You are on page 1of 386

Modeling and Inversion Methods

for the Interpretation of


Resistivity Logging Tool Response

Barbara Ina Anderson


Modeling and Inversion Methods
for the Interpretation of
Resistivity Logging Tool Response

PROEFSCHRIFT

ter verkrijging van de graad van doctor


aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof. ir. K.F. Wakker,
voorzitter van het College voor Promoties,
in het openbaar te verdedigen op maandag 15 oktober 2001 om 13.30 uur

door

Barbara Ina ANDERSON

B. Sc., Western Connecticut State University


geboren te Danbury, Connecticut, USA
Dit proefschrift is goedgekeurd door de promotoren:
Prof. dr. ir. H. Blok
Prof. dr. ir. J.T. Fokkema

Samenstelling promotiecommissie:
Rector Magnificus, voorzitter
Prof. dr. ir. H. Blok, Technische Universiteit Delft, promotor
Prof. dr. ir. J.T. Fokkema, Technische Universiteit Delft, promotor
Prof. dr. ir. P.M. van den Berg, Technische Universiteit Delft
Prof. ir. C.P.J.W. van Kruijsdijk Technische Universiteit Delft
Prof. dr. S. Luthi, Technische Universiteit Delft
Prof. dr. ir. C.P.A. Wapenaar, Technische Universiteit Delft
Dr. T.M. Habashy, Schlumberger-Doll Research, USA, guest

Published and distributed by: DUP Science


DUP Science is an imprint of:
Delft University Press
P.O. Box 98
2600 MG Delft
The Netherlands
Telephone: +31 15 27 85 678
Telefax: +31 15 27 85 706
E-mail: DUP@Library.TUDelft.NL

ISBN 90-407-2231-5

Copyright 
c 2001 by Schlumberger Technology Corporation

All rights reserved. No part of the material protected by this copyright notice may be repro-
duced or utilized in any form or by any means, electronic or mechanical, including photocopying,
recording or by any information storage and retrieval system, without written permission from the
publisher: Delft University Press

Printed in The Netherlands


Contents

1 Introduction 1
1.1 Overview of the thesis . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Introduction to well logging . . . . . . . . . . . . . . . . . . . 3
1.3 Computer modeling in log interpretation . . . . . . . . . . . . 7
1.4 Anisotropy in log interpretation . . . . . . . . . . . . . . . . . 15
1.5 Inversion in layered anisotropic media . . . . . . . . . . . . . 19

2 Electromagnetic relations for logging 27


2.1 Overview of logging environments . . . . . . . . . . . . . . . . 27
2.1.1 Borehole effect . . . . . . . . . . . . . . . . . . . . . . 29
2.1.2 Coaxial layers; invasion . . . . . . . . . . . . . . . . . 30
2.1.3 Thin beds (bed boundary discontinuities) . . . . . . . 32
2.1.4 Invaded thin beds . . . . . . . . . . . . . . . . . . . . 33
2.1.5 Dipping beds . . . . . . . . . . . . . . . . . . . . . . . 34
2.1.6 3D geometries; horizontal wells . . . . . . . . . . . . . 35
2.1.7 Anisotropy in layered media; laminated formations . . 36
2.2 Description of logging tool configurations . . . . . . . . . . . 38
2.3 Electromagnetic field equations and notation . . . . . . . . . 41
2.4 Time domain equations . . . . . . . . . . . . . . . . . . . . . 43
2.5 Frequency domain equations . . . . . . . . . . . . . . . . . . . 44
2.6 Anisotropic media; the conductivity tensor . . . . . . . . . . . 45
vi CONTENTS

2.7 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . 47


2.8 Transform for axisymmetric configurations . . . . . . . . . . . 49

3 Electrical well-logging measurements 51


3.1 What do “resistivity” tools measure . . . . . . . . . . . . . . 51
3.2 Induction tools . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.2.1 Two-coil sonde response . . . . . . . . . . . . . . . . . 56
3.2.2 Early induction tools; “focused” sondes . . . . . . . . 64
3.2.3 6FF40 and the Dual Induction tool; a standard is set . 70
3.2.4 Phasor processing and deconvolution . . . . . . . . . . 80
3.2.5 Array Induction Tool (AIT) . . . . . . . . . . . . . . . 87
3.2.6 Russian induction tools . . . . . . . . . . . . . . . . . 95
3.3 Propagation tools . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.3.1 2-MHz tools for logging while drilling . . . . . . . . . 100
3.3.2 Deep Propagation Tool (DPT) . . . . . . . . . . . . . 108
3.3.3 Electromagnetic Propagation Tool (EPT) . . . . . . . 112
3.4 Electrode (laterolog) tools . . . . . . . . . . . . . . . . . . . . 116
3.4.1 The Normal . . . . . . . . . . . . . . . . . . . . . . . . 117
3.4.2 The Lateral . . . . . . . . . . . . . . . . . . . . . . . . 124
3.4.3 Russian BKZ tools . . . . . . . . . . . . . . . . . . . . 127
3.4.4 Laterolog 7 (LL7) . . . . . . . . . . . . . . . . . . . . 130
3.4.5 Laterolog 3 (LL3) . . . . . . . . . . . . . . . . . . . . 136
3.4.6 Laterolog 8 (LL8) . . . . . . . . . . . . . . . . . . . . 139
3.4.7 The Dual Laterolog tool (DLT) . . . . . . . . . . . . . 143
3.4.8 The Spherically Focused Log (SFL) . . . . . . . . . . 153
3.4.9 High Resolution Laterolog Array (HRLA) . . . . . . . 160
3.5 Microresistivity tools . . . . . . . . . . . . . . . . . . . . . . . 166
3.5.1 The Microlog . . . . . . . . . . . . . . . . . . . . . . . 168
3.5.2 The MicroLaterolog . . . . . . . . . . . . . . . . . . . 170
3.5.3 The Proximity log . . . . . . . . . . . . . . . . . . . . 171
CONTENTS vii

3.5.4 The MicroSpherically Focused Log (MSFL) . . . . . . 173


3.6 Imaging tools . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
3.6.1 Formation MicroScanner (FMS) . . . . . . . . . . . . 175
3.6.2 Azimuthal Resistivity Imager (ARI) . . . . . . . . . . 178
3.6.3 Resistivity-At-the-Bit tool (RAB) . . . . . . . . . . . 179
3.6.4 Oil-Base MicroImager tool (OBMI) . . . . . . . . . . . 180
3.7 Resistivity through casing . . . . . . . . . . . . . . . . . . . . 182

4 Modeling of tool response 185


4.1 Analytical methods . . . . . . . . . . . . . . . . . . . . . . . . 185
4.1.1 Doll’s induction geometrical factor theory . . . . . . . 186
4.1.2 Induction skin effect in homogeneous media . . . . . . 190
4.1.3 Induction real axis, spectral integration . . . . . . . . 197
4.1.4 The induction Born response function . . . . . . . . . 205
4.1.5 Laterolog response . . . . . . . . . . . . . . . . . . . . 212
4.2 Numerical methods . . . . . . . . . . . . . . . . . . . . . . . . 214
4.2.1 The finite element method . . . . . . . . . . . . . . . . 216
4.2.2 The finite difference method . . . . . . . . . . . . . . . 225
4.3 Hybrid methods . . . . . . . . . . . . . . . . . . . . . . . . . 233
4.3.1 Fast semi-analytic (mode matching) . . . . . . . . . . 233
4.3.2 With/without skin effect hybrid . . . . . . . . . . . . 244
4.4 Glossary of computer codes . . . . . . . . . . . . . . . . . . . 247
4.4.1 Induction codes . . . . . . . . . . . . . . . . . . . . . . 247
4.4.2 Laterolog codes . . . . . . . . . . . . . . . . . . . . . . 249

5 Using modeling in log interpretation 253


5.1 Relating resistivity logs to rock physics . . . . . . . . . . . . . 253
5.2 Early 1D plus 1D “inversion” efforts . . . . . . . . . . . . . . 257
5.2.1 Deconvolution and boosting . . . . . . . . . . . . . . . 258
5.2.2 Correction chartbooks and departure curves . . . . . . 263
5.3 A 2D iterative forward modeling case study . . . . . . . . . . 274
viii CONTENTS

5.4 A least squares inversion example in thin beds . . . . . . . . 279

6 Parametric inversion 285


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
6.2 Forward modeling . . . . . . . . . . . . . . . . . . . . . . . . 289
6.3 2-MHz inversion in layered media . . . . . . . . . . . . . . . . 293
6.3.1 2-MHz tool response in anisotropic media . . . . . . . 293
6.3.2 The inversion algorithm . . . . . . . . . . . . . . . . . 295
6.3.3 2-MHz inversion results . . . . . . . . . . . . . . . . . 305
6.4 Triaxial inversion in layered media . . . . . . . . . . . . . . . 307
6.4.1 Triaxial tool response in some limiting cases . . . . . . 307
6.4.2 Triaxial inversion results . . . . . . . . . . . . . . . . . 318
6.5 Summary and future plans . . . . . . . . . . . . . . . . . . . . 326

Bibliography 333

Summary 361

Samenvatting 365

About the author 369

Acknowledgments 373

Index 375
Chapter 1

Introduction

Summary: This chapter introduces the reader to the world of borehole well logging
from a historical perspective. In addition to describing the evolution of resistivity
measurements, nuclear and acoustic measurements are briefly summarized as well.
The important role that mathematical modeling has played in the design and in-
terpretation of resistivity measurements is illustrated with computed log examples.
The unique log interpretation problems created by resistivity anisotropy are traced
back to experiments performed by Conrad Schlumberger in the 1920’s. Paramet-
ric inversion is proposed as a method for quantifying resistivity anisotropy from
borehole measurements.

1.1 Overview of the thesis

The purpose of this thesis is twofold:

1. To provide an overview of the use of mathematical modeling in resis-


tivity log interpretation, and
2. To describe a new inversion method for determining formation electri-
cal properties in anisotropic layered media.

When computationally efficient 2D modeling codes were first developed


in the 1980’s, the author of this thesis began to experiment with incorporat-
ing forward modeling directly in the log interpretation process [14, 22]. One
2 CHAPTER 1. INTRODUCTION

of the first practical uses of modeling was to improve estimates of reserves in


dipping, thinly-bedded reservoirs [117, 19]. When these improved estimates
were validated by production history, the use of modeling exploded.
Today, forward modeling and inversion methods based on forward model-
ing are routinely used to improve the accuracy of log interpretation and even
to steer the drilling of directional wells in complex reservoirs [169]. Modeling
accurately accounts for the multidimensional and often nonlinear aspects of
tool physics–aspects that were previously “corrected” on a point-by-point
basis by using 1D charts or algorithms.
The foremost reason for opening this thesis with a modeling overview is to
respond to requests to “take the mystery out of modeling,” that is, to bridge
the gap between log interpretation and computational physics. There are
many excellent books that explain the geophysical and petrophysical aspects
of log interpretation [253, 86, 147, 208, 111]. There are also many excellent
texts describing methods for solving problems in electromagnetic theory [161,
241, 68, 153, 243]. However, log analysts and petroleum engineers often
complain that there are no texts that explain how to go from Maxwell’s
equations to computing resistivity tool response in understandable terms.
The objective of the overview is to fulfill this need in sufficient detail for
an interested reader to construct elementary codes for computing synthetic
resistivity tool response.
Another important reason for including and overview is to examine in
detail the various environmental effects that can complicate inversion. These
include effects of the borehole, invaded zone, shoulder beds and formation
dip. Often some nearly linear effects, such as those of the borehole, can be
addressed separately prior to the inversion process. This prior treatment
can simplify the inversion for formation anisotropy.
The overview begins with a description of typical logging environments
and a review of the fundamental electromagnetic field equations and rela-
tions for logging tools. There will be an examination of the electrical logging
tools under consideration (e.g., laterologs, induction, 2-MHz for logging while
drilling, propagation tools), including a description of the measurement char-
acteristics and volumes of investigation of each tool. Next, there will be a
review of forward modeling methods and computer codes commonly used
to simulate logging tool response. The evolution of inversion in log inter-
pretation will be illustrated with several case studies demonstrating simple
inversion methods, such as the use of departure curve charts and iterative
1.2. INTRODUCTION TO WELL LOGGING 3

forward modeling of multiple tools with user intervention.


These studies eventually lead to the parametric inversion of resistivity
tool response in layered anisotropic media. The problem will encompass both
the forward model development and the inversion. In the forward model, the
general problem of anisotropy in planar layered media will be considered in
the frequency domain. The axis of anisotropy is assumed to be oriented in
a general direction which varies from the direction of the tool axis. The
anisotropy is assumed to occur in conductivity, permittivity and magnetic
permeability of the layered-earth formation. In the inversion, the formation
will be described by layers whose parameters are to be retrieved through a
nonlinear optimization scheme.
The potential applications of the inversion will cover tool response in
dipping beds or deviated wells. Environments where invasion contributes
significantly will not be addressed.

1.2 Introduction to well logging

Electrical well logging was the first logging method used below ground in
boreholes by the petroleum industry. Of all the rock parameters measured
by logging tools, the electrical resistivity is of particular importance. Re-
sistivity measurements are essential for determining the relative amount of
hydrocarbons in a formation. In simplest terms, high resistivity indicates the
possible presence of oil or gas in rock pores, since hydrocarbons are insula-
tors. On the other hand, low resistivity indicates water, the other fluid that
may be present. The specific formulas that are used to determine the exact
amounts of hydrocarbons and water present from resistivity measurements
are given in Section 5.1.
Borehole logging was an outgrowth of prior techniques for exploring the
underground from the surface by means of electrical measurements. The
first electrical surface prospecting experiments were carried out in 1912 by
Conrad Schlumberger. These experiments consisted of sending an electrical
current between two metallic rods driven into the earth and drawing a map
of lines of constant potential observed at the surface. The shape of these lines
indicated the nature and geometrical configuration of subsurface geological
bodies permeated by the electric field (equipotential lines elongated when
an adjacent resistive body was approached [5]). From 1912 until World War
I measurement techniques were progressively improved, and in 1920 Conrad
4 CHAPTER 1. INTRODUCTION

Figure 1.1: The first log: points plotted on graph paper by Henri Doll and
annotated with a description of the formation layers.

Schlumberger and his brother Marcel founded a surface prospecting company


bearing the family name.
The first electrical log in a borehole was recorded on September 5, 1927,
by the Schlumberger brothers and Henri Doll (Conrad’s son-in-law and the
company’s chief theoretician) in the Pechelbronn field in Alsace-Lorraine. A
portion of this log is reproduced in Figure 1.1 [220]. The electrical resistivity
of the rock formation cut by the borehole was recorded at approximately one
meter depth intervals and plotted by hand.
The concept of “apparent resistivity” allowed the data to be scaled in
absolute units that are independent of the electrode configuration and the
intensity of the current. This generality was the major reason for the com-
mercial success of the Schlumberger company’s logging methods. The mea-
surement configuration is shown in Figure 1.2 [238]. In early literature this
1.2. INTRODUCTION TO WELL LOGGING 5

Figure 1.2: Electrode configuration of the first electrical well logging tool.

is referred to as a “lateral” or “inverse” sonde.


Three electrodes, A, M and N were lowered into the borehole, each at the
end of an insulated conductor. The current (I) emitted by A flows through
the mud inside the borehole and spreads across the formation as it returns
to B near the surface. The difference in potential between M and N (V ) is
transmitted to the surface and measured. The apparent resistivity (Ra ) is
evaluated using the formula
Ra = K · V /I, (1.1)
where K is a tool constant determined by the geometry of the electrode
system AM N B. Ra characterizes resistivity of the formation layer at the
M N level.
“Carrotage électrique” was the name that the Schlumberger brothers
gave to borehole electrical prospecting. This translates from the French as
electrical coring, meaning that sensors lowered into a well on a cable were
a replacement for the time-consuming and expensive cutting of cylinders
of rock (cores) as wells were drilled. Until that time coring was the main
method of learning about the formations that the drill bit had penetrated. In
the 1930’s the term “electrical coring” was replaced by “electrical logging.”
The word “log” referred to the strip of paper on which the curves were
6 CHAPTER 1. INTRODUCTION

plotted, borrowing from nautical science where it denotes the recording of


the position of a ship in terms of time.
The petroleum industry quickly recognized the usefulness of resistivity
logging for the identification of potential hydrocarbon-bearing zones and for
correlation purposes. In 1929 electrical resistivity logging was introduced
on a commercial basis in the United States, Venezuela and Russia. In 1931
the spontaneous potential (SP) measurement was included along with the
resistivity curve on the electrical log, after it was discovered (by chance) that
the M N circuit could measure natural potentials of electrochemical origin
which indicated permeable layers.
In 1931, the Schlumberger brothers also perfected a method of continuous
recording and the first pen recorder replaced the point-by-point system. The
automatic photographic film recorder (single galvanometer) was introduced
in 1936, eliminating tedious hand-copying of logs. By that time, the log-
ging suite consisted of the SP curve, the lateral and long and short normals
(normal tools have the N electrode at or near the surface). This combina-
tion dominated logging until the 1950’s, when focused electrode tools and
induction tools (described in Chapter 3) came into use.
Experimental dipmeter tools were used in the 1930’s to help identify
major geologic structures. They were greatly improved during the 1940’s and
became the principal logging tool for describing internal lithologic features.
Dipmeter tools consist of four evenly spaced pads that are pressed against
the borehole wall. Each pad contains a short micro-resistivity device. The
four micro-resistivity curves are correlated to find the difference in depth
between bedding markers around the borehole, which yields the magnitude
and azimuth of formation dip [225].
The major uses of electrical logging tools were to infer geologic structure
and to determine the nature of fluids in sedimentary rocks from the measured
resistivity. Over the years, other types of tools were introduced to determine
additional physical properties, such as rock density, porosity, radioactivity
and sound transmission.
Nuclear measurements were developed after World War II. The gamma
ray and neutron tools were the first borehole measurements of radioactive
properties (they were also the first tools to use downhole electronics). Unlike
resistivity tools, nuclear tools are able to log formations through steel casing.
The basic gamma ray (GR) log was introduced in the 1950’s as a per-
meability indicator. It measures the natural formation radioactivity which
1.3. COMPUTER MODELING IN LOG INTERPRETATION 7

reflects shale content.


In the 1960’s the compensated neutron tool (CNL) gained acceptance as
a porosity measurement, inferring porosity from the energy loss of emitted
neutrons (compensation refers to the use of two sets of antennas whose re-
sponses are averaged together to cancel errors from sonde tilt and hole size
changes). The compensated density tool (CDL) was also introduced in the
1960’s. It infers bulk density, a property primarily dependent on porosity,
from the attenuation of emitted gamma rays.
Since the 1930’s, geophones had been lowered into oil wells on logging
cables to measure long-interval acoustic travel times from sound sources
at the surface. In the 1950’s the borehole sonic log gained acceptance as
a reliable porosity log because its travel time measurement is essentially
independent of fluid saturation.
It may seem redundant to have three porosity tools in common use when
only one porosity value is needed. However, the three tools respond not only
to porosity but also to the type of the rock matrix and the nature of the
fluid filling the pore space. When the rock and fluid types are unknown, all
three measurements are needed to sort out parameters [86].
From the 1950’s through the 1980’s, a typical logging suite consisted of
a focused resistivity tool (laterolog or induction), SP, a neutron/density log
and a sonic log. The advent of digital signal processing and transmission in
the 1980’s lead to the modernization of existing tools and the eventual intro-
duction of array induction, laterolog and sonic tools in the 1990’s (modern
resistivity tools are described in detail in Chapter 3). Other currently evolv-
ing measurements include nuclear magnetic resonance, nuclear spectrometry
and electrical and acoustic imaging.
The 1990’s also saw the growth of logging while drilling (LWD), that is
the placement of electrical, nuclear and acoustic tools on the drill string to
record measurements just behind the drill bit as it cuts through the forma-
tion. This early time data is used along with information from exploration
wells to steer drilling in the direction of hydrocarbon-bearing zones.

