apan: Geophysical Research Letters, 28, 4371-
4374.
Monastero, F. C., 2002. Model for success: Geothermal Resources Council Bulletin,
September/October, 188-194.
Monastero, F. C., A. M. Katzenstein, J . S. Miller, J . R. Unruh, M. C. Adams, and K.
Richards-Dinger, 2005, The Coso geothermal field: a nascent metamorphic core
complex: Geological Society of America Bulletin, 117, 1534-1553.
Newman, G., and D. L. Alumbaugh, 1995, Frequency-domain modeling of airborne
electromagnetic responses using staggered finite differences: Geophysical
Prospecting, 43, 1021-1042.
Newman, G.A., and D. L. Alumbaugh, 2000, Three-dimensional magnetotelluric
inversion using non-linear conjugate-gradients: Geophysical J ournal International,
140, 410-424.
Newman, G.A., and D. L. Alumbaugh, 2002, Three-dimensional induction logging
problems, Part 2: a finite-difference solution: Geophysics, 67, 484-491.
Newman, G. A., M. Hoversten, E. Gasperikova, and P. Wannamaker, 2005a, 3D
magnetotelluric characterization of the Coso geothermal field: Proceedings 30th
Workshop on Geothermal Reservoir Engineering, SGP- TR-176.
Newman, G. A., M. Hoversten, and E. Gasperikova, 2005b, 3D Magnetotelluric
investigation of the Coso Geothermal Field: Geothermal Resources Council
Transactions, 29, 493-496.
125


Newman, G. A., E. Gasperikova, M. Hoversten, and P. Wannamaker, 2008, Three-
dimensional magnetotelluric characterization of the Coso geothermal field:
Geothermics, 37, 369-399.
Ogawa, Y., 2002, On two-dimensional modeling of magnetotelluric field data: Surveys
in Geophysics, 23, 251-272.
Ogawa, Y., and Y. Honkura, 2004, Mid-crustal electrical conductors and their
correlations to seismicity and deformation at Itoigawa-Shizuoka Tectonic Line,
Central J apan: Earth Planets Space, 56, 1285-1291.
Ogawa, Y., and T. Uchida, 1996, A two-dimensional magnetotelluric inversion assuming
Gaussian static shift: Geophysical J ournal International, 126, 69-76.
Oskooi, B., L. B. Pedersen, M. Smirnov, K. rnason, H. Eysteinsson, and A. Manzella,
2005, The deep geothermal structure of the Mid-Atlantic Ridge deduced from MT
data in SW Iceland: Physics of the Earth and Planetary Interiors, 150, 183-195.
Parasnis, D. S., 1988, Reciprocity theorems in geoelectric and geoelectromagnetic work:
Geoexploration, 25, 177-198.
Pellerin, L., and G. W. Hohmann, 1990, Remedy for static shift: Geophysics, 56, 951-
960.
Pellerin, L., J . M. J ohnston, and G. W. Hohmann, 1996, A numerical evaluation of
electromagnetic methods in geothermal exploration: Geophysics, 61, 121-130.
Pluhar, C. J ., R. S. Cose, J . C. Lewis, F. C. Monastero, and J . M. G. Glen, 2006, Fault
block kinematics at a releasing stepover of the Eastern California shear zone:
partitioning of rotation style in and around the Coso geothermal area and nascent
metamorphic core complex: Earth and Planetary Science Letters, 250, 134-163.
Rodi, W., and R.L. Mackie, 2001, Nonlinear conjugate gradients algorithm for 2-d
magnetotelluric inversion: Geophysics, 66, 174-187.
Roquemore, G., 1980, Structure, tectonics, and stress field of the Coso Range, Inyo
County, California: J ournal of Geophysical Research, 85, 24342440.
Sasaki, Y., 2001, Full 3-D inversion of electromagnetic data on PC: J ournal of Applied
Geophysics, 46, 45-54.
Sasaki, Y., 2004, Three-dimensional inversion of static-shifted magnetotelluric data:
Earth, Planets and Space, 56, 239-248.
Sasaki, Y., and M. A. Meju, 2006, Three dimensional joint inversion for magnetotelluric
resistivity and static shift distributions in complex media: J ournal of Geophysical
Research, 111, 11 pp.
126