1.3 History of computer modeling in log interpretation

Technological progress in tool development was accompanied by the evolu-


tion of the new discipline of log interpretation. Very few of the petrophys-
8 CHAPTER 1. INTRODUCTION

Figure 1.3: Sample chart for interpreting an early lateral tool. Lateral ap-
parent resistivity is plotted in ordinate and distance relative to bed thickness
is plotted in abscissa. Four ratios of bed thickness (e) to tool spacing (L)
are shown. The resistivity of the central bed is 25 ohm-m and the resistivity
of the surrounding beds is 5 ohm-m. Note that the shapes of the logs are
considerably different for the four bed thicknesses. Also note the large dif-
ference between the log resistivity and the actual resistivity in each of the
central beds. In addition, note the large overshoot above 25 ohm-m on the
top left log.
1.3. COMPUTER MODELING IN LOG INTERPRETATION 9

ical properties needed to evaluate the amount of hydrocarbons in place in


a reservoir can be measured directly. The most important petrophysical
properties used in formation evaluation are porosity (pore volume per unit
volume of formation), water and hydrocarbon saturation (fraction of pore
volume occupied by fluid) and permeability (ease with which fluids flow).
Log interpretation applies known physical relationships to the parameters
measured by logging tools (resistivity, bulk density, travel time, radioactiv-
ity, etc.) in order to obtain a quantitative evaluation of the above mentioned
petrophysical properties.
Mathematical modeling has been intimately associated with electrical log
interpretation since the time that the first log was run for two basic reasons:
(1) electrical tools survey large volumes of formation making it necessary to
quantify parasitic effects caused by regions adjacent to beds of interest, and
(2) electrical tool response is highly nonlinear.
The Schlumberger brothers used small electrodes in saltwater baths to
perform early experimental modeling. Soon afterwards, mathematicians
from the École des Mines in Paris were enlisted to solve the problem of a
point electrode tool logging perpendicular to vertical layers using Maxwell’s
image theory [5]. This solution served as a basis for calculating numerous
sets of theoretical departure curves for normal and lateral tools which were
published in booklets throughout the 1930’s (departure, in this case, refers
to the difference in resistivity between tool response in a thin bed and the
unperturbed response in an infinitely thick bed). Interpretation consisted
of superimposing a transparent chart over a log and finding the theoretical
curve that gave the best coincidence. However, this method was only de-
pendable for at most three layers. An example of an early chart is shown in
Figure 1.3 [183].
Chartbooks of theoretical departure curves were routinely used to inter-
pret resistivity logs from the 1930’s through the 1970’s (some of the most
commonly used charts are described in Chapter 5). After induction tools
were introduced in the 1950’s, charts were produced to correct these tools
separately for both the effect of shoulder beds and invasion of the borehole
mud into the formation [215]. These charts were generated using computer
programs that modeled 1D analytical solutions of Maxwell’s equations [194].
2D interpretation was achieved by applying 1D corrections in sequence; a
layered media correction for shoulder bed effect was performed first, followed
by a cylindrical media correction for invasion effect.
10 CHAPTER 1. INTRODUCTION

Figure 1.4: The resistor network.

In 1950 a resistor network [131] was introduced for simulating electrode


tool response in more realistic 2D logging environments consisting of a bore-
hole and multiple thin beds with invaded zones. The network, shown in
Figure 1.4, consisted of tens of thousands of resistors and was in effect an
analog computer. Charts generated by the resistor network soon replaced
the earlier layered media charts, which suffered from inaccuracy caused by
ignoring borehole effect.
Starting in the late 1960’s, work was begun on 2D axisymmetric finite
element and finite difference codes for modeling both electrode tool and in-
duction response [175, 182, 176]. Although these numerical methods had
been successful for small-scale problems in the power industry, they proved
impractical for simulating resistivity tool response at that time because ex-
isting computer memory and speed were insufficient for modeling electric
currents that penetrated tens of meters from the borehole.
Large improvements in computing capabilities in the late 1970’s reduced
the time required to compute finite element and finite difference simulated
1.3. COMPUTER MODELING IN LOG INTERPRETATION 11

logs from weeks to hours. In 1980, the resistor network was “retired” to
the Schlumberger museum in France and replaced by a 2D finite difference
code [119]. Shortly after this, a 2D finite element code for modeling induction
tool response came into common use [62].
These codes were at first used to aid in tool design and to generate inter-
pretation charts. However, two changes occurred in the petroleum industry
in the 1980’s which led to computer modeling assuming a more active role
in log interpretation. The first was the growth in economic importance of
thinly bedded reservoirs. Resistivity tools of that time were designed to
be relatively free of effects of adjacent layers in beds thicker than six feet.
After the era of “easy oil” was over, one-to-two foot beds needed to be in-
terpreted. The application of 1D plus 1D chartbook corrections proved to
be highly inaccurate in these thin beds.
The second change was the advent of horizontal drilling. All published
charts had been generated for vertical wells, with tools logging perpendicular
to bed boundaries. These charts no longer applied when tools logged parallel
to boundaries in horizontal wells. As more and more charts became obsolete,
it became clear that another approach to interpretation was needed.
Fortunately the 1980’s also saw a continued evolution in computer power.
Personal computers were introduced that could run 2D modeling codes which
previously required large mainframe parallel machines. Continued advances
in numerical techniques [70] made it possible to compute simulated logs
in minutes instead of hours. This set the stage for the integration of tool
modeling with log interpretation.
In the 1980’s several papers were written by the author of this the-
sis [14, 15, 16] which demonstrated the power of iterative forward modeling
in log interpretation for the first time. These papers showed how forward
modeling could be applied to accurately determine formation resistivity in
complex formations that were beyond the scope of chartbook interpreta-
tion. Soon after this, a user-friendly electromagnetic modeling package called
ELMOD [19] was made available to Schlumberger log analysts for use on
personal computers at regional log interpretation centers.
The systematic application of forward modeling in log interpretation is
illustrated in the flowchart in Figure 1.5. Estimates of formation resistivities
and bed boundary dimensions are obtained from either visual inspection or
computer algorithms (i.e., bed boundaries from log inflection points and re-
sistivities from maximum/minimum values). These parameters are used to
12 CHAPTER 1. INTRODUCTION

Figure 1.5: Flowchart illustrating the use of forward modeling in log inter-
pretation.

set up an initial formation model for a given section of log. The modeling
code is then run to simulate tool response in this formation in an attempt to
generate a computed log that overlays the field log. If the two logs disagree,
then the formation model is refined, either by systematically varying param-
eters or by incorporating additional information from other logs or cores.
The process is repeated until reasonable agreement is achieved. The final
formation model provides the resistivity values in each layer. Even though
solutions obtained in this way are not necessarily unique, modeling can nev-
ertheless serve to eliminate impossible scenarios and validate the most likely
interpretation.
The first successful use of ELMOD was in improving the determination
of hydrocarbon reserves from induction logs in deviated wells in the North
Sea. A series of ELMOD runs was used to find a squared resistivity profile
that would reproduce an induction field log. The steps involved in finding
the solution are illustrated in Figure 1.6 [19], using a section of an actual
field log.
Dipmeter logs in this well indicated that the combined hole deviation and
formation dip gave a total dip of 38◦ . In order to determine the characteristic
response of induction tools at that dip angle, the log analysts involved first
consulted published examples of dip effect [32, 143]. Inflection points on
1.3. COMPUTER MODELING IN LOG INTERPRETATION 13

Figure 1.6: Three iterative forward modeling runs are used to find the for-
mation resistivity in a North Sea case study.
14 CHAPTER 1. INTRODUCTION

the induction curve were used as initial bed boundary locations. Using
this information, a trial formation was set up and induction response was
modeled.
In Figure 1.6, the log on the left (simulation 1) shows the first assumed
formation resistivity profile (square Rt ), along with the field log and the
computed log (IDPH is the deep Phasor induction tool). Although the two
logs agree fairly well in the center of most beds, the shape of the computed
log isn’t correct near the bed boundaries. The second model in the middle
(simulation 2) adjusts bed boundary locations and fine-tunes some resistivity
values, making the computed log agree more nearly with the field log. The
final model on the right (simulation 3) adjusts for overcompensation and
further refines the shape of some beds. The square formation now gives a
computed log that agrees very closely with the field log.
Note the difference in resistivity level between the field log and the final
square formation resistivity in the two resistive pay zones (40 ohm-m versus
200 ohm-m at 1040 feet, and 60 ohm-m versus 150 ohm-m at 1100 feet).
This difference is a result of dip effect. If the resistivity read by the tool
was used in reserve calculations, the amount of hydrocarbons in place would
be severely underestimated. The log analysts involved in this study cited
an additional benefit of modeling: it gave them a better insight into tool
physics which they could apply to future interpretations.
The iterative forward modeling process could of course be replaced by in-
version. Indeed, in the 1980’s several authors proposed inverse solutions [170,
106, 115, 138, 71, 267] for resistivity logging. However, computers at that
time were still too slow to make inverse solutions practical. In addition,
inverse solutions for the tools of the 1980’s were plagued by nonuniqueness
to an even greater extent than iterative forward modeling solutions.
The problem of nonuniqueness is illustrated by the two logs in Figure 1.7.
The log in the 2 ohm-m bed on the left is identical to the log in the alternat-
ing 1–100 ohm-m laminated zone on the right. Nonuniqueness caused by a
tool’s poor vertical resolution, such as in this case, is not a major problem in
iterative forward modeling. During the iterative process, formation models
can be severely constrained by local knowledge from cores or higher resolu-
tion logs (such as nuclear or imaging logs). Commercial inversion software
for resistivity logging is not implemented to access non-resistivity informa-
tion, although this problem is currently receiving considerable attention.
The introduction of high resolution array tools with multiple depths of
1.4. ANISOTROPY IN LOG INTERPRETATION 15

Figure 1.7: Identical 6FF40 logs generated by two different formation models
illustrating the problem of nonuniqueness in resistivity log inversion.

investigation in the 1990’s has made reliable inverse solutions possible. Re-
cently, maximum entropy log inversion (MERLIN) [49] was developed for the
Schlumberger AIT Array Induction tool to provide more accurate Rt and in-
vasion interpretation in highly deviated wells. AIT response to invasion in
vertical wells has also been inverted to generate fractional flow logs which
display saturations [207]. For the Schlumberger HRLA Array Laterolog, 2D
imaging inversion [237] is used to obtain Rt and the invasion profile. Baker-
Atlas has also developed and documented inversion algorithms for both their
array induction [248] and array laterolog [142] tools.

1.4 Anisotropy in log interpretation

Anisotropy (the variation of properties with direction) is not uncommon in


sedimentary strata. Many solid particles have flat or elongated shapes that
are usually oriented parallel to the plane of deposition as shown in Fig-
ure 1.8 [21]. This results in a pore structure that allows electric current to
flow more easily parallel to the bedding plane than perpendicular to it [112]
16 CHAPTER 1. INTRODUCTION

Figure 1.8: Scanning electron photomicrograph showing aligned grains in a


limestone sample.

(the conducting medium is the water saturating the rock pores). Sedimen-
tation of this type produces transversely isotropic (TI) anisotropy, that is,
the horizontal resistivity (Rh ) is the same in every direction in the horizon-
tal bedding plane, while the vertical resistivity (Rv ) normal to the bedding
plane is different. Particle shape anisotropy is most commonly found in
shales, and may also occur in sands and carbonates.
Although we are concerned with electrical anisotropy, it is important to
note that the same sedimentary processes that cause electrical anisotropy
can result in anisotropy in other physical parameters. Permeability anisot-
ropy is particularly important in determining hydrocarbon flow in reservoirs.
Currently work is being carried out to find relationships between electrical
anisotropy and permeability anisotropy [254, 159].
Anisotropy depends very much on scale. In addition to microscopic an-
isotropy occurring at the particle scale, formations consisting of a series of
isotropic beds of different lithology (such as sequences of sand and shales)
also behave anisotropically if a logging tool is significantly longer than the
bed thickness. This is referred to as macroscopic anisotropy. The two logs
in Figure 1.7 are identical because the eight foot induction tool averages the
one foot resistive and conductive layers (on the right), reading an effective
resistivity which is equivalent to the resistivity in the thick bed (on the left).
1.4. ANISOTROPY IN LOG INTERPRETATION 17

When logging perpendicular to bed boundaries in cases such as this, resistiv-


ity tools read the effective horizontal resistivity, Rh , which can be calculated
from the volume average of the layer conductivities (inverse resistivities),

1 1 1
= Vsand · + Vshale · , (1.2)
Rh Rsand Rshale

where resistivities are expressed in ohm-m and Vsand and Vshale are the bulk
volume fractions (percentages) distributed throughout the layered region
(layers are all assumed to be approximately uniform in thickness). The
effective vertical resistivity, Rv , can be calculated in a similar manner from
the volume average of the layer resistivities,

Rv = Vsand · Rsand + Vshale · Rshale . (1.3)

As early as 1920, Conrad Schlumberger recognized that anisotropy af-


fected surface prospecting measurements [223]. In 1932, Maillet and Doll [181]
presented a method for interpreting surface potential measurements in aniso-
tropic formations. They showed that a TI anisotropic medium could be
rescaled to an isotropic medium using the anisotropy coefficient λ, defined
as 
λ = Rv /Rh . (1.4)
The isotropic medium was assigned an effective resistivity (geometric mean)
denoted as R, with 
R = Rv · Rh . (1.5)
These results were used to design an experimental electromagnetic surface
prospecting device that determined the direction of formation dip from mea-
surements of the horizontal and vertical components of the magnetic field [5].
For both electrode and induction tools, the apparent resistivity (Ra ) in
a TI anisotropic medium can be calculated using the approximation [193]

Ra = R/ sin2 α + λ2 cos2 α, (1.6)

where α is the angle between the tool axis and vertical. For α = π/2
(surface prospecting or horizontal wells), Ra = R. For α = 0 (vertical
wells), Ra = Rh . Thus the vertical resistivity cannot be detected at all by
conventional resistivity logging tools in vertical wells. This is sometimes
referred to as the “paradox of anisotropy”.
18 CHAPTER 1. INTRODUCTION

Although both electrode and induction tool response is theoretically the


same in homogeneous anisotropic media, in the 1950’s it was noticed that
16 inch Normal logs sometimes read higher resistivity values than induction
logs in shales. This prompted Kunz and Moran [165] to investigate borehole
effect in anisotropic formations. Adding a borehole filled with conductive
mud to the vertical well model, they showed that Rv can affect electrode
tool response since current has a considerable vertical component as it travels
between the source and the return.
Twenty years later, this work was extended by Moran and Gianzero [193]
to model both induction and electrode tool response to dipping bedding
planes (with no borehole). In the same paper, they proposed a technique for
measuring anisotropy using a combination of horizontal and vertical coils,
since vertically oriented coils are sensitive to Rv . However, they concluded
that borehole and bed boundary effects would make the method impractical.
In a later paper [112], the same authors proposed a sidewall pad device to
overcome borehole effect.
When laterolog tools were introduced, it was assumed that they measured
Rh with negligible influence of Rv since these tools use bucking currents to
force the survey current laterally into the formation (see a yet unnumbered
figure in Chapter 3 showing current lines). However, discrepancies between
induction and laterolog measurements were still noted in shales and also in
laminated sand–shale sequences. Chemali, et al. [64], showed that laterologs
still responded appreciably to Rv , although to a lesser degree than unfocused
electrode tools. They generated charts for evaluating λ from differences be-
tween laterolog and induction logs in dipping and horizontal beds. However,
the method is seldom used because in most cases the difference is so small
that it is less than the precision of the measurements.
Anisotropy can also enter into the interpretation of ULSEL logs. The
ULSEL (Ultra Long Spaced Electrical Logging) tool was developed in the
1960’s [214]. It is used to locate distant resistive anomalies such as salt domes
which act as traps for hydrocarbons. The ULSEL tool consists of four to six
long normal arrays with spacings ranging from 75 to 2400 feet. The depth
of investigation of ULSEL is approximately 2000 feet from the wellbore.
Accurate location of a salt dome (normally to the side of a well) with such
long arrays depends on knowing both Rh and Rv . Since no measurement of
Rv is available, an induction or laterolog log in the same well is used to set up
a layered model of the formation. The theoretical ULSEL response is then
computed in this formation. The presence of a lateral salt dome is indicated
1.5. INVERSION IN LAYERED ANISOTROPIC MEDIA 19

when the ratio of the actual log (with salt dome) to the computed log (no
salt dome) is significantly greater than one. A comparison of differences
between the ratios of the various normals indicates the distance to the salt
dome.
In the early 1990’s ULSEL started to be used in large-scale reservoir
description [197], and ULSEL interpretation was updated. Borehole seismic
measurements, dipmeter logs and modeling codes including anisotropy now
help ULSEL predict distance and direction to any resistive or conductive
anomaly more accurately.

1.5 Parametric inversion in layered anisotropic media

From the 1920’s through the 1980’s, anisotropy was regarded as a secondary
effect on resistivity logs. Even though papers were written describing the
mathematical modeling of anisotropy and occurrences of anisotropy were
flagged on logs, anisotropy effect was rarely included in routine log inter-
pretation. Because most wells drilled up to the mid-1980’s were vertical or
only slightly deviated, resistivity tool sensitivity to Rv was negligible and
the effect of anisotropy was masked. Therefore modeling and inversion to
evaluate parasitic effects on beds of interest from adjacent zones (borehole,
neighboring beds, invasion) received primary attention.
However, the increased use of horizontal drilling in the late 1980’s and
the subsequent introduction of 2 MHz LWD resistivity tools revealed that
anisotropy could not be ignored in horizontal well interpretation. In fact,
anisotropy effect was often surprisingly larger than shoulder bed or invasion
effects in horizontal wells.
The interpretation of horizontal well data is a multi-step process. Prior
to drilling a horizontal well, potential hydrocarbon-bearing zones are first
located using vertical exploration wells. Then a horizontal well is drilled
toward a target bed, with marker beds used to maintain the wellbore tra-
jectory. Resistivity logs recorded behind the bit are compared to logs from
the exploration wells to identify the marker beds. Computer modeling of
predicted resistivity tool response at different well deviation angles (called
geosteering [9]) is used to modify the well path as needed. After a horizontal
well penetrates a hydrocarbon-bearing bed, drillers attempt to keep it inside
the bed for as long as possible. This procedure allows the well to drain a
large area, making a horizontal well more cost effective than several vertical
20 CHAPTER 1. INTRODUCTION

Figure 1.9: Wireline induction (left) and 2-MHz CDR (right) response to
anisotropy for Rv /Rh = 10 with Rh = 10 ohm-m.

wells.
When comparing resistivity logs in a horizontal well to logs from a verti-
cal exploration well, it was noticed that the resistivity values often differed in
shales and in laminated zones. This made identification of beds ambiguous,
posing a problem in steering a horizontal well toward a target bed. After a
closer examination of all available logs, cores and modeling, these differences
were attributed to anisotropy for the first time in 1991 [169].
Figure 1.9 illustrates typical differences between resistivity tool readings
in vertical wells (0◦ ) and horizontal wells (90◦ ) caused by anisotropy. At
0◦ dip, both the dual induction and Compensated Dual Resistivity (CDR)
tools accurately read Rh . As the dip (or deviation) angle increases, the deep
and medium induction curves both increase in the direction of Rv with little
separation between them. The CDR curves also increase in the direction
of Rv , with the phase shift resistivity reading higher than the attenuation
resistivity (this curve order is also characteristic of CDR response for values
of Rv /Rh other than 10).
The induction and CDR tools both generate azimuthally polarized elec-
tric fields which induce current loops that are tilted with respect to the
transverse anisotropy. These tilted current loops sense a weighted average
1.5. INVERSION IN LAYERED ANISOTROPIC MEDIA 21

of Rv and Rh which depends on dip angle. The response for induction tools
can be approximated from Equation (1.6). The low frequency (20-kHz) in-
duction response is fairly linear and not strongly sensitive to anisotropy. In
contrast, extensive modeling and analysis of the higher frequency (2-MHz)
CDR response by Lüling, et al. [180], using the approach of Moran and
Gianzero [193], has demonstrated that radiation effects control the phase
shift measurement more strongly than the attenuation measurement. Thus
separation between 2-MHz phase shift and attenuation logs provide a good
indication of anisotropy (in the absence of invasion and shoulder bed effect),
with sufficient resolution for inversion.
Resistivity tool sensitivity to Rv revealed by horizontal well interpreta-
tion prompted a reassessment of the phenomenon known as “low resistivity
pay” [59], which in turn led to proposals for tools that could measure Rv
directly. In some reservoirs, particularly in the Gulf of Mexico, hydrocar-
bons are produced from vertical wells in zones with resistivities between 0.5
to 5 ohm-m, values usually associated with fresh water production. With
such low resistivities, these zones were often bypassed. However, high reso-
lution resistivity imaging tools introduced after the late 1980’s (such as the
Formation MicroScanner and the LWD Resistivity-At-the-Bit tool) revealed
that many of these low resistivity zones consisted of laminated conductive
shales and resistive oil-bearing sands. The conductive shales were lowering
the average resistivity read by the induction tools. Occasionally these reser-
voirs were penetrated by horizontal wells, and resistivity tools read higher
than in the vertical wells, confirming anisotropy. The effective resistivity in
horizontal wells was influenced more by Rv , which was higher and nearer to
values normally expected in hydrocarbon-bearing zones.
Naturally, this generated interest in designing a tool that could measure
Rv in vertical exploration wells so that these productive zones would not
be bypassed. Calculations of vertical coil response in homogeneous aniso-
tropic media [193] have demonstrated that a transverse magnetic dipole tool
(TMD) is moderately sensitive to Rv in vertical wells. Unfortunately, more
recent calculations [201] have shown that TMD antennas are extremely sen-
sitive to borehole effect. Methods are currently being investigated to cancel
TMD borehole effect, either by means of hardware or software. With bore-
hole effect removed, layered media inversion algorithms are more accurate
and easier to implement. Triaxial antennas, which provide more information
for inversion, are also being investigated using a 3D anisotropic media finite
difference code [82].
22 CHAPTER 1. INTRODUCTION

Figure 1.10: CDR response at 0◦ dip (left) and 80◦ dip (right) as the tool
logs an isotropic bed above an anisotropic bed.

Historically, the first method used to solve for Rh and Rv in horizontal


wells was iterative forward modeling using a laminated formation model [20],
both for modeling laminations and to approximate bulk anisotropy using
Equation (1.2) and Equation (1.3). Subsequently, a code was written to
model induction and CDR tool response in anisotropic layered media [137],
eliminating the tedious task of setting up a lamination model. Examples of
typical CDR synthetic logs are shown in Figure 1.10. The log on the right
in Figure 1.10 illustrates separations between phase shift and attenuation
curves that are typically seen in anisotropic media in highly deviated wells.
The log on the left shows the insensitivity of the CDR tool to anisotropy in
a vertical well in the same formation.
In the early 1980’s, software to invert CDR response for Rh and Rv based
on the homogeneous anisotropic media solution of Moran and Gianzero [193]
was implemented for commercial use by Rosthal [211]. Results obtained by
applying this inversion to the 80◦ log in Figure 1.10 are shown in Figure 1.11
on the left. Since the log input to the inversion was generated by a layered
medium code [137], it is free of noise. The known information used in the
homogeneous medium inversion is the relative dip angle and the apparent
1.5. INVERSION IN LAYERED ANISOTROPIC MEDIA 23

Figure 1.11: Inversion for Rh and Rv for the 80◦ log of Figure 1.10. Results
based on the homogeneous medium solution are on the left and parametric
inversion results are on the right.

phase shift and attenuation resistivities. The closed form analytical solution
for tool response in homogeneous anisotropic media is solved iteratively by
a Newton-Raphson algorithm. An initial guess for Rh and Rv is obtained
from the log apparent resistivities and used to compute the corresponding
phase shift and attenuation resistivities at the given dip angle. The iteration
scheme uses the computed resistivities and their gradient with respect to
changes in Rh and Rv to obtain the next estimate. Iteration continues
until a solution is found to a specified accuracy or a maximum number of
iterations is exceeded. Typically about five iterations are required to reach
convergence.
Often a solution does not exist or it is physically unrealistic. Many other
environmental effects exist (invasion, borehole effect, response to dielectric
rock properties, shoulder bed effect) that cause separations between phase
shift and attenuation resistivity curves similar to those caused by anisot-
ropy. Note that in Figure 1.11 (left) the solution for Rh and Rv in the
anisotropic bed is only correct at distances greater than eight feet below the
bed boundary. In this case, the homogeneous medium inversion cannot ac-
24 CHAPTER 1. INTRODUCTION

curately account for shoulder bed effect and the polarization horn [20] that
occurs at bed boundaries at high dip angles. The height of a polarization
horn depends on resistivity contrast and horns are a common occurrence
near resistive hydrocarbon-bearing zones. In fact, horns are often used in
Geosteering as an indication that the well path has crossed into a target bed,
so they must be accurately taken into account in the model.
Parametric inversion based on a layered-earth model provides a means of
accounting for shoulder bed effect and polarization horns more accurately.
Results obtained using parametric inversion in the same 80◦ formation are
shown on the right in Figure 1.11, and will be described in greater detail in
Chapter 6. In this case, the known information used in the inversion is the
relative dip angle, the bed boundary location obtained from a boundary de-
tection algorithm and the apparent phase shift and attenuation resistivities.
It is assumed that the bed boundary location is known within an accuracy
of ±2 inches. Errors greater than 2 inches will degrade the inversion.
Triaxial measurements are proposed as a means of overcoming this dif-
ficulty. Triaxial measurements have sufficient sensitivity to anisotropy to
directly solve for the bed boundary locations and dip angle, in addition to
Rh and Rv .
The general geometry considered in this thesis consists of multiple, dip-
ping anisotropic thin beds. Borehole effect is not taken into consideration
because it is fairly linear and can be decoupled from the problem (com-
mercial software exists for pre-processing resistivity tool response to correct
for borehole effect). Invasion is also not considered here because it is nor-
mally shallow at early times during logging while drilling, the area where
anisotropy interpretation is of most interest.
The objective is to invert for the horizontal and vertical resistivities
within each bed from the apparent resistivity log. Two cases are consid-
ered. For the CDR inversion it is assumed that a fixed deviation angle can
be obtained from a dipmeter or imaging log. Fixed bed boundary locations
are obtained from inflection points on logs for small dip angles, or from peak
values of polarization horns for large dip angles. For the triaxial inversion,
the bed boundary locations and dip angle are not fixed, but are included
in inversion solution. In both cases, the initial guesses for Rh and Rv are
obtained from measured center-bed resistivity readings.
The inversion algorithm is an iterative approach based on the Gauss-
Newton method that employs a quadratic model of the cost function. The
1.5. INVERSION IN LAYERED ANISOTROPIC MEDIA 25

cost function is defined as the square of the sum of the relative residual er-
rors given by the difference between the log data and the estimated response
normalized to the log data. The step length is adjusted by line search to suf-
ficiently decrease mismatch between measured and predicted responses after
each iteration. The method is based on constrained minimization where up-
per and lower bounds are imposed on the inverted parameters. The forward
model is generated from the code ANISBEDS [137] which is an AC model
for arbitrarily oriented point dipoles. The same general parametric inver-
sion method has been applied to laterolog tools in isotropic invaded beds in
vertical wells by Habashy, et al. [136].
26 CHAPTER 1. INTRODUCTION
Chapter 2

Basic electromagnetic field relations for


logging tools

Summary: This chapter relates Maxwell’s equations to resistivity tool antenna


configurations and the borehole logging environment. Basic electromagnetic con-
cepts such as notation, boundary conditions and the conductivity tensor representa-
tion are defined as they apply to resistivity measurements. The logging environment
is characterized in terms of both geometry and geology, with emphasis on the de-
positional processes that give rise to anisotropy. The need for accurate modeling
and inversion is demonstrated by showing how readily resistivity measurements in
beds of interest can be corrupted by adjacent media because of the large volumes
of investigation of resistivity tools.

2.1 Overview of logging environments

The parameter of greatest interest in evaluating a reservoir for its hydrocar-


bon content is Rt , the resistivity of a bed under consideration which has not
been contaminated by borehole fluids. Logging tools measure the over-all
apparent resistivity, Ra , and in order to accurately determine Rt , perturba-
tions caused by adjacent regions must be taken into account. These regions
are shown in Figure 2.1 [222], and include:

- The borehole of diameter dh (6 to 16 inches), filled with drilling mud


28 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

Mud
t bed
Rm Adjacen

Rs

Uninvaded zone
Transition zone
Invaded zone
hmc

or Annulus
R xo
Bed thickness

Rt

Mudcake
h

dii
dj
meters
Invasion dia Adjacent
bed
Rs

dh

Borehole
diameter

Figure 2.1: The logging environment.

of resistivity Rm ,
- Zones encircling the borehole flushed by the borehole mud called in-
vaded zones, with resistivity Rxo and diameter di (ranging from dh to
200 inches, and occasionally larger),
- Adjacent layers of differing resistivity called shoulder beds, with resis-
tivity Rs and thickness h (ranging from several inches to 100 feet).

The effects of the borehole and adjacent beds can be decreased by de-
signing tools to minimize their effect or by computer processing. Invasion
can be resolved by using tools with several depths of investigation.
The first half of this chapter addresses the geometry of the logging envi-
ronment and the formation electrical characteristics which affect resistivity
tool modeling and inversion. The second half defines the subset of Maxwell’s
equations used for modeling resistivity tool response.
2.1. OVERVIEW OF LOGGING ENVIRONMENTS 29

2.1.1 Borehole effect

Because well logging is carried out with the tool immersed in the borehole
mud, mud properties and borehole size can affect the accuracy of the mea-
surement of Rt . For example, highly conductive mud can short-circuit lat-
erolog currents and prevent them from penetrating deeply into a formation.
Therefore it is important to accurately account for borehole effect.
Most wells are drilled with a rotary bit located at the end of a long
string of drill-pipe. A liquid mud is pumped down inside the drill-pipe and
out through holes in the bit, and returns to the surface in the annular space
between the drill-pipe and the borehole wall. The mud lubricates the bit
and carries cuttings to the surface. In addition, the mud prevents blowouts
by providing a weighted column of liquid whose hydrostatic pressure can be
adjusted to exceed that of the pore fluids in the formation [253].
The majority of drilling muds are water-based. These muds contain
weighting materials (usually clays) for adjusting the density, chemicals for
maintaing a desired pH and gels to adjust flow properties. The resistivity of
water-based mud is dependent mainly on its salinity. Muds made from sea
water can be very conductive, ranging from 0.005 to 0.1 ohm-m at downhole
temperatures. Muds made from fresh water are less conductive, ranging
from 0.01 to 5 ohm-m, depending on the blend of the additives [177].
Oil-based muds are also commonly used. These muds consist of a complex
mixture of oil, water, salt and surfactants necessary to keep the oil-water
mixture in emulsion. Although oil is the continuous phase, some oil-based
muds may contain as much as 40% water. The resistivity of oil-based mud is
typically about 1000 ohm-m or greater. Oil-based muds usually do not invade
the formation very deeply. However, high down-hole temperatures and the
effects of the surfactants can sometimes combine to produce moderately deep
invasion of either the water-phase or the oil-phase [168].
Borehole sizes commonly range between 6 and 10 inches in diameter, but
may be as large as 20 inches. The larger the hole, the greater the volume
of mud around the tool, and therefore the stronger its effect on the tool
response. Corrections for borehole size and mud resistivity are performed
either on-line on the logging truck by means of computer algorithms, or after
the log is recorded by using correction charts (see Section 5.2.2).
In soft or poorly cemented formations, the borehole may be eroded to
a diameter much larger then the bit size by the action of the mud flow.
30 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

This enlargement is called a cave. Caves may increase borehole effect either
smoothly or irregularly with depth.
Because a pressure drop is maintained across the borehole wall, a mud’s
liquid phase (mud filtrate) displaces the movable connate liquid in permeable
formations. Particles in the mud are filtered out and adhere to the borehole
wall to form a mudcake. Filtrate flow diminishes rapidly at first and then
more slowly until it reaches equilibrium [86]. The mudcake formed usually
ranges from 0.1 to 1 inch in thickness. The thin mudcake has little effect
on the response of mandrel tools, such as induction or laterologs. Mudcake
corrections are only needed for pad-type tools, which are applied against the
borehole wall and have shallow depths of investigation.

2.1.2 Coaxial layers; invasion

In permeable formations, the mud filtrate flushes away most of the connate
water and much of any hydrocarbons that may be present in the region close
to the borehole. This flushed zone is referred to as the invaded zone (see
Figure 2.1). Further out from the borehole, the displacement of formation
fluid may become less and less complete, resulting in a transition zone (for
modeling simple invasion, the transition zone is normally ignored and step
contact is assumed between Rxo and Rt ). Saturations in the transition zone
range between those of the mud filtrate and the original formation fluid. The
extent of the invaded and transition zones depends on several parameters:
drilling mud properties, formation porosity and permeability, the pressure
differential and the time since the formation was first drilled [220].
Sometimes in oil and gas-bearing formations, where the mobility of the
hydrocarbons is greater than that of water because of relative permeability
differences, the hydrocarbons move away faster than the interstitial water.
In this case, an annulus with high formation water saturation may be formed
between the invaded zone and the uninvaded formation. Figure 2.2 shows
typical saturation and resistivity profiles for an annulus region. Annuli prob-
ably occur to some degree in most hydrocarbon-bearing formations. Their
influence on log measurements depends on the radial location of the annulus
and the severity of the resistivity contrast. Annuli typically develop near the
borehole shortly after drilling and gradually broaden and migrate outward
until they disappear in time through dispersion [7].
In fractured formations the invasion pattern is usually quite different.
2.1. OVERVIEW OF LOGGING ENVIRONMENTS 31

Figure 2.2: Saturation (a) and resistiv- Figure 2.3: 1D coaxial cylindri-
ity (b) profiles for a representative ex- cal geometry for modeling bore-
ample of annulus invasion. hole and invasion effects.

Unless fractures are very thin, they are generally invaded by bulk mud and
no mudcake is formed [133]. Most shales have extremely low permeabilities,
and it may be assumed that shales are not invaded (occasionally heavy oil-
based mud can cause hydraulic fracturing of shales [18]).
Early 1D analytical codes for modeling borehole and invasion effect as-
sumed coaxial layers with smooth cylindrical boundaries, as shown in Fig-
ure 2.3 [27].
This simplification of the environment sometimes led to optimistic eval-
uations of tool performance. Since the 1980’s, 2D and 3D finite element
and finite difference codes have allowed features such as caves [14] and non-
uniform invasion caused by gravity segregation [105] or permeability anisot-
ropy [18] to be assessed more accurately.
32 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

Figure 2.4: 1D layered formation geometry.

2.1.3 Thin beds (bed boundary discontinuities)

Most reservoir forming rocks were laid down in strata like a layer-cake. The
uniformity of layers is dependent on the conditions present at the time of de-
position. For first-order interpretation purposes, the resistivity within a layer
is assumed to be relatively uniform in all directions (i.e., anisotropy is not
taken into consideration). Boundaries between layers with different physical
characteristics are assumed to be planar and parallel to first approximation.
This familiar layer-cake representation of sedimentary geological structure
is shown in Figure 2.4 [27].
The main property that determines the resistivity of an individual layer
is its porosity, since electrical current only flows through the water saturating
the pore structure. The higher the porosity, the greater the amount of water
that can be present, and therefore the lower the resistivity. The salinity of the
water also contributes to the over-all resistivity, with high salt concentrations
further reducing the resistivity.
Porosity of subsurface layers can vary widely. Carbonates (limestones
and dolomites) and evaporites (salt, anhydrites and gypsum) show prac-
tically zero porosity [220]. Their resistivities are usually in excess of 100
ohm-m.
Shales or clays may contain over 40% water-filled porosity. However,
2.1. OVERVIEW OF LOGGING ENVIRONMENTS 33

individual pores are so small that the rock is impervious to the flow of
fluids. Shale resistivities typically range from 0.5 to 5 ohm-m [86].
Well-consolidated sandstones have porosities between 10 to 15%; uncon-
solidated sands may have 30% or more porosity. If sands are saturated with
salt water, as often occurs in offshore wells, the resistivity may be as low
as 0.2 ohm-m. Oil-bearing sands that are interspersed with shale lamina-
tions (so-called low-resistivity pay) have resistivities averaging around 1 to
2 ohm-m [59]. “Normal” pay sands have resistivities ranging from 2 to over
1000 ohm-m.
Since tool response to a bed of interest can be strongly affected by ad-
jacent layers, thin bed modeling has historically played an important role
in both tool design and log interpretation (for early tools of the 1950’s, a 6
foot bed was considered thin). The geometry shown in Figure 2.4 is assumed
by 1D analytical codes that model induction response to thin beds with the
tool logging perpendicular to bed boundaries (vertical wells).
1D codes have served well for the Dual Induction tool, which was de-
signed to have minimal borehole effect and is often run in oil-based muds
where invasion is shallow or nonexistent (borehole effect for laterologs is of-
ten large and therefore cannot be neglected). 1D layered media codes were
used to evaluate the ability of early tools to resolve thin beds and to generate
shoulder correction charts (described in Chapter 5.) In the 1980’s, 1D thin
bed modeling, supplemented by 2D modeling of beds with invasion, was
used to design Phasor processing which extended Dual Induction vertical
resolution down to 2 feet [221].

2.1.4 Invaded thin beds

The 2D geometry for modeling thin beds with invasion, shown in Fig-
ure 2.5 [27], is very much a combination of the 1D coaxial cylindrical ge-
ometry (Figure 2.3) and the 1D layered formation geometry (Figure 2.4).
Bed boundaries are assumed parallel to each other and perpendicular to the
borehole axis (z). Radial boundaries (borehole and invasion, if it exists) are
perfectly cylindrical and centered around the borehole axis. This is the ge-
ometry commonly assumed by 2D finite element, finite difference and hybrid
codes for modeling induction and laterolog response in vertical wells.
Invasion that arises in thin beds normally occurs in the more porous and
permeable sandstones. Shales and “tight” carbonates usually do not invade
34 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

Figure 2.5: 2D layered formation geometry with borehole and invasion.

and act as permeability barriers to prevent interaction between invasion in


different beds.

2.1.5 Dipping beds

For the purpose of modeling tool response, dipping beds are considered to be
any beds whose boundaries are not perpendicular to the tool axis. As such,
dip has three causes: (1) geologic tilting of the formation, (2) deviation of
the wellbore from vertical and (3) a combination of formation tilt and well
deviation.
The effect of dip on resistivity tool response was virtually ignored until
the mid-1980’s when horizontal drilling became common practice. Before
that time, formation dips encountered were usually less than 30◦ , and were
shown to have little effect on induction [32] or laterolog [65] response. How-
ever, the 60◦ to 90◦ dips encountered in horizontal drilling often rendered
resistivity logs uninterpretable.
Figure 2.6 [49] illustrates the reason for this complication. In vertical
wells, the volume of investigation of a tool is normally within the bed where
it resides. However, in horizontal wells, the volume of investigation may
extend over several beds.
2.1. OVERVIEW OF LOGGING ENVIRONMENTS 35

Figure 2.6: Induction response in vertical (A) and nearly horizontal (B)
sections of a deviated well showing how dip causes the region probed by the
tool to cut across several beds.

Fast analytical codes for modeling induction response in dipping beds


(without borehole effect) were developed in the 1980’s [32, 143]. 3D finite
element codes are required for modeling laterolog response in dipping beds,
since borehole effect cannot be ignored.

2.1.6 3D geometries; horizontal wells

The drilling of horizontal wells has accelerated the development of 3D finite


element and finite differences codes. Indeed, if invasion is present in devi-
ated wells, it is practically impossible to interpret induction and laterolog
response without 3D modeling. Two examples of the type of complex inva-
sion geometries that can arise in deviated wells are illustrated in Figure 2.7
and Figure 2.8.
36 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

Figure 2.8: 3D horizontal well with


Figure 2.7: 3D deviated well with non- the wellbore passing below an im-
cylindrical invasion caused by gravity permeable cap shale; there is annu-
segregation. lus invasion in the pay sand below.

Figure 2.7 [27] shows a deviated well, where gravity segregation has
caused invasion to spread out above an impermeable bed. Figure 2.8 [18]
shows annulus invasion which is truncated above a horizontal borehole by
a cap shale. In addition to solving specific interpretation problems such as
these, 3D modeling is also prompting research in the areas of tool design, log
inversion and invasion physics by identifying deficiencies in existing methods.

2.1.7 Anisotropy in layered media; laminated formations

Physical characteristics within a bed (i.e., resistivity, permeability) are usu-


ally relatively uniform in all radial directions parallel to the plane of depo-
sition and slightly different perpendicular to that plane. This gives rise to
some degree of TI anisotropy [220] (transversely isotropic anisotropy, which
denotes having the same resistivity in every direction in the horizontal bed-
ding plane, but a different resistivity normal to it). On the macroscopic
scale (between grain-size and bed-size) there are two main types of deposi-
tion that can cause anisotropy. They are: (1) alternating thin sand–shale
laminae, and (2) alternating fine and coarse microlayering.
Sand–shale laminae are composed of fairly conductive shales and sands
that can be quite resistive if they are hydrocarbon saturated. The anisotropy
resulting from this combination is one of the primary causes of low-resistivity
pay, where hydrocarbons are recovered from zones that look like either shales
2.1. OVERVIEW OF LOGGING ENVIRONMENTS 37

Figure 2.9: Whole-core photograph from a well in the Gulf of Mexico showing
the relative distribution of shale (dark) and sand laminations. (Note that
the length of each of the three sections of core is slightly over one foot.)

or wet sands. The inherent conductivity of the shale contributes to the low
resistivity by reducing Rh read by resistivity tools in vertical wells (Equation
(1.2)). Interpretation in deviated wells is further complicated because tools
respond to both Rv and Rh as a function of deviation (Equation (1.6)). Thus
logs from a vertical well and a deviated well in the same reservoir will give
different values of “Rt ”. Figure 2.9 [189] illustrates the relative size of sand
and shale layers in a representative low resistivity pay reservoir. Individual
layer thicknesses typically range from a fraction of an inch to several inches.
Electrical and density image logs can be used to improve the interpre-
tation of sand–shale anisotropy in deviated wells. Image logs provide an
estimation of sand and shale layer thicknesses and apparent dip. This infor-
mation, along with resistivity from 2-MHz logging while drilling logs, can be
used to derive Rh and Rv and to isolate the resistivity of the hydrocarbon-
38 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

Figure 2.10: Photograph of a fluvial deposit of the Colorado river showing


fine and course microlayering with crossbedding formed by ripples. (Note
the pencil near the top of the photograph indicating scale.)

bearing sand layers from the shale resistivity, giving a more accurate deter-
mination of oil in place than traditional shaly sand methods [246].
Alternating fine and coarse microlayering can cause anisotropy in per-
fectly clean sands with no shale content. If both hydrocarbons and water are
present, the water saturation of the fine-grained layers will be higher than
that of the coarse-grained layers, leading to alternating resistive and conduc-
tive layers with high anisotropy [159]. This type of anisotropy is often asso-
ciated with crossbedding, that is, wind or water-deposited strata arranged
at different angles relative to the main bedding plane. In some cases there
may be thin cemented sandstone layers separating crossbeds [264], which
further complicates interpretation. Figure 2.10 [213] illustrates alternating
fine and course microlayering in a crossbedded dune.
It is also possible for fine and coarse sand microlayering to exist in com-
bination with shaly layers or shaly sands. In general, pronounced electrical
anisotropy in porous sediments is a good indicator of hydrocarbon pay.