Sheridan, J ., K. Kovac, P. Rose, C. Barton, J . McCulloch, B. Berard, J . Moore, S. Petty,
and P. Spielman, 2003, In situ stress, fracture and fluid flow analysis east flank
of the Coso geothermal system: Proceedings, 28th Workshop on Geothermal
Reservoir Engineering, SGP-TR-173.
Simpson, F., and K. Bahr, 2005, Practical Magnetotellurics: University Press, Cambridge,
270 pp.
Siripunvaraporn, W., M. Uyeshima, and G. Egbert, 2004, Three-dimensional inversion
for network-magnetotelluric data: Earth, Planets and Space, 56, 893-902.
Siripunvaraporn, W., G. Egbert, Y. Lenbury, and M. Uyeshima, 2005, Three-dimensional
magnetotelluric inversion: data-space method: Physics of the Earth and Planetary
Interiors, 150, 3-14.
Smith, J . T., 1996a, Conservative modeling of 3-D electromagnetic fields, Part I:
Properties and error analysis: Geophysics, 61, 13081319.
Smith, J . T., 1996b, Conservative modeling of 3-D electromagnetic fields, Part II:
Biconjugate gradient solution and an accelerator: Geophysics, 61, 13191324.
Spichak, V., and A. Manzella, 2009, Electromagnetic sounding of geothermal zones:
J ournal of Applied Geophysics, 68, 459-478.
Spies, B. R., and F. C. Frischknecht, 1991, Electromagnetic sounding, in M. N.
Nabighian, ed., Electromagnetic Methods in Applied Geophysics: Society of
Exploration Geophysicists, 285386.
Spitzer, K., 2001, Magnetotelluric static shift and direct current sensitivity: Geophysical
J ournal International, 144, 289-299.
Sternberg, B.K., J . C. Washburne, and L. Pellerin, 1988, Correction for the static shift in
magnetotellurics using transient electromagnetic soundings: Geophysics, 53,
1459-1468.
Stratton, J ., 1941, Electromagnetic Theory: McGraw Hill, New York, 615 pp.
Tarantola, A., 1987, Inverse Problem Theory: Elsevier, New York, 613 pp.
Tarantola, A., 2005, Inverse Problem Theory and Methods for Model Parameter
Estimation: SIAM, Philadelphia, 352 pp.
Tikhonov, A. N., 1950, On determining electrical characteristics of the deep layers of the
earth's crust: Deki Akud Nuck, 73, 295-297.
Torres-Verdin, C., and F. X. Bostick, J r., 1992a, Principles of spatial surface electric field
filtering in magnetotellurics: electro-magnetic array profiling (EMAP):
Geophysics, 57, 603-622.
127


Torres-Verdin, C., and F. X. Bostick, J r., 1992b, Implications of the Born approximation
for the magnetotelluric problem in three-dimensional environments: Geophysics,
57, 587-602.
Uchida, T., and Y. Sasaki, 2006, Stable 3D inversion of MT data and its application to
geothermal exploration: Exploration Geophysics, 37, 223-230.
Unruh, J . R, S. K. Pullammanappallil, and W. Honjas, 2001, New seismic imaging of the
Coso geothermal field, eastern California: Proceedings of the 26th workshop on
Geothermal Reservoir Engineering, Stanford University.
Unruh, J . R., E. Hauksson, F. C. Monastero, R. J . Twiss, and J . C. Lewis, 2002,
Seismotectonics of the Coso RangeIndian Wells Valley region, California:
Transtensional deformation along the southeastern margin of the Sierran
microplate, in A. F. Glazner, J . D. Walker, and J . M. Bartley, eds., Geologic
Evolution of the Mojave Desert and Southwestern Basin and Range: Geological
Society of America Memoir 195, 277294.
Volpi, G., A. Manzella, and A. Fiordelisi, 2003, Investigation of geothermal structures by
magnetotellurics (MT): an example from the Mt. Amiata area, Italy:
Geothermics, 32, 131-145.
Vozoff, K., 1972, The magnetotelluric method in the exploration of sedimentary basins:
Geophysics, 37, 98-141.
Vozoff, K., 1991, The magnetotelluric method, in M. N. Nabighian, ed., Electromagnetic
Methods In Applied Geophysics: Society of Exploration Geophysics, 641-711.
Vuduc, R., J . W. Demmel, K. A. Yelick, S. Kamil, R. Nishtala, and B. Lee, 2002,
Performance optimizations and bounds for sparse matrix-vector multiply:
Proceedings of the 2002 ACM/IEEE conference on Supercomputing, 1-35.
Wan, L., M. S. Zhdanov, K. Key, and S. Constable, 2006, Rigorous 3-D inversion of
marine MT data: Proceedings of Annual Meeting of Consortium for
Electromagnetic Modeling and Inversion, 229-254.
Wannamaker, P. E., 1999, Affordable magnetotellurics: interpretation of MT sounding
profiles from natural environments, in M. Oristaglio and B. Spies, eds., Three-
Dimensional Electromagnetics: Development Series no. 7, Society of Exploration
Geophysics, Tulsa, 349-374.
Wannamaker, P. E., 2004, Creation of an enhanced geothermal system through hydraulic
and thermal stimulation magnetotelluric surveying and monitoring: Quarterly
Project Information and Planning (PIP) Report, U.S. Dept. of Geothermal
Technology.
Wannamaker, P. E., G. W. Hohmann, and S. H. Ward, 1984, Magnetotelluric responses
of three-dimensional bodies in layered earths: Geophysics, 49, 1517-1533.
128