2.2 Description of logging tool configurations; mandrel


tools vs. dipole approximations
Today, well logging is completely controlled by a computer located on a
logging truck. Logging data are recorded and processed by the computer and
2.2. DESCRIPTION OF LOGGING TOOL CONFIGURATIONS 39

output to either paper or magnetic media for additional processing offsite.


The logging tools themselves are composed of two main components: (1)
a sonde containing the sensors used for making measurements (electrodes
for laterologs or coils for induction tools), and (2) a cartridge containing
electronics that power the sensors, process the measured signals and transmit
the data uphole [220]. Most logging tools are combinable, that is, the sondes
and cartridges of several tools can be connected together in order to make
multiple measurements on a single trip into the borehole. The logging string
is typically 3.5 to 4 inches in diameter and 20 to 50 feet long [86].
In wireline logging, the tool is suspended from the end of a cable and
lowered into the borehole by means of a powered winch-drum. The cable
both supplies power to the tool and digitally transmits recorded data uphole
to the truck computer. In logging while drilling (LWD), tools are mounted
on the drill string and powered by batteries. Data is either transmitted to the
surface in real time by pulsing the mud or stored in memory within the tool
for downloading when the bit is pulled to the surface [8]. LWD tools have
the advantage of acquiring early-time data that is relatively uncorrupted
by invasion and can be used for steering the bit. However, the slow real-
time data transmission rate of LWD tools (12 bits per second for mud-
pulse compared to 500 kilobits per second for wireline) prohibits the use of
sophisticated array tools for LWD.
The transmitters and receivers of induction-type tools consist of coils
wound coaxially around a mandrel, as shown in Figure 2.11 (a). The mandrel
of present-day tools is made of steel; early tools had a fiberglass mandrel.
The entire tool is enclosed in an epoxy-composite housing. The induction
transmitter coil is driven by a high-frequency alternating current (in the kHz
to several MHz range) of constant intensity which creates a primary magnetic
field around the tool. This magnetic field induces currents in the formation
which flow in circular loops centered around the tool axis. These current
loops in turn set up a secondary magnetic field which induces a voltage in
a receiver coil. This voltage is approximately proportional to the formation
conductivity. Commercial tools consist of arrays of transmitters and receiver
coils which focus induction currents in regions of interest (i.e., to make the
depth of investigation deeper or shallower.) Coil strengths are weighted by
adjusting the number of turns and direction of winding (induction focusing
is described in greater detail in Chapter 3).
Although induction coils are wound on a mandrel that is several inches
in diameter, calculations of magnetic fields generated by finite-size coils both
40 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

(a) (b) (c)

Figure 2.11: Three source representations commonly used for modeling re-
sistivity tool response shown in a borehole: (a) loop around a mandrel, (b)
thin ring, (c) point dipole.

with and without a mandrel (Figure 2.11 (a) and (b)) show that coils may
be replaced by idealized point dipoles (Figure 2.11 (c)) for modeling most
cases of practical interest [33]. One notable exception is eccentricity effect in
resistive formations with conductive boreholes [123, 178, 84]. In most other
cases, tool effects are small in comparison to effects from the formation
(such as anisotropy, shoulder-bed effect and dip). Therefore the extra time
required for numerical analysis and modeling of finite-size coils and a mandrel
is not justified.
Borehole effects for induction tools in general are also small. Conse-
quently the borehole is often omitted from induction modeling in order to
further speed up calculations. 3D modeling has shown [18] that the ar-
ray induction borehole corrections algorithm [129] removes borehole effect
so accurately that modeling tool response without a borehole is effectively
equivalent to the field performance of the borehole-corrected tool. This is
true even for the shortest spacings (under two feet).
Laterologs, however, cannot be accurately modeled as point sources [177],
and the mandrel and borehole are always included in laterolog response cal-
culations. Laterolog tools inject current into the formation from conductive
2.3. ELECTROMAGNETIC FIELD EQUATIONS AND NOTATION 41

metallic electrodes which are directly in contact with the borehole mud. In
a homogeneous isotropic medium, the amount of voltage required to drive
a unit current between two electrodes is approximately proportional to the
resistance of the formation, as indicated by Equation (1.1). Currents radiate
outward from a source in straight lines, and surfaces of constant potential are
spheres. However when a borehole is present, laterolog current lines bend
as they cross the borehole wall, with the degree of bending being a func-
tion resistivity contrast between the mud and the beds between the current
source and return. Laterologs are often run in salty muds where the Rt /Rm
contrast is as high as 10,000. In cases such as these, borehole effect can be
large and highly nonlinear and cannot be neglected. In order to minimize
the effect of the borehole and shoulder beds, additional electrodes are in-
troduced to focus currents in regions of interest (various types of laterolog
focusing are described in Chapter 3.)

2.3 Electromagnetic field equations and notation

The response of all electrical logging tools is calculated from numerical or


analytical solutions of Maxwell’s equations with the appropriate source and
boundary conditions. Maxwell’s equations describe the behavior of electro-
magnetic fields in space and time. Position in space is specified by (x, y, z)
coordinates in a right-handed Cartesian reference frame consisting of three
mutually perpendicular base vectors {iix , iy , iz } that are of unit length each.
(All vector quantities will be represented by bold-face symbols.) The posi-
tion of a vector A is the linear combination of A = Axix + Ay iy + Az iz as
shown in Figure 2.12. The Cartesian coordinate system is chosen so that its
x-axis and y-axis are parallel to planar beds in the logging environment (see
Figure 2.1). The borehole axis does not necessarily coincide with the z-axis;
the borehole can be deviated as shown in Figure 2.7.
Electromagnetic quantities considered in this thesis and their units in
the International System of Units (SI) are, in the frequency domain:

E = electric field strength (V/m)


H = magnetic field strength (A/m)
J e = volume density of external (source) electric current (A/m2 )
J = volume density of electric current (A/m2 )
42 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

A Az
iz

iy y
ix Ax

Ay
x

Figure 2.12: Cartesian coordinate system.

K e = volume density of external (source) magnetic current (V/m2 )


K m = volume density of material magnetic current (V/m2 )
D = electric flux density (C/m2 )
B = magnetic flux density (T)
ρ = volume density of electric charge (C/m3 )
j S = surface current density (A/m)
σ S = surface charge density (C/m2 )
σ = conductivity (S/m)
 = dielectric permittivity = r 0 (r is relative permittivity) (F/m)
0 = dielectric permittivity of free space = 8.8541878 × 10−12 (F/m)
µ = magnetic permeability = µr µ0 (µr is relative permeability) (H/m)
µ0 = magnetic permeability of free space = 4π × 10−7 (H/m)

Corresponding electromagnetic quantities in the time domain are ex-


pressed in script notation.
2.4. TIME DOMAIN EQUATIONS 43

2.4 Electromagnetic field equations in the time domain

Maxwell’s equations describe the manner in which an electric current pro-


duces a magnetic field, and in which a magnet can produce an electric cur-
rent, as well as how both electric charges and magnetic poles can set up fields
consisting of lines of force. Maxwell showed that electric or magnetic fields
could not be considered in isolation. The two are present together, giving
rise to a single electromagnetic field which propagates outward in all direc-
tions. The result is the radiation of electromagnetic waves with frequencies
equal to that in which the electromagnetic field oscillates [40].
Maxwell’s equations in the time domain are given by

∇ × H + ∂t D + J = −J
−∇ J e, (2.1)
Ke ,
∇ × E + ∂t B = −K (2.2)

in which the hypothetical magnetic current density, K e , is introduced for


convenience.
Applying the divergence operator, ∇·, to both sides of Equations (2.1)
and (2.2) leads to the compatibility relations

∂t ∇ · D + ∇ · J = −∇ ∇ · J e, (2.3)
∇ · Ke.
∂t ∇ · B = −∇ (2.4)

Historically, the volume density of electric charge is introduced as ρ = ∇ · D .


Equation (2.1) without the ∂t D displacement current term (added by
Maxwell) is related to Ampère’s circuit law. Equation (2.2) is related to
Faraday’s induction law.
The constitutive relations describe the properties of media and provide
additional information to solve for electromagnetic field vectors. Restriction
to instantaneous, anisotropic media give the constitutive relations

D =  · E, (2.5)
B = µ · H, (2.6)
J = σ · E, (2.7)

where, for example, J = σ · E means

Ji = σij Ej (2.8)
44 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

in subscript notation, which is short for



3
Ji = σij Ej . (2.9)
j=1

In the case of isotropic media,

ij =  δij , (2.10)


µij = µ δij , (2.11)
σij = σ δij , (2.12)

where δij is the symmetric unit (Kronecker) tensor of rank two (δij = 1
when subscripts are equal, and δij = 0 when subscripts are different). The
constitutive are relations are

D = E, (2.13)
B = µ H, (2.14)
J = σ E. (2.15)

2.5 Electromagnetic field equations in the frequency do-


main

In the time-harmonic or steady-state case, it is assumed that all fields depend


sinusoidally on time. Thus

E , H , D , B , J , J e , K e }(x
{E E , H , D , B , J , J e , K e }(x
x, t) = Re[{E x, ω) exp(−iωt)],
(2.16)
where ω is the angular frequency (2π·frequency). Maxwell’s equations in the
frequency domain are then found as

∇ × H − iωD
−∇ J e,
D + J = −J (2.17)
∇ × E − iωB K e.
B = −K (2.18)

The compatibility relations lead to

∇ · D = ρ, (2.19)
∇ · B = 0, (2.20)

when there is no external magnetic current.


2.6. ANISOTROPIC MEDIA; THE CONDUCTIVITY TENSOR 45

The constitutive relations for instantaneous, anisotropic media are then


given by
D =  · E or Di = ij Ej , (2.21)
B = µ · H or Bi = µij Hj , (2.22)
J = σ · E or Ji = σij Ej . (2.23)
The constitutive relations for isotropic media are
D =  E, (2.24)
B = µ H, (2.25)
J = σ E. (2.26)

2.6 Anisotropic media; the conductivity tensor repre-


sentation

The electromagnetic properties of anisotropic media are characterized by


three real, symmetric tensors of rank two: σij , ij and µij . These tensors
can be represented mathematically by real, symmetric 3-by-3 matrices, and
geometrically by second-degree surfaces. For example, for the conductivity
tensor, this geometric surface is an ellipsoid [76]. The directions of the rep-
resentation surface’s major axes are the principal directions of the relevant
tensor. If the reference coordinate axes are parallel to the major axes of the
ellipsoid, the off-diagonal terms in the representation matrix vanish. The
terms on the main diagonal are the principal values (σx , σy , σz ) of the ten-
sor. The geometrical representation of a conductivity tensor as an ellipsoid
in the principal axis system is shown in Figure 2.13 [199].
The highest tensor symmetry is obtained when a material is isotropic.
Then all three principal values are equal. If two of the principal values
are equal but differ from the third, the geometric surface is an ellipsoid
of revolution. This configuration has cylindrical symmetry. A practical
example of this situation is thin-bedded sequences of alternating high and low
resistivity layers that occur in logging environments. This type of anisotropy
is commonly referred to as transversely isotropic or TI anisotropy. In the
principal axes system with TI anisotropy
   
σx 0 0 σh 0 0
σ= 0 σy 0 = 0 σh 0 , (2.27)
0 0 σz 0 0 σv
46 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

1/σ z
1/σ x 1/σ y
y
x

Figure 2.13: Orientation of a conductivity tensor represented as an ellipsoid


in the principal axis system.

where σh is the horizontal conductivity and σv is the vertical conductivity. If


all three principal tensor elements are different, the anisotropy is classified as
biaxial. In the TI anisotropic case, the anisotropy is classified as uniaxial. In
borehole logging configurations, the principal directions (axes) of the various
beds normally do not coincide with the chosen coordinate system.
The c-axis, shown in Figure 2.14, is chosen along the principal conduc-
tivity axis of the uniaxial conducting medium. A unit vector, ck , along the
c-axis has the components

cx = sin θ cos φ (2.28)


cy = sin θ sin φ (2.29)
cz = cos θ (2.30)

with 0 ≤ θ < 2π, 0 ≤ φ < 2π and ck ck = 1. To convert to a chosen


coordinate system, a rotation through the angles θ and φ is performed as
shown in Figure 2.14. Such a conversion is performed, for example, when
the tool axis does not coincide with the principal conductivity axis. This is
the case for the inversion results described in Chapter 6.
The overall rotation in θ and φ can be described [193] in terms of a
rotation matrix, R , given by
 
cos θ cos φ cos θ sin φ − sin θ
R =  − sin φ cos φ 0  (2.31)
sin θ cos φ sin θ sin φ cos θ
2.7. BOUNDARY CONDITIONS 47

iz
cz
c
θ

cy
iy
cx φ

ix

Figure 2.14: Orientation of the c-axis within the Cartesian reference frame.

Using this rotation matrix, the conductivity tensor can be straightfor-


wardly expressed in the Cartesian coordinate system in terms of the elements
σ , of a
of the principal tensor. For TI anisotropy, the conductivity tensor, σ̂
bed in the Cartesian coordinate system is then found to be
σ = R −1σ R ,
σ̂ (2.32)
where R −1 is equal to the transpose of R . For the inversion problem consid-
ered in this thesis, it is assumed that only the conductivity exhibits anisot-
ropy.

2.7 Boundary conditions

Boundary conditions are necessary for relating the electromagnetic field


quantities on either side of an interface between two regions where the con-
stitutive properties differ. Let S denote an interface and assume that S has
everywhere a unique tangential plane. Furthermore, let ν denote the unit
vector along the normal to S such that upon crossing S one passes from
region M1 to region M2 as shown in Figure 2.15.
If there is no surface charge, the boundary equations on S are
ν × H 1 = ν × H 2, (2.33)
ν × E 1 = ν × E 2. (2.34)
48 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

Figure 2.15: Interface between two Figure 2.16: Interface between two
isotropic regions with different elec- regions with M1 electrically impen-
tromagnetic properties. etrable.

If a surface charge exists on S , the boundary equations on S are

ν × H 2 − ν × H 1 = jS , (2.35)
ν × E 2 − ν × E 1 = 0. (2.36)

If region M1 is electrically impenetrable as shown in Figure 2.16, then


E ≡ 0 in M1 . Consequently,

x) × E (x
ν (x x) → 0, (2.37)

when x → S .
An electrically impenetrable region is either perfectly conducting (i.e.,
σ → ∞) or  → ∞. From Maxwell’s equations it also follows that H (x x) = 0
in M1 . As a result,

x) × H (x
ν (x x) → j S (x
x), (2.38)
x) · B (x
ν (x x) → 0, (2.39)

when x → S .
In an isotropic medium, the electric field lines are locally perpendicu-
lar to S in region M2 , while the magnetic field lines are tangential to S .
Equation (2.38) states that the tangential component of the magnetic field
strength has a surface current density j S as a limiting value on S .
2.8. TRANSFORM FOR AXISYMMETRIC CONFIGURATIONS 49

0
y
ρ
φ

Figure 2.17: Transformation from Cartesian to cylindrical coordinates.

2.8 The transform-domain equation for axisymmetric


well-logging configurations

For problems involving the borehole, a cylindrical coordinate system is some-


times employed. Cartesian coordinates are transformed to cylindrical coor-
dinates using the relationships

x = ρ sin φ, (2.40)
y = ρ cos φ, (2.41)
z = z, (2.42)

as shown in Figure 2.17, with 0 ≤ ρ < ∞, 0 ≤ φ < 2π and −∞ < z < ∞.


The divergence operator, ∇· indicates the total outward flux from a point.
The divergence, which in Cartesian coordinates is

∇ · A = ∂x Ax + ∂y Ay + ∂z Az , (2.43)

becomes in cylindrical coordinates

1 1
∇·A = ∂ρ (ρAρ ) + ∂φ Aφ + ∂z Az . (2.44)
ρ ρ
50 CHAPTER 2. ELECTROMAGNETIC RELATIONS FOR LOGGING

The curl operator, ∇× indicates the amount of rotation a field has. The
curl, which in Cartesian coordinates is

∇ × A = ix (∂y Az − ∂z Ay ) + iy (∂z Ax − ∂x Az ) + iz (∂x Ay − ∂y Ax ), (2.45)

becomes in cylindrical coordinates


 
1 1 1
∇ ×A
A = iρ ∂φ Az − ∂z Aφ +iiφ (∂z Aρ − ∂ρ Az ) +iiz ∂ρ (ρAφ ) − ∂φ Aρ .
ρ ρ ρ
(2.46)
In the frequency domain, Maxwell’s equations in cylindrical coordinates
for isotropic media are then found to be

1 
ρ ∂φ Hz − ∂z Hφ + iωDρ − Jρ = J ρ
e


∂z Hρ − ∂ρ Hz + iωDφ − Jφ = J e φ , (2.47)


1 1 
ρ ∂ρ (ρHφ ) − ρ ∂φ Hρ + iωDz − Jz = J z
e


1 
ρ ∂φ Ez − ∂z Eφ − iωBρ = −K ρ
e


∂z Eρ − ∂ρ Ez − iωBφ = −K e φ . (2.48)


1 1 
ρ ∂ρ (ρEφ ) − ρ ∂φ Eρ − iωBz = −K z
e
Chapter 3

Overview of electrical well-logging


measurements

Summary: This chapter provides an overview of the response characteristics of


some of the most commonly used resistivity logging tools. The emphasis is on
developing an understanding of tool physics, since this understanding furnishes
valuable insights for writing efficient modeling and inversion software. Specific tool
parameters used by modeling and inversion codes are given (such as induction coil
locations and turns, and electrode tool focusing conditions). The vertical resolu-
tion and radial depth of investigation of all major induction an laterolog tools are
systematically compared for the first time by analyzing their computed logs in the
same benchmark formation.

3.1 What do “resistivity” tools measure

The success of resistivity logging as a hydrocarbon detection and formation


evaluation technique stems from the fact that the formation electrical re-
sistivity (Rt ) is strongly dependent on the concentration of hydrocarbons,
which are electrical insulators. Electrical conduction takes place via any wa-
ter present in rock pores. The fraction (or percentage) of the rock volume
that is pore space is referred to as its porosity (φ).
An idealized view of a porous, hydrocarbon-bearing rock is shown in Fig-
52 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.1: Idealized view of hydrocarbon-bearing rocks containing oil and


water (left), and oil, gas and water (right).

ure 3.1 [86]. The rock matrix commonly consists of grains of sand, limestone
or dolomite. The pore space between the grains is filled with water, oil, and
perhaps gas. The water exists as a film around the rock grains and also
occupies very fine crevices, forming a continuous, tortuous path through the
rock matrix. Oil occupies the larger pore spaces. If gas is present, it will
occupy the largest pores [86].
Both porosity and water saturation are used to determine the quantity
of hydrocarbons in place. Porosity is measured by nuclear or acoustic tools.
Resistivity tools provide a measurement the water saturation (Sw ), which is
the fraction (or percentage) of the pore space containing water. The remain-
ing fraction of the pore space which contains oil or gas is the hydrocarbon
saturation (Sh ). The fraction of the total formation volume containing hy-
drocarbons is therefore φ Sh or φ (1 − Sw ).
The resistivity of the water in the rock pores (Rw ) also enters into the
interpretation of resistivity logs. Fresh water is fairly resistive. Saline water,
which is often found at depths where hydrocarbons are located, is usually
quite conductive. The value of the water resistivity is generally obtained
from a nearby clean (non-hydrocarbon bearing) sand.
Figure 3.2 [247] shows the dependence of Rt on water saturation and
water resistivity for a formation with 10% porosity and one with 30% poros-
ity. For illustrative purposes, note that in Figure 3.2, in a 10% porosity
formation impregnated with a mixture of hydrocarbons and sea water, Rt
3.1. WHAT DO “RESISTIVITY” TOOLS MEASURE 53

Figure 3.2: Formation resistivity (Rt ) as a function of water saturation (Sw )


and water resistivity (Rw ) for a formation with 10% porosity (a), and 30%
porosity (b).

is approximately 20 ohm-m at 100% water saturation. Rt increases to 80


ohm-m at 50% water saturation, and to 500 ohm-m at 20% water saturation.
(Resistivity data is normally plotted on a logarithmic scale.)
The resistivity of the mud filtrate (Rmf ) which may invade permeable
formations also has an important effect on the measurement Rt , since the
invaded zone resistivity (Rxo ) close to the tool can significantly perturb tool
response. (Figure 3.2 can also be used to estimate Rxo by substituting
Rmf for Rw .) Muds can be made with fresh or salt-saturated water, or an
emulsion of water and oil. Depending upon the resistivity of the uninvaded
zone and the water saturation, two basic types of radial invasion profiles can
be found in practice. These two profiles are shown in Figure 3.3. Profile (a)
generally corresponds to cases where the mud filtrate resistivity is greater
than the formation water resistivity. Profile (b) is typical of formations
drilled with salt-saturated muds. Subsets of these two basic profiles also
exist which include an annulus or transition zone between Rxo and Rt , as
shown in Figures 2.1 and 2.2.
Laterolog (or electrode) tools are optimally suited to salt mud condi-
tions (i.e., Rxo < Rt ) because they require direct contact with a conductive
mud column in order to inject current into the formation. Laterologs are
considered to be DC measurements for practical purposes, although actual
operating frequencies are in the 10’s to 100’s of hertz range in order to elim-
54 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Rxo Rt
Resistivity (relative)

Rm (a) (b)
Rxo

Rt
Rm

rh ri Radius (relative) rh ri Radius (relative)

Figure 3.3: Radial resistivity profiles encountered in formations drilled with


resistive (a), or conductive mud (b). Solid lines denote the actual shape of
the resistivity profile and dashed lines denote the step profile approximation
often used for modeling invasion effect.

inate electrode polarization and interference from natural potentials. When


currents are emitted from an electrode, current lines bend each time they
cross boundaries between regions with different resistivities. It eventually
becomes difficult to evaluate the individual influence of each region because
a change in resistivity affects current lines not only where it occurs, but
also over their entire path [5]. Therefore, laterologs are said to respond to
formation resistances in series.
In contrast, induction tools are more suited to fresh water or oil-based
mud conditions (i.e., Rxo > Rt ). Induction tools operate in the 10’s of
kilohertz range, and the time-varying electromagnetic fields permit circula-
tion of currents beyond resistive regions. In spite of the fact that induction
is referred to as a “resistivity” measurement, the voltage induced in a re-
ceiver coil is actually proportional to formation conductivity. In isotropic
formations, induction currents are circular and coaxial with the sonde. This
results in parallel current lines which remain within a medium of uniform
conductivity (at least for vertical wells and horizontal beds). Furthermore,
as long as frequency and conductivity are not too high, current loops have
negligible mutual interaction, and the effect of each region may be considered
separately [5]. Therefore, induction tools are said to respond to formation
conductivity in parallel.
This remainder of this chapter examines the response characteristics of
specific resistivity tools. A historical perspective is taken, showing how lim-
3.2. INDUCTION TOOLS 55

First induction log recorded


1950
5FF27 - First focused tool
5FF40 - Deeper investigation
6FF40 - Standard deep induction

1960
DIL - Dual Induction (ILD and ILM) with LL8

DIT-B - Improved ILM


1970
DIT-D - LL8 replaced by SFL

EPT - 1.1 GHz dielectric tool on a pad


DPT - 25 MHz dielectric & conductivity tool
1980

DIT-E - X-signal and Phasor processing

CDR - 2 MHz LWD tool


1990 AIT - Array Induction Imager tool

ARC5 - 2 MHz array tool for LWD

2000

Figure 3.4: Time-line of Schlumberger induction tool development.

itations of early tools were used as opportunities for innovation in designing


their replacements. Because modeling has always played a major role in
the development of new resistivity devices, modeled response functions and
simulated logs are used to characterize each tool.

3.2 Induction tools

The first drilling muds used were water-based. However, early water-based
muds could invade formations to depths of several feet or more, which
severely decreased hydrocarbon production rates. (This deep invasion still
occurs with current inexpensive water-based muds.) Oil-based muds became
a popular substitute in the late 1930’s because they invaded formations to
depths of several inches or less. Only the difficulty of handling oil-based muds
and their high cost kept them from replacing water-based muds entirely.
Because oil-based muds are extremely resistive, they caused major prob-
56 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

lems for the existing normal and lateral tools which needed a conductive
mud column to provide an electrical connection between electrodes and the
formation (wire “scratcher” brushes were sometimes used with limited suc-
cess [5]). Eddy current measurements were considered as a solution to the
oil-based mud problem, since patents for their use in surface prospecting
dated back to the early 1900’s. However, the technology required to develop
a borehole logging tool was nonexistent in the 1930’s.
The first practical induction logging technique was invented by H.G. Doll
in the mid-1940’s [88, 93, 92]. Doll derived the borehole logging tool from a
jeep-mounted mine detector that he developed for the U.S War Department
during World War II. The first induction log was recorded on May 3, 1946 in
a Humble well near Tyler, Texas [5]. Although induction measurements were
originally intended as a replacement for electrode tools in oil-based muds,
the induction tool has come to dominate the resistivity market because it
makes an accurate measurement of formation resistivity over a wide range
of drilling environments, including moderately conductive muds.
Figure 3.4 shows a time-line summarizing milestones in induction tool
development. Individual tools will be described in detail in the remainder
of this section.

3.2.1 Two-coil sonde response

Commercial induction tools consist of multiple coil arrays designed to op-


timize vertical resolution and depth of investigation. However, in order to
illustrate induction tool fundamentals, it is instructive to first examine the
basic building block of multiple coil arrays, the two-coil sonde.
A two-coil sonde consists of a transmitter and receiver mounted coaxially
on a mandrel, as shown in Figure 3.5. Typical coil separations range from
one to ten feet apart. In practice, each coil can consist of from several to a
hundred or more turns, with the exact number of turns determined by design
considerations. The operating frequency of commercial induction tools is in
the tens of kilohertz range, with 20 kHz being the most commonly used
frequency prior to 1990.
The induction transmitter coil is driven by a constant-amplitude sinu-
soidal current with a time dependence of e−iωt which creates a primary mag-
netic field around the tool. The field patterns are illustrated in Figure 3.5.
The primary magnetic field causes eddy currents to flow in circular loops
3.2. INDUCTION TOOLS 57

rR
ρ
L

rT

Figure 3.5: Basic two-coil induction sonde showing electromagnetic field


patterns and the coordinate system used for computing response functions.

(often called “ground loops”) centered around the borehole axis. The eddy
currents are proportional to the formation conductivity, and they in turn
generate a secondary magnetic field, which induces an alternating voltage in
the receiver coil.
Early induction logs were displays of the real part of this induced voltage,
commonly called the R-signal, or resistive signal, plotted on a resistivity
scale. The first logs were plots of raw data, while modern tools use both
electronics and software to “correct” the R-signal for various environmental
effects prior to display. At low conductivities, the R-signal is 180◦ out of
phase with the transmitter current (the eddy current lags the transmitter
current by 90◦ and the induced voltage lags the formation current by an
additional 90◦ [203]). The remaining imaginary part of the signal, which
is 90◦ out of phase with the transmitter current, is called the X-signal, or
reactive signal. The X-signal consists of signals from the formation related
58 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

to overall conductivity level, as well as the signal resulting from the direct
mutual inductance between the transmitter and receiver coils.
At 20 kHz the X-signal is usually much larger than the R-signal. Conse-
quently, Doll had to develop electronic circuitry that could precisely differ-
entiate the R-signal from the X-signal in order to obtain accurate apparent
resistivity measurements. He also found it necessary to introduce an auxil-
iary transmitter “bucking coil” connected in series with the main transmitter
to cancel out the sizable mutual inductance portion of the X-signal [93, 92].
In addition, the depth of investigation and vertical resolution of a com-
mercial tool had to be maintained within desired limits. These constraints
complicated tool design, and Doll introduced geometrical factor theory as
a computational aid in optimizing coil configurations. Because geometrical
factor theory is necessary for analyzing induction logging measurements, it
is summarized below. It will be treated in greater detail in Chapter 4.
In essence, geometrical factor theory defines a response function that
describes the amount of signal coming from each part of the formation.
Doll reasoned that as a first-order approximation, the fields generated by a
transmitter in a wellbore are essentially the same as in a vacuum. Therefore,
the voltage at a receiver is the sum of the contributions from an infinite
number of eddy current loops. Using the Biot-Savart law, Doll defined [88]
the contribution of a single loop having a unit cross-sectional area to the
total conductivity signal as

σLoop = gD (ρ, z) σ(ρ, z) dρ dz, (3.1)

where gD is the “Doll” geometrical factor or relative weight, σ is the con-


ductivity of the formation within the loop, and ρ and z are the radial and
vertical distances as shown in Figure 3.5. For the zero conductivity limit

L ρ3
gD (ρ, z) = 3 , (3.2)
2 rT3 rR

where L is the spacing between the transmitter and receiver coils, and rT
and rR are the vector distances from the formation loop to the transmitter
and receiver respectively, as indicated in Figure 3.5. The total real part of
the apparent conductivity signal, σR , is then given by
 ∞  ∞
σR = gD (ρ, z) σ(ρ, z) dρ dz. (3.3)
−∞ 0
3.2. INDUCTION TOOLS 59

Figure 3.6: Low conductivity Born response function for a two-coil sonde
with a coil separation of 40 inches.

However, Doll’s geometrical factor theory is valid only at the zero con-
ductivity limit. In the early 1980’s, several theoreticians generalized geomet-
rical factor theory to finite conductivities [121, 192, 251, 275]. The method
of Moran [192] is currently the most widely used in induction logging. It
considers a homogeneous formation of conductivity σ with a loop of the for-
mation at conductivity σ + δσ. Because this solution is analogous to the
Born approximation in quantum mechanics and involves a single scattering
in the same manner, it is often called the Born response function.
Starting from a homogeneous formation of conductivity σ instead of a
vacuum, Moran derived an expression [192] for the complex Born response
function, gB (ρ, z), which is given by

gB (ρ, z, σ) = gD (1 − ikrT )eikrT (1 − ikrR )eikrR , (3.4)


60 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

where gD , is the Doll response function from Equation (3.2), rT and rR are
defined as above, and k is the propagation constant (k 2 = iωµσ, neglecting
displacement current at 20 kHz). The measured complex conductivity signal,
σR +i σX , is obtained by integrating over the entire space using the expression
 ∞  ∞
σ R + i σX = gB (ρ, z, σ) σ(ρ, z) dρ dz. (3.5)
−∞ 0

There is a more detailed discussion of Born response functions in Section 4.1.4.


Several other useful response functions can be computed using Equa-
tion (3.4) as a basis. Sonde response to thin formation layers is given by the
vertical response function, gz , which is defined as
 ∞
gz (z, σ) = gB (ρ, z, σ) dρ. (3.6)
0

Response to thin cylindrical shells of formation is given by the radial response


function, gρ , which is defined as
 ∞
gρ (ρ, σ) = gB (ρ, z, σ) dz. (3.7)
−∞

Experience has shown that it is more informative to study the integrated


radial response function, which gives the contribution from a cylindrical
volume of formation (such as an invaded zone) rather than just a thin shell.
The integrated radial response function, Gρ is defined as
 ρ ∞
Gρ (ρ, σ) = gB (ρ, z, σ) dz dρ. (3.8)
0 −∞

Figure 3.6 shows the two-dimensional Born response function at low con-
ductivity for a two-coil sonde with L = 40 inches. The height of the function
at any point is the relative weight given to the loop of formation at that lo-
cation.
Figure 3.7 shows the real and imaginary parts of the vertical response
function for a two-coil sonde. Figure 3.8 shows the radial response function,
and Figure 3.9 shows the integrated radial response function. Curves are
computed at several conductivity values for a coil separation, L, of 40 inches.
Distances have all been normalized to L. The zero conductivity response
is independent of L. However, the finite conductivity responses are not
independent of L because of skin effect, which is nonlinear and greater at
larger distances.
3.2. INDUCTION TOOLS 61

Figure 3.7: Normalized real (top) and imaginary (bottom) parts of the ver-
tical Born response function for a two-coil sonde with a coil separation, L,
of 40 inches.

The term skin effect is commonly used to describe the amplitude re-
duction and phase shift observed in induction signals at high conductivities
and high frequencies. It is borrowed from electrical engineering, where it
describes the phenomenon in which currents in good conductors tend to
concentrate in a thin surface layer or “skin”. This layer has a thickness or
skin depth, δ, defined as

2
δ= (meters). (3.9)
ωµσ
62 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.8: Normalized real (top) and imaginary (bottom) parts of the radial
Born response function for a two-coil sonde with a coil separation, L, of 40
inches.

At the average induction frequency of 20 kHz, skin depths range from 370 feet
at 1 mS/m to 3.7 feet at 10,000 mS/m.
Figure 3.7 shows that most of the signal comes from the portion of the
formation between the coils (−L/2 to L/2). However, there is still a signif-
icant amount of signal contributed from the area outside of the coils. The
imaginary response in Figure 3.7 (and also in Figure 3.8 and Figure 3.9) is
nonexistent at zero conductivity and increases as the conductivity increases.
Some modern signal processing methods use the relationship between the
3.2. INDUCTION TOOLS 63

Figure 3.9: Normalized real (top) and imaginary (bottom) parts of the in-
tegrated radial Born response function for a two-coil sonde with a coil sep-
aration, L, of 40 inches.

X-signal and conductivity level to correct the R-signal for skin effect.
The real radial and integrated radial response functions in Figure 3.8
and Figure 3.9 show that over half of the signal comes from the formation
within a radial distance of L. Note that the signal drops off more slowly in
the radial direction than in the vertical direction.
The depth of investigation of an induction tool is usually defined as
the mid-point (or 50%) of the integrated radial response function, which
is indicated by the dotted line in Figure 3.9 (top). Note that the radial
64 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

distance of the mid-point decreases with increasing conductivity. The depth


of investigation gets shallower because the electromagnetic fields become
increasingly attenuated at higher conductivities due to skin effect. These
response functions show graphically where the induction signal is coming
from, and are an invaluable aid in tool design, log interpretation and signal
processing.

3.2.2 Early induction tools; “focused” sondes

The preceding overview of geometrical factor theory and skin effect has
touched on some of the constraints that must be taken into account in order
to design a commercial tool. Before examining tool design and interpreta-
tion in greater detail, it is helpful to first summarize the basic requirements
for a practical induction measurement. The primary requirements are:

1. The depth of investigation of a tool should be deep enough to accu-


rately estimate the formation resistivity beyond any invasion. Invasion
radii are typically less than 4 feet, but may occasionally extend to 8
feet or more.
2. The apparent resistivity reading for a bed of interest should not be
significantly affected by adjacent shoulder beds. A bed thickness of 5
feet was selected as a suitable target for the vertical resolution of early
induction tools, but today it has become desirable to achieve a vertical
resolution of 2 feet or less.
3. There should be little borehole effect in holes ranging from 8 to 16
inches in diameter, or if borehole effect is non-negligible, the correction
procedure should be extremely simple.
4. A tool should operate at a frequency high enough to generate a low-
noise signal, but not so high as to be significantly influenced by skin
effect.
5. The mutual inductance signal should not be so large that it obscures
the apparent conductivity signal coming from the formation.

The first commercial induction tool was the 5FF27, introduced in the
Texas–Louisiana Gulf Coast in 1952. “5” refers to the number of coils, “FF”
indicates fixed focusing in both radial and vertical directions [203], and “27”
is the spacing in inches between the main transmitter and receiver coils. The
5FF27 coil configuration is listed in Table 3.1.
3.2. INDUCTION TOOLS 65

Turns Position(in.)
Transmitters: 42. 13.5
-25. 40.5
Receivers: 43. -13.5
-24. -40.5
-3. 1.75

Table 3.1: 5FF27 coil configuration.

For multi-coil sondes, the total tool response is the normalized summa-
tion of the individual two-coil responses, weighted by the appropriate coil
strengths and spacings
 Ti Rj σa i,j
i,j Li,j
σa T otal = , (3.10)
 Ti Rj
i,j Li,j

where T and R are transmitter and receiver turns, respectively, L is the


spacing between a transmitter–receiver coil pair, and σa is the apparent
conductivity signal. The normalization factor in the denominator of the
above equation is often referred to as the sensitivity of the sonde. The
sensitivity is a meaningful quantity in itself, since if it is too low, the signal
level of a tool may be so small that the measurement is impractical.
The introduction of auxiliary coils on a basic two-coil sonde in order to
have as large a portion of the signal as possible coming from a particular
region of interest is referred to as focusing. Focusing is the superposition
of two-coil responses to either subtract signals from unwanted regions of
the formation (such as the borehole, invaded zones and adjacent shoulder
beds) or to add signals from desired regions (such as hydrocarbon-bearing
thin beds). For induction tools, focusing is accomplished by varying the coil
spacings, number of turns and direction of windings. All transmitter coils are
connected in series, as are all receiver coils. The frequency and transmitter
current remain constant for a given induction sonde (unlike laterolog tools,
where currents are adjusted dynamically while logging).
66 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.10: Vertical response functions for 5FF27 at several conductivities.

Designers of early induction tools were faced with a choice of two basic
operating configurations:

1. Gathering multiple data samples downhole, combining these samples


while they are still downhole in a rapid and intelligent manner and
sending one composite data sample uphole,
2. Gathering multiple data samples downhole and sending them all uphole
separately for post-processing and analysis.

The designers of the induction sondes of the 1950’s chose the first configu-
ration by default because the wireline was not capable of transferring large
amounts of data in real time. However with today’s digital telemetry, up-
hole post-processing and analysis have become a reality, leading to the array
induction tools now run by all major service companies.
The cancellation of first order effects downhole by focusing allowed de-
signers of early induction tools to obtain a great deal of information with
a minimum amount of data transfer to the surface. Focusing also extended
the range of accuracy of departure curves used to estimate the true forma-
tion resistivity Rt from the log reading Ra (the use of departure curves is
described in Chapter 5).
The 5FF27 was designed to have low skin effect (necessary in the high-
conductivity Gulf Coast environment), which also implies a shallow depth
3.2. INDUCTION TOOLS 67

Figure 3.11: Integrated radial response functions for 5FF27 at several con-
ductivities.

of investigation. Figure 3.10 shows vertical response functions, and and Fig-
ure 3.11 shows integrated radial response functions for 5FF27. The responses
in both figures change little with conductivity, indicating that there is only
a small amount of skin effect. A comparison of Figure 3.11 with Figure 3.9
shows that the 5FF27 is significantly shallower than the 40-inch two-coil
sonde.
Figure 3.12 shows a computed 5FF27 log in a theoretical formation with
both invaded and noninvaded beds. Also shown in Figure 3.12 is a com-
puted log for the short normal (or 16-inch normal), an electrode tool that
measures resistivity by injecting current into the formation and measuring
the voltage drop 16 inches from the current source (see Section 3.4 for a
more detailed description of normal tools). The short normal was often run
with early induction tools. This combination is referred to as an induction
electrical survey, or “IES.” Separation between induction and normal curves
was an indication of invasion. Typical curve separations caused by invasion
can be observed in the uppermost and lowermost 10-foot invaded beds in
Figure 3.12. The invasion radius is indicated in the left track. Rxo and Rt
are indicated in the right track by dashed and thin solid lines, respectively.
Note that the 5FF27 and short normal curves also separate in the un-
invaded beds, with the 5FF27 reading closer to Rt than the short normal.
Although 5FF27 reads Rt in the center of the uninvaded 10-foot beds, the
68 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.12: 5FF27 log.

3-foot beds are too thin for the tool to resolve (see also the vertical response
functions in Figure 3.10). Ambiguity in the visual interpretation of curve
separations caused by invasion and vertical resolution has provided ongoing
motivation for research in the areas of signal processing, tool design, 3D
modeling and inversion.
Note also in Figure 3.12 that both tools read closer to Rt in the conduc-
tive 3-foot bed than in the resistive 3-foot bed. In addition, the apparent
thickness of the conductive 3-foot bed indicated by both logs is greater than
that of the resistive 3-foot bed. This is caused by currents flow preferentially
in the more conductive medium. Conductive beds between resistive shoul-
ders are sometimes referred to as a squeeze configuration because currents
appear to be squeezed into the bed (see also Figure 3.84 in Section 3.4 on
electrode tools). Conversely, resistive beds between conductive shoulders are
3.2. INDUCTION TOOLS 69

Figure 3.13: Vertical response functions for 5FF40 at several conductivities.

referred to as an anti-squeeze configuration.


Because the geometry modeled in Figure 3.12 graphically illustrates dif-
ferences in vertical resolution and depth of investigation, it will be used as
a benchmark formation to compare the responses of all resistivity tools de-
scribed subsequently in this chapter. Although early logs were sometimes
plotted on linear resistivity or conductivity scales, the conventional logarith-
mic resistivity scale is used throughout this thesis to maintain continuity
over the history of resistivity tools and to facilitate comparison.

Turns Position(in.)
Transmitters: 57. 20.
-35. 50.
Receivers: 58. -20.
-34. -50.
-4. 2.5

Table 3.2: 5FF40 coil configuration.

The 5FF40 induction tool was introduced in 1956. As Table 3.2 shows,
it was a scaled-up version of 5FF27. Figure 3.13 shows vertical response
functions, and and Figure 3.14 shows integrated radial response functions
for 5FF40. A comparison of Figure 3.14 with Figure 3.11 shows that 5FF40
70 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.14: Integrated radial response functions for 5FF40 at several con-
ductivities.

was deeper, and still had low skin effect. However, the depth of investigation
was not sufficient for formation conditions in areas of the world outside of
the Gulf Coast, where invasion could be much deeper and resistivities much
higher.
Figure 3.15 shows a 5FF40 log in the same benchmark formation as
5FF27. A comparison of Figure 3.15 with Figure 3.12 shows that the deeper
5FF40 reads closer to Rt than 5FF27 in the uppermost and lowermost in-
vaded beds. The vertical resolution of 5FF40 in the 3-foot beds is slightly
poorer than that of 5FF27, as might be expected for a longer tool.