Wannamaker, P. E., G. R. J iracek, J . A. Stodt, T. G. Caldwell, A. D. Porter, V. M.
Gonzalez, and J . D. McKnight, 2002, Fluid generation and movement beneath an
active compressional orogen, the New Zealand Southern Alps, inferred from
magnetotelluric (MT) data: J ournal of Geophysical Research, 107, 22pp.
Wannamaker, P. E., P. E. Rose, W. M. Doerner, B. Berard, J . McCulloch, and K. Nurse,
2004, Magnetotelluric surveying and monitoring at the Coso geothermal area,
California, in support of the enhanced geothermal systems concept: survey
parameters and initial results: Proceedings, 29th Workshop on Geothermal
Reservoir Engineering, SGP- TR-175.
Wannamaker, P. E., 2005, Anisotropy versus heterogeneity in continental solid earth
electromagnetic studies: fundamental response characteristics and implications for
physicochemical state, invited review paper: Surveys in Geophysics, 26, 733-765.
Ward, S. H., and G. W. Hohmann, 1988, Electromagnetic theory for geophysical
applications in M. N. Nabighian, ed., Electromagnetic Methods In Applied
Geophysics: Society of Exploration Geophysics, 131-311.
Weidelt, P., 1975, Electromagnetic induction in three-dimensional structures:
Geophysical J ournal Royal Astronomical Society, 41, 85-109
Whitmarsh, R.S., 2002a, Geological map of the Cactus Peak 7.5 quadrangle, Inyo
County, California, CD-ROM map, in Glazner, A.F., Walker, J .D., Bartley, J .M.
eds., Geologic Evolution of the Mojave Desert and Southwestern Basin and
Range: Geological Society of America Memoir 195.
Whitmarsh, R.S., 2002b, Geological map of the Petroglyph Canyon 7.5 quadrangle, Inyo
County, California, CD-ROM map, in Glazner, A.F., Walker, J .D., Bartley, J .M.
eds., Geologic Evolution of the Mojave Desert and Southwestern Basin and
Range: Geological Society of America Memoir 195.
Wilson, C. K., C. H. J ones, and H. J . Gilbert, 2003, Single-chamber silicic magma system
inferred from shear wave discontinuities of the crust and uppermost mantle, Coso
geothermal area, California, J ournal of Geophysical Research, 108, 16 pp.
Wright, P. M., S. H. Ward, H. P. Ross, and R. West, 1985, State-of-the-art geophysical
exploration for geothermal resources: Geophysics, 50, 2666-2699.
Yee, K. S., 1966. Numerical solution of initial boundary problems involving Maxwell's
equations in isotropic media: IEEE Trans. Ant. Prop., 14, 302-307.
Zhdanov, M. S., 2002, Geophysical Inverse Theory and Regularization Problems:
Elsevier, Oxford, 609 pp.
129


Zhdanov, M. S., 2009, Geophysical electromagnetic theory and methods: Methods in
Geochemistry and Geophysics, 43, Elsevier, 868 pp.
Zhdanov, M. S., and N. G. Golubev, 2003, Three-dimensional inversion of
magnetotelluric data in complex geological structures, in J . Macnae, and G. Liu,
eds., Three-dimensional Electromagnetics III: Australian Society of Exploration
Geophysicists.
Zhdanov, M. S., and A. Gribenko, 2008, J oint three-dimensional of magnetotelluric and
magnetovariational data: Proceedings of Annual Meeting of Consortium for
Electromagnetic Modelling and Inversion, 349-378.
Zhdanov, M. S., I. M. Varentsov, J . T. Weaver, N. G. Golubev, and V. A. Krylov, 1997,
Methods for modelling electromagnetic fields. Results from COMMEMI - the
international project on the comparison of modeling methods for electromagnetic
induction: J . Appl. Geophys., 37, 133-271.

Das könnte Ihnen auch gefallen