3.2.3 6FF40 and the Dual Induction tool; a standard is set

After experimenting with these preliminary designs, Schlumberger intro-


duced what was to become the industry-standard induction tool, the 6FF40,
in 1959 [249]. This tool has been licensed and run, with minor variations,
by all logging service companies. The 6FF40 (and its Dual Induction ana-
logue, the deep induction array, ID) remained the industry standard for over
30 years until the introduction of array induction tools in the early 1990’s.
The optimization of the 6FF40 coil configuration will be examined in greater
detail because it illustrates the basic problems encountered in designing in-
duction tools and the use of modeling in their solution.
3.2. INDUCTION TOOLS 71

Figure 3.15: 5FF40 log.

The 6FF40 coil configuration is shown in Table 3.3. In designing the


6FF40, the depth of investigation was optimized using the integrated radial
geometrical factor, and the vertical resolution was optimized using the verti-
cal geometrical factor. A main coil spacing of 40 inches was chosen because
its median depth of investigation is deeper than average invasion and its
vertical resolution is near the target resolution of 5 feet. The 5 foot resolu-
tion refers to a bed that gives a solid deflection, not the thickness at which
the true resistivity can be read. The latter definition, in wider acceptance
today, shows the vertical resolution of the 6FF40 to be about 8 feet [15]. It
should be noted that although 5 to 8 foot vertical resolution was considered
adequate in the 1950’s, 1 to 2 foot resolution is desired in today’s thinly
bedded producing reservoirs.
Because borehole and invasion effects are near-field phenomena, an ad-
72 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Turns Position(in.)
Transmitters: 60. 20.
-15. 10.
-4. -50.
Receivers: 60. -20.
-15. -10.
-4. 50.

Table 3.3: 6FF40 coil configuration.

ditional receiver coil with a smaller spacing and reverse windings was in-
troduced between the main coils to cancel out currents circulating in the
borehole and to reduce the currents in the invaded zone. A transmitter
with the same number of turns and a symmetrical spacing was introduced
concurrently to symmetrize thin bed response. The exact number of turns
on these two auxiliary coils was determined by first selecting values of L
for which cancellation of the radial geometrical factor was desired, and then
solving for the number of turns that would bring about cancellation when all
of the transmitter-receiver contributions were summed according to Equa-
tion (3.10). These coils with opposite windings from the main coils also
act as “bucking coils” which serve to cancel a large portion of the mutual
contribution to the X-signal.
One way to improve vertical resolution is to make a sonde as short as
possible. However, this improved vertical resolution is bought at the expense
of reduced depth of investigation. A second way is to try to cancel out
shoulder bed response by introducing exterior coils. The latter method was
applied in the case of the 6FF40. An outer set of coils, again with reverse
windings from the main coils, was added to subtract contributions to the
total vertical geometrical factor response from the area outside the main
coils. The exact location of these coils was allowed to vary within a fraction
of an inch in order to exactly zero out the remaining mutual signal. The
turns on these outer coils had to be kept at a minimum in order to keep the
effective length of the sonde from becoming too large, and thus deteriorating
the vertical resolution.
The concept of effective length was used in designing early induction
sondes in order to compare multi-coil sondes to equivalent two-coil sondes.
The effective length, Le , weights the contributions of each coil pair by the
3.2. INDUCTION TOOLS 73

Figure 3.16: Low conductivity Born response function for 6FF40, also show-
ing the location of transmitter coils (light) and receiver coils (dark).

appropriate transmitter and receiver turns, and is defined as



Ti Rj
i,j
Le = , (3.11)
 Ti Rj
i,j Li,j

where T , R and L are defined as in Equation (3.10).


The effective length of the 6FF40 is 61 inches, which is significantly larger
than the main coil spacing of 40 inches. Knowing that the effective length
of the 6FF40 is 61 inches helps to explain why the sonde is unable to resolve
beds thinner than 5 feet, and also why it reads much deeper than a 40-inch
two-coil sonde.
Figure 3.16 shows the two-dimensional Born response function at low
74 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.17: Vertical response functions for 6FF40 at several conductivities.

conductivity for the 6FF40 tool. Figure 3.17 shows 6FF40 vertical response
functions, and and Figure 3.18 shows integrated radial response functions.
Large changes in the shape of curves and region of investigation are notice-
able in comparison with any of the older tools. These changes reflect the
much larger depth of investigation of the 6FF40 and the resulting inclusion of
deep regions of the formation volume where currents are significantly phase

Figure 3.18: Integrated radial response functions for 6FF40 at several con-
ductivities.
3.2. INDUCTION TOOLS 75

Figure 3.19: Low conductivity integrated radial response of 6FF40 compared


to that of 5FF40 and a 40 inch two-coil sonde.

shifted (see Equation (3.4)).


The improved vertical resolution of the 6FF40 sonde over a 40-inch two-
coil sonde is can be seen by comparing the vertical geometrical factor curves
in Figure 3.17 and Figure 3.7. Note that negative lobes occur on the 6FF40
curves in the shoulders at the position of the outer coil pairs, and positive
lobes occur where the main coils are located. The two-dimensional response
function shown in Figure 3.16 includes the location of the transmitter and
receiver coils, and better illustrates how features in the response coincide
with the coil positions. The large negative excursions that fall within the
borehole region are caused by unequal cancellations of the individual coil pair
responses. They can give rise to a phenomenon known as cave effect. Cave
effect manifests itself in rugose boreholes filled with conductive mud when
the negative response lobes coincide with mud-filled caves in the borehole
wall to cause spurious oscillations or spikes on logs.
The improvement in depth of investigation of the 6FF40 over 5FF40
and a 40-inch two-coil sonde can be seen by comparing the integrated radial
geometrical factor curves for the three tools, shown in Figure 3.19. The im-
proved borehole response of the 6FF40 is illustrated by the integrated radial
geometrical factor curves at small radii plotted on an expanded scale, which
are shown in Figure 3.20. Note that a portion of the 6FF40 curve is negative
from a radius of 4 to 14 inches. This effect is significantly reduced when the
76 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.20: Low conductivity integrated radial borehole response of 6FF40


compared to that of 5FF40 and a 40 inch two-coil sonde.

tool is run eccentered, because the negative borehole response is averaged at


least in part with the signal coming from the formation. Eccentering is not
a problem for field tools since it is easier to introduce a fixed standoff from
the borehole wall than it is to guarantee a tool’s exact centering.
A scaled-down version of 6FF40, called the 6FF28, was developed in the
late 1960’s for use in slim holes and hostile
√ environments. The spacings
of 6FF28 were reduced by a factor of 2 from those listed in Table 3.3
for 6FF40, giving the tool its name. The 6FF28 operated at a frequency
of 40 kHz. Because the skin effect term in the series approximation for
induction response is L/δ (see Equation (4.35)), the skin depth √
(defined in
Equation (3.9)) remains the same when spacings are reduced by 2 and the
frequency is doubled. Thus the correction algorithms and charts developed
for 6FF40 could be used for 6FF28 without modification.
Although the 6FF40 provided an improved deep measurement, its re-
sponse was still influenced by deep invasion. The Dual Induction tool was
introduced in 1962 [249] in an attempt to quantify the effect of the invaded
zone. The Dual Induction tool (DIT) kept the 6FF40 as the deep measure-
ment (renamed ID). A shallower induction measurement (induction medium,
or IM) was added which used the ID transmitter coils in combination with its
own new receiver configuration. The design of the IM was based on 5FF40.
The IM receiver arrangement was optimized to provide a shallower depth
3.2. INDUCTION TOOLS 77

Turns Position(in.)
Common Transmitters: 105. 20.
-26.2 10.
-7. -49.
-2. 30.4
ID Receivers: 59.6 -20.
-15. -10.
-4. 49.4
IM Receivers: 66.1 54.
-32. 80.
-11.3 39.4
-8. -14.
4. -35.

Table 3.4: Dual Induction tool coil configuration.

of investigation using the same procedure described above for 6FF40. This
tool was called the DIT-A. In 1968, with the introduction of the second-
generation DIT-B, an additional small transmitter coil was added to both
arrays in order to improve the borehole response of IM. However, this coil
does not significantly effect the deeper ID response, which is identical to
6FF40 for all practical purposes. The coil configuration for the DIT-B Dual
Induction tool is shown in Table 3.4.
Figure 3.21 shows the two-dimensional Born response function at low
conductivity for the IM array. Figure 3.22 shows IM vertical response func-
tions, and and Figure 3.23 shows integrated radial response functions. Note
in Figure 3.23 that 50% of the IM signal comes from within a 30 inch radius,
while the corresponding 50% point for ID is approximately 60 inches (see
Figure 3.18).
A shallow measurement provided by a laterolog tool was also included
when the Dual Induction tool was run. The Laterolog-8 (LL8) was used on
early induction tools. It was replaced in the mid 1970’s by the Spherically
Focused Resistivity Log (SFL) [231, 244], which had considerably reduced
borehole response compared to the LL8. (LL8 and SFL are both described
in Section 3.4). The relative depths of investigation of ID, IM and SFL
are illustrated by the integrated radial geometrical factor curves shown in
Figure 3.24.
78 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.21: Low conductivity Born response function for IM.

Figure 3.22: Vertical response functions for IM at several conductivities.


3.2. INDUCTION TOOLS 79

Figure 3.23: Integrated radial response functions for IM at several conduc-


tivities.

Measurements provided by tools with three different depths of investi-


gation allow one to solve for the three primary unknowns of the borehole
environment, namely, the formation and invasion resistivities (Rt and Rxo )
and the invasion diameter (di ). The procedure for doing so is described in
Section 5.2.2 under “tornado charts”.
The integrated radial response function curves in Figure 3.24 can also be
used to estimate tool response to step profile invasion. For example, at the
30-inch invasion radius indicated by dashed lines, the response function or
weight is 0.17 for ID, 0.47 for IM and 0.73 for SFL. Induction tool response is
calculated from the response function, G, and the formation conductivities,
σxo and σt , while laterolog response is computed from the response function,
J, and the formation resistivities, Rxo and Rt using

σa = G σxo + (1 − G) σt , (3.12)
Ra = J Rxo + (1 − J) Rt . (3.13)

The laterolog response function, J, is called a pseudo-geometrical factor


because of it’s extreme dependence on resistivity contrast. Using 2.5 ohm-m
for Rxo and 0.5 ohm-m for Rt (the values in the uppermost invaded bed in
Figure 3.25), one obtains apparent resistivity readings of 0.8 ohm-m for IM
and 0.6 ohm-m for ID from Equation (3.12), and 2 ohm-m for SFL from
Equation (3.13). These are nearly the same apparent resistivities shown in
80 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.24: Integrated radial geometrical factors showing the differences in


depths of investigation between ID, IM and SFL.

Figure 3.25, keeping in mind that the logs in the 10-foot bed are subject to
shoulder bed effect.
DIT response in the benchmark formation is shown in Figure 3.25. Note
the improved vertical resolution of IM in the thin 3-foot beds compared
to 5FF40 (Figure 3.15), and the clear separation of the three DIT curves
in the invaded beds. The induction logs in Figure 3.25 have undergone
the conventional processing of the 1960’s and 70’s, which was performed
using a “panel” analog computer on the logging truck. ID and IM measured
voltages were boosted to compensate for their reduction due to skin effect.
In addition, ID was deconvolved using a three-station windowing filter in
an attempt to reduce shoulder bed effect. Deconvolution and boosting are
described in greater detail in Chapter 5.

3.2.4 Phasor processing and deconvolution

Figure 3.25 also illustrates several of the major problems encountered inter-
preting induction field logs processed using the algorithms of the 1970’s:

- Poor vertical resolution of IM and ID logs, particularly in thin resistive


beds, as shown in the bed between 80 and 83 feet.
- Separation between ID, IM and SFL curves in thick uninvaded resistive
3.2. INDUCTION TOOLS 81

Figure 3.25: Dual Induction log.

beds which could be mistaken for invasion, as shown in the bed between
93 and 103 feet.
- Horns and overshoots on ID and IM logs in low resistivity beds, as
seen in the series of beds between 40 and 70 feet.

In extreme cases, some of these parasitic effects on logs were mistaken for
geological features. Although effects such as these are fully predictable from
electromagnetic theory, automatic algorithms to correct for them were un-
successful due to the nonlinearity of the R-signal, which was the only mea-
surement made at that time.
In the early 1980’s, advances in electronics technology and modern sig-
nal processing theory led to the development of Phasor processing [219] for
82 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

improving induction vertical resolution. Phasor processing is a nonlinear


deconvolution technique that corrects induction logs in real time for shoul-
der effect and skin effect over the full range of formation conductivities.
Digital electronics made it possible to use the X-signal to measure the con-
ductivity nonlinearity directly. The remainder of this section gives a brief
overview of Phasor processing. A more detailed description is contained
in [33, 42, 43, 221].
The induction tool makes a measurement that is a blurred average of
formation conductivity. Tool response is dominated by formation layers in-
side the main lobe of the response function. However, layers outside of the
main lobe can also make a significant contribution. This outside response
trails off gradually, particularly at high resistivities, where as much as 100
feet of formation can contribute to the apparent conductivity signal. The
mathematical process that describes how a tool adds the conductivity of each
layer through a response function is called convolution. The blurring pro-
cess is different at different conductivity levels because the response function
changes as a function of conductivity to produce different logs. The response
of an induction tool can be derived by means of convolution by substitut-
ing Equation (3.6) in Equation (3.5). This gives the convolution integral
(expressed in terms of Phasor processing parameters)
 ∞
σa (z) = gv (z − z  ) σ(z  ) dz  , (3.14)
−∞

where σa is the apparent resistivity log and σ is the formation conductivity.


gv is the vertical response function which characterizes a very thin layer of
formation. The coordinate z  represents a position in the formation with
respect to a given tool depth z.
If the blurring process is well-behaved mathematically, one can theoret-
ically determine an inverse function that would restore the true formation
conductivity profile when applied to the log data. The process of finding an
inverse blurring function is known as inverse filtering. The inverse filter is
applied to the log data in a manner mathematically identical to the convolu-
tion process that produced the log in the first place. Since the conductivity
log is a sequence of measurements at discrete values of z, the inverse filter is
a set of discrete weights. Each weight is multiplied by a corresponding log
reading and then summed to produce a single depth sample of the corrected
log. The construction of the corrected log in this manner is called deconvo-
lution. Finding an inverse filter for gv (z) is finding a weighting function h(z)
3.2. INDUCTION TOOLS 83

such that  ∞
σ(z) = h(z − z  ) σa (z  ) dz  , (3.15)
−∞
where σ, σa , z and z  all have the same meaning as in Equation (3.14).
During the 1960’s, Doll [98] and others made significant progress on
the induction deconvolution problem. The commercial result of this work
was the three-point filter used to generate the ID log in Figure 3.25 and
described in detail in Chapter 5. The development of deconvolution filters
was greatly simplified by modern signal processing theory. One conventional
signal processing technique [206] is to tailor the frequency domain response
after Fourier-transforming the spatial response function. The deconvolution
filter is then computed with the Remez algorithm [185].
In the early 1980’s, such a filter was developed for the ID tool using its
low-conductivity (Doll) response function [219]. Applying the filter gives
a log that more closely resembles the true formation resistivity profile in
high resistivity formations. However, this filter yields poor results in low
resistivity formations because it does not accommodate changes in resistivity
level; it neglects the fact that when a response function changes, the filter is
no longer its inverse due to differences in the spatial extent of skin effect.
A solution to this nonlinear problem was first suggested by Moran and
Kunz [194]. They introduced the concept of “skin effect error signal,” which
is the difference between a log, σG , generated by Doll geometrical factor
theory and a measured R-signal log, σR . They showed that the X-signal
measures the part of the error signal caused by phase shift, and that the
phase shift is the dominant loss mechanism at induction frequencies for con-
ductivities up to 5 S/m. The skin effect error signal can then be defined
as
σER = σG − σR . (3.16)
The corresponding response functions, g, also obey the same relation
gER = gG − gR . (3.17)
Applying the deconvolution filter, h, to the measured R-signal in Equa-
tion (3.16) results in a deconvolution error defined by the relation
σDER = σDG − σDR , (3.18)
where D denotes deconvolved. These signals have response functions, f ,
that obey the same relation as the signals themselves,
fER = fG − fR . (3.19)
84 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Subsequently, Moran showed [191] that the X-signal spatial response, gX ,


strongly resembles the deconvolution error, fER , over a wide conductivity
range and could be used as a first-order correction for skin effect.
The Phasor processing algorithm was developed using this theoretical
background. A transformation was derived to match gX to fER which when
applied to the X-signal yielded the correct error information needed in Equa-
tion (3.16) to obtain an estimate of the formation conductivity free of skin
effect. As a second-order correction, a nonlinear magnitude fitting function
was computed to fit σX to σER . The resulting algorithm consists of linear in-
verse filter deconvolution based on the low-conductivity induction response,
combined with a skin effect correction based on the X-signal to compensate
for the nonlinearity problem. The expression summarizing Phasor process-
ing [48] is
z
max   z
max
σP (z) = h(z − z  ) σR (z  ) + α σX (z) b(z − z  ) σX (z  ), (3.20)
z  =zmin z  =zmin

where σP is the Phasor corrected apparent conductivity, σR is the measured


R-signal and σX is the measured X-signal. h is the deconvolution filter and
b is the X-signal fitting filter. h and b are both finite impulse response (FIR)
filters, and α is the nonlinear element. The algorithm is termed “Phasor”
deconvolution because the error resulting from skin effect is recovered from
phase information contained in the X-signal. Because Equation (3.20) was
derived using 1D functions computed in homogeneous media (to obtain a
real-time correction), extensive validation was performed by Phasor process-
ing a large number of computed logs in 2D formations (such as Figure 3.26).
Prior to the introduction of array tools, the Phasor algorithm was up-
graded in order to further enhance vertical resolution. To improve the ver-
tical resolution of ID, high-frequency information is “borrowed” from IM
and incorporated in the ID response. To accomplish this, there is one fil-
ter for ID, another for IM and a third for extracting the information from
IM that is missing from ID. This allows the vertical response of both tools
to be matched exactly. The filters are also designed to keep the original
depth of investigation of each tool intact for invasion interpretation. This
modification is known as Enhanced Resolution (ER) Phasor processing [44].
Figure 3.26 shows ER Phasor processed IM and ID logs in the bench-
mark formation. Note the improved vertical resolution of IM and ID, par-
ticularly in resistive beds, compared to the conventionally processed logs in
3.2. INDUCTION TOOLS 85

Figure 3.26: Phasor processed Dual Induction log.

Figure 3.25. The Phasor processed logs have retained the same depth of
investigation as the conventionally processed logs. In addition, Phasor pro-
cessing has significantly reduced the horns in the conductive beds between
40 and 70 feet.
A dip correction algorithm [45] is also incorporated in Phasor processing
for use in deviated wells and dipping formations. A series of step profiles
was computed for various conductivity contrasts and dip angles, and used
to derive a set of vertical response functions with dip effect. Inverse filters
were developed from these results and then applied in the same manner as
the conventional filters to correct the dip-induced shoulder effect. Phasor
dip correction is accurate up to 60◦ .
86 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

1.00

12 ft

4 ft

Earth
0.10

1.0 10.0

Figure 3.27: Two-height lift curve for correcting the R-signal of the AIT-H
39 inch array for sonde error.

The Phasor induction tool is called DIT-E. The ID and IM coil con-
figurations are identical to those of the previous versions of DIT, shown
in Table 3.4. The major upgrade in DIT-E was the introduction of digi-
tal electronics. Because data was sent uphole via digital telemetry, more
measurements could be accommodated, with the X-signal being the primary
addition.
DIT-E was also the first tool to correct sonde error using the modeled
response of an induction tool in air above the earth [51]. Sonde error is a spu-
rious voltage generated by a coupling between metal parts of a logging tool
(e.g., laterolog electrodes, pressure bulkheads, mechanical components) and
the induction coils, which causes an error of several mS/m in the received
signal. Sonde error was traditionally measured and adjusted by placing a
sonde high enough above ground level for the signal from the earth to be
small. However, with the exception of some desert locations, there is usu-
ally a residual ground signal, which causes an offset error in the sonde error
correction. In the model-based correction method, the response of a horizon-
tal sonde is computed at two different heights above the ground for a wide
range of earth conductivities. These results are used to construct a chart
3.2. INDUCTION TOOLS 87

relating the signal difference at the two heights to the earth signal at the
uppermost measurement position. Entering the actual measured differences
in the chart gives the true earth signal at the upper measurement position.
This method is also used to correct sonde error for the AIT array induction
family of tools. The correction procedure is illustrated in Figure 3.27 [33].

3.2.5 Array Induction Tool (AIT)

With Phasor processing, the amount of information that could be extracted


from Dual Induction logs reached its limit. As grosser environmental effects
were corrected by Phasor or similar processing, it became apparent that
improvements were still needed in:

- Estimating Rt in the presence of deep invasion,


- Interpreting complex invasion profiles, such as transition zones, annu-
lus invasion and vertical segregation caused by gravity,
- Extending vertical resolution to one to two feet,
- Correcting errors in resolution enhancement caused by cave effect.

Making the deep induction tool longer has often been suggested as a
way to obtain logs that read closer to Rt . However, long tools have poorer
vertical resolution and are influenced to a significant degree by skin effect.
“Superdeep” induction arrays have been proposed that enhance far field
response by canceling the near field [232]. However, skin effect also re-
duces the depth of investigation of these arrays as formation conductivity
increases. An approach which provides more information about both Rt
and the invasion profile is recombining multiple arrays to produce a set of
measurements with several different depths of investigation, and then in-
verting the measurements radially to obtain an estimate of Rt . This concept
was first proposed in the 1950’s by Pupon [205], but the limited amount of
data that could be returned to the surface on the logging cable at that time
prevented the development of such an array tool. By the late 1980’s data
transmission was no longer a problem, and this concept became the basis of
the AIT Array Induction Imager family of tools. These tools abandon the
fixed-focusing principle of previous induction tools; they are constructed of
eight independent arrays, each with its own unique spatial response.
There are presently two types of AIT tools in the field: the AIT-B (stan-
dard) tool and the shorter AIT-H (Platform Express) tool. Figure 3.28 [33]
88 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.28: Coil configurations for AIT-B (left) and AIT-H (right).

shows the coil configurations for these two tools (AIT coil spacings and turns
are not listed in a table because they are proprietary at this time). Each
array is mutually balanced and consists of a single transmitter coil, which is
common to all arrays, and two receivers. Coil spacings range from 6 inches
to 6 feet.
The AIT-B tool [151] operates at three frequencies: 25, 50 and 100 kilo-
hertz. Both the R-signal and the X-signal are acquired by each array at
one or two frequencies suitable for that array length. The AIT-H tool [46]
operates at 25 kilohertz and also acquires the R-signal and the X-signal from
each array. All measurements are recorded every 3 inches in depth. Stability
is maintained over wide temperature and pressure ranges by using a metal
mandrel [50] and ceramic coil forms. Stable response is particularly impor-
tant for the shortest arrays, which are used to characterize borehole effect
and the near-borehole environment.
3.2. INDUCTION TOOLS 89

Integrated Radial Response

Radius (inches)

Figure 3.29: Integrated radial response of the five AIT logs.

AIT tools use nonlinear processing methods to combine several of the


eight array measurements at a time in such a way as to create a log focused
at a designated region of the formation. A total of five logs are generated.
These logs have median depths of investigation of 10, 20, 30, 60 and 90
inches. Integrated radial response functions for these five logs are shown
in Figure 3.29 [33]. Because AIT processing yields median responses that
are constant both radially and vertically over a wide range of formation
conductivities, response function curves do not vary with conductivity.
The vertical resolution of each AIT log is closely matched to that of
the others. There are three resolution widths available: 1, 2 and 4 feet.
Vertical response characterizing these three resolution widths is shown in
Figure 3.30 [33]. AIT processing contains no assumptions about the invasion
profile, so all five logs are interpretable in the same manner as DIT logs.
All environmental corrections are automatically built into the five AIT
logs prior to display. Therefore, there are no published correction charts for
AIT. The first step before log formation is the correction of all raw array
responses for borehole effect. The borehole correction algorithm [118, 129,
188] is based on tables of forward modeled array responses as a function of
four parameters: borehole radius, mud conductivity, formation conductivity
and the standoff of the tool from the borehole wall. An exact description of
all tool dimensions is included in the model. Mud conductivity is measured
by a small electrode tool at the nose of the sonde. A caliper samples the
90 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Vertical Response

Depth (inches)

Figure 3.30: The three vertical responses of AIT logs.

borehole size at 1-inch intervals, which is fine enough to determine cave


effect.
The five AIT logs are each formed as weighted sums of the the borehole
corrected array measurements. Details of the weighting method are found in
Barber and Rosthal [47], and are briefly summarized here. The response of
each individual array can be characterized by a 2D Born response function.
The relationship is given by
 ∞  ∞
σa (n)
(z) = gn (ρ, z − z  ) σ(ρ, z  ) dρ dz  , (3.21)
−∞ 0

where σa (n) is the measured log from the nth array, gn is the Born response
function for that array and σ is formation conductivity. z  is a position in
the formation with respect to a given tool depth z, and ρ denotes radial
distance. To generate an AIT log, a weighted sum of the individual array
measurements is computed over a depth range from zmin to zmax surrounding
the output log value. This log formation process is given by

N z
max
σlog (z) = wn (z  ) σa (n) (z − z  ), (3.22)
n=1 z  =zmin

where σlog is one of the five AIT logs, σa (n) is defined as above, N is the
total number of arrays contributing to the log and wn is the appropriate set
of weights for the nth array.
3.2. INDUCTION TOOLS 91

Figure 3.31: Schematic representation of the AIT log forming process.

The log produced by this process is different from logs produced by any
of the individual arrays, and can also be described by a 2D response func-
tion, glog (ρ, z). Each AIT log then has the the relation to the formation
conductivity distribution σ(ρ, z  )
 ∞  ∞
σlog (z) = glog (ρ, z − z  ) σ(ρ, z  ) dρ dz  . (3.23)
−∞ 0

The composite response function, glog (ρ, z), is a weighted sum of the re-
sponse functions of each of the individual n arrays (for both the R-signal
and the X-signal). This relationship is given by

N z
max
glog (ρ, z) = wn (z  ) gn (ρ, z − z  ). (3.24)
n=1 z  =z min

The filter weights, wn (z  ), are determined using Equation (3.24). The


problem is conceptualized in Figure 3.31 [33], which illustrates how the 2D
raw array responses (left) are combined so that the 2D AIT response (right)
and the resulting logs are optimum.
92 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

The optimization process is an extension to two dimensions of a method


first proposed by Doll [98] for reducing shoulder effect. A system of linear
equations is constructed which relates the weighting functions so that the
resulting log has: (1) the desired radial response, (2) the desired vertical
response, and (3) a smooth near-borehole response. The linear equations set
is over-determined to ensure smoothness and must satisfy several constraints
(such as the correct answer must be obtained in any homogeneous medium,
and the median depth of investigation is equal to some preset value).
A least squares technique is used to derive the weighting functions. After
the weights are determined for a given log response, Equation (3.22) is used
to produce the output log, with a different sets of weights developed for each
of the five output logs. Each output log is a weighted combination of several
input array measurements.
The log processing formalism is valid only at the one background con-
ductivity at which the Born responses are determined, and the weights in
Equation (3.22) are functions of this effective background conductivity. To
find the effective background conductivity for actual log values, the array
response is compared with synthetic logs modeled at several different con-
ductivity levels using Born response functions. The effective background
conductivity at any point is the conductivity for which the Born log matches
the array response. Changes in formation conductivity are handled by de-
veloping several filter sets for a range of background conductivities. The
background conductivity also represents an average of the formation con-
ductivity that is responsible for skin effect. X-signals are used in skin effect
correction, as in Phasor processing.
Because the Born-based processing becomes less accurate at high con-
trast bed boundaries, it is being replaced by a multi-array version of Phasor
processing. If the zero conductivity filter set in Equation (3.24) is used on
the unprocessed R-signals, σlog can be treated as σR in Equation (3.20), and
a Phasor correction based on the X-signal can then be derived. This new
algorithm was released to the field in early 2001.
AIT logs in the benchmark formation are shown in Figure 3.32 for the
2-foot vertical resolution width. Vertical resolution is greatly improved in
comparison to the conventionally processed DIT logs in Figure 3.25. Res-
olution is also improved in comparison the the Phasor processed logs in
Figure 3.26 (in Figure 3.26 in the 3-foot resistive bed between 80 and 83,
only SFL reads close to Rt of 50 ohm-m, with ID and IM reading lower).
3.2. INDUCTION TOOLS 93

Figure 3.32: AIT log.

The main advantage of AIT over all versions of DIT is the additional
information provided about invasion. The 10 inch AIT curve reads very
close to Rxo in the invaded beds between 27 and 37 feet, and between 113
and 123 feet. In addition, the 90 inch curve reads much closer to Rt than
the Phasor processed ID curve in these invaded beds. The 5 AIT curves
provide sufficient information to invert complex invasion profiles including
transition zones or annuli. A four-parameter inversion algorithm [149] is
available which computes real-time logs of Rt , Rxo and invasion radius at
the well site. The algorithm includes an option for introducing a smoothly
varying transition zone between Rxo and Rt which is bounded by inner and
outer radii. Inversion results can be displayed either as logs or as color
94 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.33: Limits of use for AIT tools.

resistivity images.
The first AIT logs were corrected for dip effect using a modified version
of the Phasor dip correction algorithm. However, this method is relatively
useless for horizontal well interpretation because it is limited to angles less
than 60◦ . In 1998, a maximum entropy inversion algorithm (MERLIN) [49]
was developed to provide more accurate Rt and invasion interpretation in
highly deviated and horizontal wells. MERLIN inversion was tested on a
wide range of invaded dipping bed benchmark cases generated by a 3D finite
difference code [18] and shown to be accurate up to 85◦ dip. Results can be
displayed both as logs and as non-axisymmetric resistivity images.
Traditionally, induction tools were run primarily in oil-based muds, or in
fresh muds where Rxo > Rt . However, the radial processing algorithm for
AIT works as well for Rxo < Rt as for Rxo > Rt , within limits. The main
restriction to using AIT in salty muds is the ability to perform accurate
borehole corrections. The chart in Figure 3.33 [33] characterizes the limits
of use for the AIT family of tools as a function of borehole and formation
resistivity and borehole size. Obviously, if the mud is very salty or the
borehole is very large or in bad shape, the laterolog remains the resistivity
tool of choice. For most applications where Rt /Rm > 500, laterologs provide
a better estimate of Rt . When it is possible to run AIT and laterolog tools
3.2. INDUCTION TOOLS 95

together, the two tools always provide a better total answer than either tool
alone, especially in the presence of invasion.
Other service companies have also introduced array tools to replace their
dual induction tools. BPB (now Reeves Wireline) had one of the earliest
array tools [218]. Their tool has main spacings of 20, 30, 40 and 60 inches
and takes the name of its high resolution processing, VECTAR [110]. In
1996, Baker-Atlas introduced an array induction tool, called the HDIL High
Definition Induction Log [53, 54]. Like the AIT, this tool generates five logs
with depths of investigations of 10, 20, 30, 60 and 90 inches. A 120 inch
curve is also available. In 2000, Halliburton announced the introduction
of their new array induction tool called the HRAI High Resolution Array
Induction [56].

3.2.6 Russian induction tools

After the former Soviet Union opened exploration to outside companies in


the early 1990’s, Western countries found a need to model Russian resistiv-
ity tool response in order to aid in log interpretation. Induction tools first
appeared in the former Soviet Union in the late 1950’s [146]. Induction tool
development was motivated by the poor performance of unfocused lateral
tools in thin beds, in low resistivity formations and in wells drilled with oil-
base mud [269]. Like Western induction tools, Russian tools are designed
with a main transmitter–receiver coil pair and from two to six auxiliary fo-
cusing coils. The most significant difference between Russian and Western
induction tools is in instrumentation; Russian tools are designed for sim-
plicity rather than for ease of interpretation. Russian induction logs are
often displayed on a linear conductivity grid rather than the conventional
logarithmic resistivity scale.
Coil configurations for several of the most commonly encountered Rus-
sian induction tools [162] are given in Table 3.5 through Table 3.9. 6E1,
4I1, 4F0.75 and 8I1.4 all operate at 20 kHz. 6F1 operates at 50 kHz. Multi-
coil focused tool response is computed in the same manner as Western tools,
using Equation (3.10). Coil arrangements are described by nomenclature
similar to that used by Western logging companies. For example, 6E1 refers
to a six coil sonde with a one meter spacing between the main transmitter
and receiver coils. “E” indicates a tool combined with other electrical mea-
surements, “I” indicates a general induction tool and “F” indicates a focused
tool. The fourth coil of the four-coil sondes in Table 3.7 and Table 3.8 is
96 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Moment Position(in.)
Transmitters: 1. 20.
-0.25 10.
-0.077 -50.39
Receivers: 1. -20.
-0.25 -10.
-0.077 50.39

Table 3.5: 6E1 coil configuration.

Moment Position(in.)
Transmitters: 1. 19.69
-0.05 2.95
-0.075 -35.63
Receivers: 1. -19.69
-0.05 -2.95
-0.075 35.63

Table 3.6: 6F1 coil configuration.

Moment Position(in.)
Transmitters: 1. 10.63
-0.216 -5.12
Receivers: 1. -28.74

Table 3.7: 4I1 coil configuration.

Moment Position(in.)
Transmitters: 1. -5.91
-0.2 6.30
Receivers: 1. 23.62

Table 3.8: 4F0.75 coil configuration.


3.2. INDUCTION TOOLS 97

Moment Position(in.)
Transmitters: 1. 27.56
-0.07 13.78
-0.01 -7.87
0.04 -47.24
-0.10 51.18
Receivers: 1. -27.56
-0.20 -37.40
-0.11 0.

Table 3.9: 8I1.4 coil configuration.

not listed because it contributes very little to the tool response; it is used
to cancel the mutual signal and its moment is extremely small. The 6E1 is
practically identical to 6FF40, as can be seen from a comparison of Table 3.5
and Table 3.3 (the ratio of 6E1 moments is equivalent to the ratio of 6FF40
turns).
The relative depths of investigation of the sondes in Table 3.5 through
Table 3.9 are indicated by the integrated radial geometrical factors shown in
Figure 3.34. Curves for ID (6FF40) and IM are also shown for comparison.
Environmental correction charts for present-day Russian induction tools
are similar to charts used for Western tools in the 1960’s and 1970’s. Cor-
rections are applied in sequence for borehole effect and shoulder bed effect,
when needed. “Tornado” charts (see Chapter 5) for various combinations
of induction and lateral tools with different depths of investigation are also
used to determine Rt in invaded formations.
One chart that is unique to Russian induction tools is the so-called skin
effect correction chart, which is the first correction that is ordinarily applied.
For resistivities above 20 ohm-m, skin effect correction is insignificant. How-
ever, at high conductivities (low resistivities), signal level rises less rapidly
and in a progressively nonlinear fashion as formation conductivity increases,
and skin effect correction cannot be neglected. Western practice has been
to perform skin effect correction electronically prior to the display of a log,
even before the advent of computerized log processing. Russian practice is to
perform the correction manually using either a look-up table or a chart such
as the one shown in Figure 3.35. To use the chart, the measured apparent
98 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.34: Integrated radial geometrical factors showing the differences in


depths of investigation between Russian induction tools. ID (6FF40) and
IM are also shown.

Figure 3.35: Skin effect correction chart for five Russian induction tools, also
showing ID (6FF40) and IM.
3.2. INDUCTION TOOLS 99

Figure 3.36: 6E1 log.

conductivity is entered on the left, and the skin effect corrected conductivity
is read at the bottom.
6E1 response in the benchmark formation is shown in Figure 3.36, along
with 6FF40 for comparison. The need for skin effect correction of the 6E1 log
can be seen from its large departures from Rt in the 0.5 ohm-m conductive
beds and in the upper and lower thick 5 ohm-m shoulder beds. In the
50 ohm-m resistive beds, the simple 6FF40 3-point deconvolution of the
1960’s provides a slight improvement over the unprocessed 6E1 log, while
deconvolution over-corrects 6FF40 in the upper beds between 40 and 65
feet.
100 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

3.3 Propagation tools

Propagation tools is the generic term used in resistivity logging for induction-
type tools that measure the phase shift and attenuation of the voltage cre-
ated by a transmitter current between two receivers. The most well-known
propagation tools are 2-MHz tools used in logging while drilling (LWD).
2-MHz tools are similar to induction tools in that their antennas are verti-
cal magnetic dipole coils mounted on a mandrel. However, while induction
tools obtain the apparent resistivity from the skin-effect corrected R-signal,
2-MHz tools and other propagation tools measure the skin effect directly.
There are also several propagation tools which are used in wireline log-
ging. These tools operate at frequencies ranging from a few hundred kilohertz
to over a gigahertz. Schlumberger tools include the 25-MHz Deep Propaga-
tion Tool (DPT), and the 1.1-GHz Electromagnetic Propagation Tool (EPT).
As the frequency increases above the kilohertz range, dielectric effect be-
comes more significant. 2-MHz LWD measurements are first corrected for
dielectric effect and then scaled to resistivity. Measurements above 2-MHz
are ordinarily converted to an apparent resistivity and an apparent dielectric
constant using tables of tool response computed in homogeneous isotropic
media.
Computer codes developed to model induction tool response can also be
used to model propagation tool response with minor modifications. For finite
element and finite difference codes, mesh sizes must be adjusted to account
for the shorter wave lengths generated at higher frequencies. For spectral
integration codes, integration intervals and paths may have to be changed,
depending on the generality of the code.

3.3.1 2-MHz tools for logging while drilling

Commercial resistivity measurements made while drilling first became avail-


able in the late 1970’s. Because the drilling environment is much more ad-
verse than the wireline logging environment, a simple short normal mounted
behind the drill bit was used as the first LWD resistivity tool [250]. Since
short normals have a shallow depth of investigation and relatively poor ver-
tical resolution, they were only able to provide enough information for basic
interpretation, such as correlation of geological markers and estimation of
gross water saturation [154]. Normal tools had another major limitation:
they could only be run in water-based mud environments.
3.3. PROPAGATION TOOLS 101

28 in. Transmitter 1

3 in. Receiver 1
Measure
-3 in. Receiver 2 Point

-28 in. Transmitter 2

Figure 3.37: Antenna configuration for the 2-MHz CDR tool.

In 1983, NL Information Services (now part of Sperry-Sun) introduced


the Electromagnetic Wave Resistivity (EWR) tool [210, 77]. NL adopted an
induction-type measurement in order to expand the LWD resistivity market
to oil-based mud environments. Since it was difficult to engineer a conven-
tional induction tool on a steel drill collar using the technology of the early
1980’s, a higher frequency propagation measurement was considered to be
more practical. A frequency of 2-MHz was chosen because it was the lowest
frequency at which accurate phase shift measurements could be made on
a drill collar at that time (the attenuation was not used because NL had
problems making accurate attenuation measurements). The EWR tool had
a single transmitter and measured the phase shift of the voltage between
two receivers located at 24 and 30 inches from the transmitter. The choices
made in designing this tool laid the groundwork for LWD tools introduced
by other service companies in the following years.
In 1987, Schlumberger introduced the Compensated Dual Resistivity
(CDR) tool [74, 75]. The antenna configuration is shown in Figure 3.37.
The CDR is still run today. However, it is in the process of being replaced
by an array tool, the ARC5 (described later in this section). The ARC5
includes the CDR spacings as one of its arrays. The CDR tool broadcasts
102 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

a 2-MHz electromagnetic wave and measures both the phase shift and at-
tenuation of the wave between two receivers. Note in Figure 3.37 that the
CDR has two transmitters. The phase shift and attenuation generated by
Transmitter 1 between Receiver 1 and Receiver 2, and by Transmitter 2 be-
tween Receiver 2 and Receiver 1, are averaged together to symmetrize the
response. This averaging is known as borehole compensation because it also
reduces the effect of borehole rugosity.
The averaged phase shift and attenuation are transformed to two in-
dependent resistivities: RPS (phase shift, shallow) and RAD (attenuation,
deep). A correction is performed for dielectric effect before the raw data is
converted to apparent resistivity. A strong correlation between the relative
dielectric constant ( ) and resistivity (R) was determined using over 300
core samples. This correlation takes the explicit form  = 110 R−0.35 [75].
Since this equation expresses  as a function of resistivity, the dielectric con-
stant can be eliminated as an independent quantity. The dielectric-corrected
phase shift and attenuation are then converted to resistivity using a table
look-up based on polynomial approximations of computed tool response in
homogeneous isotropic media of known resistivity, Rt .
Both RPS and RAD are relatively insensitive to borehole size and mud
resistivity. Borehole correction is only necessary in conductive holes with
large washouts when the Rt /Rm contrast is greater then 100 to 1 [74].
Invasion is usually quite shallow at the time of drilling when LWD logs are
run. However, LWD logs may also be recorded each time that the drill string
is pulled to replace the drill bit. At these later times, invasion can become
much deeper. The two resistivities, RPS and RAD provide two independent
depths of investigation for the interpretation of invasion. The reason that two
depths of investigation can be obtained from a single measurement is made
clearer by examining the behavior of the electromagnetic field. Surfaces of
constant phase and constant amplitude generated by the uppermost CDR
transmitter in a 1 ohm-m formation are shown in Figure 3.38 [75]. The areas
between the constant contours passing through the two receivers is shaded
to denote the differential measurement. The surfaces of constant phase are
spheres because the wave travels with the same speed in all directions. The
surfaces of constant amplitude are toroids because the wave is stronger in
the radial direction than in the vertical direction, which is characteristic
of vertical magnetic dipole antennas. The attenuation corresponds to a
significantly deeper region than the phase shift. In this 1 ohm-m formation,
the depth of investigation (defined as 50% of the radial response) is 30 inches
3.3. PROPAGATION TOOLS 103

Figure 3.38: Surfaces of constant phase (left) and amplitude (right) for an
electromagnetic wave generated by the uppermost CDR transmitter. Each
phase surface represents an interval of 10◦ , and each amplitude surface rep-
resents an interval of 3 dB.

for RPS, and 50 inches for RAD [8]. The depths of investigation of both
RPS and RAD become shallower as the formation resistivity level decreases
(conductivity increases) because of increasing attenuation of the signal due
to skin effect.
The depth of investigation can also be studied by modeling tool response
in invaded formations. Figure 3.39 shows CDR radial response for a case
where Rxo > Rt and Figure 3.40 shows the radial response for Rxo < Rt [135].

Figure 3.39: CDR radial response for Rxo > Rt .


104 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.40: CDR radial response for Rxo < Rt .

In both figures, RPS and RAD are plotted as a function of increasing invasion
radius. In Figure 3.39, RPS reads consistently closer to Rxo , indicating that
RPS is the shallower of the two measurements. In Figure 3.40, RPS is
again consistently shallower than RAD. In this case the RPS curve extends
below the value of Rxo between a radius of 30 and 50 inches because of wave
reflection at the invasion front. In general, the depth of investigation of RPS
is 10 to 20 inches shallower than that of RAD.
The 2D response of the CDR can be characterized using Born response
functions [135]. Section 4.1.4 describes the derivation of Born response func-
tions and their application to the modeling of 2-MHz tool response. Figures
4.12 and 4.13 compare Born response functions for the CDR phase shift and
attenuation measurements in 2 ohm-m and 10 ohm-m formations. Attenu-
ation due to skin effect can clearly be seen in the 2 ohm-m formation. The
differences in the radial and vertical volumes of investigation of the phase
shift and attenuation measurement are also apparent.
Vertical resolution and depth of investigation are characterized in more
detail in Figure 3.41 which shows CDR logs in the benchmark formation. In
the 3-foot and 10-foot conductive uninvaded beds between 47 and 50 feet,
both RPS and RAD read near Rt . However in the resistive uninvaded beds
between 80 and 103 feet, only RPS reads near Rt in the 10-foot bed, and
neither RPS or RAD read near Rt in the 3-foot bed. Like conventional
induction logs, the vertical resolution of unprocessed 2-MHz logs is signifi-
3.3. PROPAGATION TOOLS 105

Figure 3.41: CDR logs.

cantly poorer in resistive beds than in conductive beds of the same size. In
both resistive and conductive beds, the vertical resolution of RPS is much
sharper than that of RAD.
In the conductive invaded bed between 27 and 37 feet, the separation
between the two curves clearly indicates the presence of invasion, although
neither curve reads the value of Rxo or Rt . In the resistive invaded bed
between 103 and 113 feet, the separation between the two curves is not as
great as in the conductive invaded bed. RPS now reads above Rxo because its
depth of investigation changes with Rxo /Rt contrast, and RAD is influenced
by shoulder effect
In 1995, Schlumberger introduced the Array Resistivity Compensated
tool (ARC5). Five independent phase shift and attenuation measurements
106 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Wear band

34 in. Transmitter

22 in. Transmitter

10 in. Transmitter
3 in. Receiver Measure
Point
-3 in. Receiver

-16 in. Transmitter

-28 in. Transmitter


Wear band

Figure 3.42: Antenna configuration for the 2-MHz ARC5 tool.

are made at 2 MHz. The number of measurements was deliberately chosen


to be the same as that of wireline array induction tools to allow the sharing
of interpretation software.
The ARC5 antenna configuration is shown in Figure 3.42 [58]. There are
five transmitters and two receivers. The phase shift and attenuation of the
signal broadcast by each transmitter is measured between the two receivers
for a total of five raw phase shifts and five raw attenuations. The raw
measurements are linearly combined using a technique called mixed borehole
compensation [58], and then transformed into five calibrated phase shift and
attenuation resistivities using a method similar to the one described above
for CDR. The resulting resistivity logs are characterized by the antenna
spacings: 10, 16, 22, 28 and 34 inches. The 28-inch spacing yields a log
identical to CDR.
Figure 3.43 shows ARC5 logs in the benchmark formation. The vertical
resolution of all five RAD (attenuation) curves is very similar. The vertical
resolution of the five RPS (phase shift) curves is also quite similar. RPS
has consistently sharper vertical resolution and reads closer to Rt in all of
3.3. PROPAGATION TOOLS 107

Figure 3.43: ARC5 log.

the uninvaded beds. In the conductive invaded bed between 27 and 37 feet,
there is considerable separation between the RAD curves to aid in invasion
interpretation, while there is not much separation between the RPS curves.
Conversely, in the resistive invaded bed between 103 and 123 feet, there is
separation between the RPS curves, while the RAD curves remained grouped
together. The separation between curves would of course be different for
different invasion radii.
There are currently three other major service companies which provide
LWD resistivity services in addition to Schlumberger. In 1991, Sperry-Sun
introduced a version of the EWR with three phase shift and three attenuation
measurements [209], and in 1993 they introduced the EWR-Phase 4 [200]
108 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

with four phase shift measurements operating at 1 MHz and 2 MHz. Both
versions of EWR have coil spacings ranging from 12 to 42 inches and are run
without borehole compensation.
In 1989, Teleco introduced the 2 MHz Dual Propagation Resistivity
(DPR) tool [113]. This tool measured the phase shift and attenuation at
receivers located 27 and 35 inches from a single transmitter (borehole com-
pensation was not used). Teleco was taken over by Baker Hughes, and in
1993 the DPR was replaced by the Multiple Propagation Resistivity (MPR)
tool [186]. This tool makes borehole compensated phase shift and attenu-
ation measurements using two transmitters and two sets of receivers, with
spacings ranging from 23 to 35 inches. The tool operates at both 2 MHz
and 400 kHz.
In 1993, Halliburton introduced the 2 MHz Compensated Wave Resistiv-
ity (CWR) tool [124]. This tool makes a set of shallow and deep phase shift
and attenuation measurements with borehole compensation. The transmit-
ter to receiver spacing is approximately 40 inches for the deep mode, and 20
inches for the shallow mode.

3.3.2 Deep Propagation Tool (DPT)

The interpretation of resistivity logs works well in situations where the for-
mation water salinity is known and is reasonably constant throughout zones
of interest. However, some reservoirs have water resistivities which vary
considerably from zone to zone. In such cases, the standard Archie in-
terpretation (see Section 5.1) is not accurate because values of the forma-
tion water resistivity, Rw , are not known for each zone. In addition, when
formation waters are extremely fresh, oil and water zones become difficult
to distinguish due to the similarity in their resistivities. The need for a
salinity-independent determination of hydrocarbon saturation led to the de-
velopment of tools which measure another basic electrical property of the
formation: the dielectric constant.
There are two factors that make the measurement of formation dielectric
constant attractive. The first is the large difference in the values of the
relative dielectric constants of oil and gas (1 to 5) compared to the dielectric
constant of water (60 to 80). The second is the greater sensitivity of dielectric
measurements to water volume rather than salinity. The dielectric constant
can be described as the measure of a material’s ability to store an electric
3.3. PROPAGATION TOOLS 109

95 in. R1 Amplitude 1
Phase 1
Far Attenuation Far Resistivity
Far Phase Shift Far Dielectric
R2 Amplitude 2
70 in. Phase 2
Cross Resistivity
Cross Dielectric
R3 Amplitude 3
50 in. Phase 3
Near Attenuation Near Resistivity
Near Phase Shift Near Dielectric
R4 Amplitude 4
25 in. Phase 4

0 in. T

Figure 3.44: DPT antenna configuration, showing how the various receiver
responses are combined to obtain apparent resistivity and dielectric constant.

charge. Because the water molecule is polar, its dielectric constant is much
higher than the dielectric constants of hydrocarbons and all other formation
materials. Thus the measured dielectric constant is primarily a function of
the water-filled porosity.
Russian scientists were active in the area of dielectric logging since the
early 1960’s, and they published extensively on the subject. In the 1970’s,
Texaco developed a 20-MHz mandrel tool for measuring the dielectric con-
stant beyond the invaded zone [78]. Prompted by this work, Schlumberger
introduced the Deep Propagation Tool (DPT) in 1981 [150]. The DPT an-
tenna configuration is shown in Figure 3.44 [220].
The DPT is a mandrel tool which was designed to be as sensitive as pos-
sible to the dielectric constant and resistivity of the formation beyond the
invaded zone. A frequency of 25 MHz was chosen because it provided good
sensitivity to dielectric permittivity, while at the same time maintaining a
deep depth of investigation over a large range of mud and formation resis-
tivities. There is one transmitter coil and four receiver coils located above
the transmitter.
110 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Phase Shift (degrees) R (ohm-m)

σ (mS/m) ε’

Attenuation (dB)

Figure 3.45: Chart illustrating the conversion of DPT phase shift and atten-
uation measurements to apparent resistivity and dielectric constant for the
“near” receiver pair.

As illustrated in Figure 3.44, the receivers are grouped in two pairs: the
“near” pair and the “far” pair. The relative phases and amplitudes of the
transmitted signal are recorded at the four receiver locations. These mea-
surements are converted into the phase shifts and attenuations between the
two receivers of each pair. Finally, the phase shifts and attenuation are trans-
formed to apparent resistivity and dielectric constant using an algorithm
based on tool response in homogeneous isotropic media. Figure 3.45 [220]
graphically illustrates this algorithm. Because signal level of the “far” atten-
uation becomes extremely small at low resistivity levels, a “cross” measure-
ment was substituted. The “cross” measurement uses the attenuation from
the “near” receiver pair and the phase shift from the “far” receiver pair.
This combination gives reasonable results because the depth of investigation
of the “near” attenuation approaches the depth of investigation of the “far”
phase shift (attenuation measurements are deeper than comparable phase
shift measurements as described in Section 3.3.1). The standard DPT log
displays only the “near” and “cross” measurements.
Both the mud and formation resistivities limit the practical use of DPT
measurements. The lower the mud or formation resistivity, the lower are
the received signal levels. Acceptable accuracy is obtained in muds with
3.3. PROPAGATION TOOLS 111

Figure 3.46: DPT log.

resistivities of 0.2 ohm-m or greater in 8 inch diameter boreholes. In addition,


the resolution of the dielectric measurement deteriorates when the formation
resistivity is low. The vertical resolution of DPT measurements ranges from
four to eight feet. Typical depths of investigation are around 20 inches for the
“near” measurement, and from 30 to 40 inches for the “cross” measurement.
Figure 3.46, shows computed DPT apparent resistivity and dielectric
logs [23] in the benchmark formation. In addition to characterizing vertical
resolution and depth of investigation, this figure also illustrates some of the
problems involved in interpreting DPT logs. In the two conductive beds
between 47 and 70 feet, an accurate inversion for apparent resistivity and
dielectric constant is not possible because of low signal level. In the invaded
conductive bed between 27 and 37 feet, the inversion gives more accurate
results because the signal level is higher on account of the resistive invasion.
112 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

In the resistive beds between 80 and 127 feet, both the “near” and “cross”
logs track the formation parameters reasonably well. In the invaded bed
between 113 and 123 feet, the 30 inch invasion is a bit too deep for the “cross”
measurement to read Rt . Note the horns that appear on the dielectric logs
near bed boundaries.
In practice, DPT logs proved to be more difficult to interpret than antici-
pated. It is well known that simultaneous measurements of conductivity and
dielectric constant are frequency dependent [243]. In general, conductivity
increases with with increasing frequency, and dielectric constant decreases
with increasing frequency. The term used to describe these changes is dis-
persion. Dispersion can also be affected by salinity. The interpretation of
dispersion becomes quite complex in the vicinity of 25 MHz. In addition,
the interpretation of shales and shaly sands is complicated by the fact that
the dielectric constant of shales can range from 5 to 25 depending on the
ratio of “bound water” to “total water.” Interpretation is still possible in
these complex situations, but accuracy becomes doubtful.
Because of these problems, the DPT was never widely used. Only a few
tools still survive in the field today. They are run occasionally in locations
where there is difficulty differentiating between fresh water and oil zones,
such as in South America.

3.3.3 Electromagnetic Propagation Tool (EPT)

The Electromagnetic Propagation tool (EPT) was designed to provide a shal-


low measurement of the conductivity and dielectric constant of the invaded
zone. The tool operates a frequency of 1.1 GHz. An advantage of measur-
ing the dielectric permittivity at this high frequency is that dispersion is
negligible because ionic relaxation effects are small. The EPT consists of
four microwave antennas (two transmitters and two receivers) mounted on
a metallic pad. The pad is rigidly attached to the sonde body and is pushed
against the borehole wall by a backup arm. Logging measurements are made
in a borehole compensated mode; each transmitter is turned on separately,
and the propagation time and attenuation of the transmitted electromag-
netic waves are measured in both directions by the receivers and averaged
together.
As shown in Figure 3.47 [126], there are two basic antenna configurations
corresponding to two different orientations of the dipole antennas on the pad.
3.3. PROPAGATION TOOLS 113

Endfire Broadside

Figure 3.47: Antenna configuration of the EPT-G Endfire (left) and Broad-
side (right) arrays, showing the power radiation patterns.

In one configuration, the antennas are mounted so that the dipole moments
point end-to-end to each other; this is called the Endfire magnetic dipole
(EMD) array. In the other configuration, the antennas are mounted so that
the dipole moments are side-by-side to each other; this is called the Broadside
magnetic dipole (BMD) array. Two different arrays are employed in order to
overcome the problems of standoff and signal level. The EMD is used under
ordinary conditions because it is the the deepest array, while the BMD is
used in lossy environments because of its higher signal level.
The power radiation patterns associated with each of the magnetic dipoles
(shown along with the antennas in Figure 3.47) illustrate why this is the case.
The BMD has most of its power transmitted directly along the surface of
the pad, as well as some power directed out into the formation. On the
other hand, the EMD has no net power flow in the direction of the receivers.
Instead, all of its power is directed outward toward the formation. Thus the
EMD can see deeper into the formation and is less sensitive to standoff, while
the BMD is more efficient (i.e., the signal falls off more slowly), making it bet-
ter suited for lossy (conductive) environments. The transmitter-to-receiver
114 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

spacings of the BMD are shorter than those of the EMD in order to obtain
as strong a signal as possible. Because the BMD has a shallower depth of
investigation, it is more susceptible to mudcake effect and pad standoff.
The dual array tool shown in Figure 3.47 is called the EPD-G. This
tool was introduced in 1987 [126] and is still currently in use in the field.
The EPT-G replaced an earlier version of the tool, called the EPT-D [61],
which was introduced in 1977. The EPT-D had large resonant slot antennas
that proved to be difficult to model and interpret [216]. These problems
prompted the development of the much smaller nonresonant dipole anten-
nas of the EPT-G. The size of these antennas is sufficiently small and their
transmitter-to-receiver spacing is sufficiently large that they can be modeled
as point magnetic dipoles. The accuracy of the point magnetic dipole model
for EPT-G has been demonstrated by the excellent agreement between ex-
periments and computer modeling studies [216, 37, 36, 30].
The response of the EMD and BMD arrays has been extensively modeled
and characterized in many different logging environments. These include
response to mudcake effect [114, 216], invasion [30], thin beds [37], dipping
beds [36], and a wide variety of 2D geometries [30]. Because of the BMD’s
shallow depth of investigation, it should always be corrected for mudcake
effect. The deeper EMD only requires correction when mudcakes are thicker
than 0.5 inches. Vertical resolution ranges from two to six inches. In general,
phase shift measurements have better vertical resolution than attenuation
measurements, and the BMD (in the absence of standoff) has slightly better
vertical resolution than the EMD. Depth of investigation ranges from one to
five inches, with response becoming shallower in more lossy environments.
The BMD is shallower than the EMD. Phase shift measurements are slightly
shallower than attenuation measurements. Skin depths at 1.1 GHz gives an
indication of the distance that an electromagnetic wave can travel. Under
normal logging conditions, skin depths range from 0.5 to 6 inches.
EPT phase shift and attenuation measurements are often scaled as travel
time in nanoseconds/m and attenuation in dB/m, which appear on log head-
ings as TPL (propagation time) and EATT (EPT attenuation). Apparent
resistivity and dielectric constant are computed from the measured phase
shift and attenuation using an iterative solution of the algebraic expression
for tool response in homogeneous isotropic media.
Most EPT interpretation methods are based on the assumption of a plane
wave model. Although this is an empirical model, it has been used with a
3.3. PROPAGATION TOOLS 115

Figure 3.48: EPT log.

reasonable degree of success, and it has been shown to be semiquantitatively


correct [116]. In the plane wave model, EPT response is represented as a
plane wave in a homogeneous medium. However, the dipole transmitters
generate an electromagnetic wave that is basically spherical. A spread loss
correction in commonly applied to the raw EPT measurements to compen-
sate for the loss in intensity of the wave as it propagates due to its geometrical
spreading. The corrected response better approximates planar propagation.
Plane wave corrected measurements are denoted on logs by TPPW and
EAPW. The plane wave model underlies the two most widely used inter-
pretation methods for EPT: CRIM (complex refractive index method) [220]
and tpo (standing for lossless formation propagation time) [220].
Because of the excellent vertical resolution of the EPT, it has been used
successfully in thin bed analysis. The computed logs shown in Figure 3.48
compare the vertical resolution of EPT (EMD) and AIT logs in a thinly
bedded sand–shale sequence. This log is is taken from a case study [36] where
116 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

EPT field logs were used to better define bed boundaries and resistivity
values for the iterative modeling of AIT response. Individual sand and shale
beds can be clearly identified by the differences in their dielectric constants.
The sands are the 2-foot bed at 105 feet and the 1-foot bed at 110 feet.
The AIT log with 2-foot processing is unable to resolve the 1-foot bed, while
EPT easily provides a high resolution value for Rt .
A computer interpretation technique called laminated sand analysis [10]
uses a joint interpretation of EPT, induction and nuclear logs to quantita-
tively evaluate the porosity, shaliness and water saturation in reservoirs with
beds as thin as two inches.

3.4 Electrode (laterolog) tools

For 20 years after the first well log was run in 1927 (see Section 1.2) the
only resistivity measurements available were conventional electrical surveys
(ES). The ES consisted of an SP measurement and three electrode tools: a
16-inch normal, a 64-inch normal and an 18-foot 8-inch lateral. Thousands
of ES logs were run every year in wells drilled all over the world [220].
Electrode tools are only used in water-based muds because they require
direct contact with a conductive mud column in order to inject current into
the formation. In relatively fresh muds, borehole effect has little influence
on apparent resistivity readings. However, in more conductive salty muds,
the current emitted by the unfocused normal and lateral tools can travel
inside the borehole for long distances before entering the formation. For this
reason, normal and lateral tools are particularly susceptible to borehole ef-
fect. Unfocused tools are also strongly influenced by other parasitic effects.
The mud that invades the formation can significantly affect the apparent
resistivity if the contrast between the mud resistivity and the connate wa-
ter resistivity is high. In addition, the long normal has difficulty resolving
beds that are less than 10 feet thick, and the unsymmetrical lateral logs are
notoriously complicated to interpret (see Figure 1.3).
Starting in the 1940’s, new electrode tools were designed to address these
problems. Each new tool that was introduced attempted to make the appar-
ent resistivity in a zone of interest as immune as possible from the proximity
of adjacent zones by means of various focusing methods. Major milestones
in electrode tool design are shown in Figure 3.49. Individual tools and fo-
cusing methods will be described in detail in the remainder of this section.
3.4. ELECTRODE (LATEROLOG) TOOLS 117

Microlog - Normal and Inverse on a pad


LL3 - Guard log
1950
LL7 - Active focusing
Microlaterolog - LL7 on a pad

HDT - Resistivity dipmeter

1960 Proximity Log - LL3 on a pad


LL8 - Combined with Dual Induction

ULSEL - Ultra-long Normals (for saltdomes)


DLT - Dual Laterolog tool
1970
SFL - Combined with Dual Induction
Micro-SFL - Rxo pad tool for DLT

1980
DLT - Groningen effect correction
FMI - Formation Micro-electrical Imager

1990
ARI - Azimuthal Resistivity Imager on DLT
HALS - Short DLT for Platform Express

HRLA - High Resolution Laterolog Array


2000

Figure 3.49: Time-line of Schlumberger laterolog tool development.

Some of the tools listed in Figure 3.49 are now obsolete (namely, the lateral,
normals, LL3, LL7 and LL8). However, their descriptions are included here
because many wells in producing fields have been logged with these tools,
and the old logs are still used to determine the placement of new wells. In
addition, some of the focusing methods developed for obsolete tools have
been adapted for use by new array laterolog tools.

3.4.1 The Normal

The normal is the simplest electrode device for measuring formation resis-
tivity. Figure 3.50 [220] shows the electrode configuration for the normal
tool. A low frequency survey current of constant intensity I is emitted by a
current electrode, A. The source of the current is a generator at the surface.
The current return is at B, which considered to be at “infinity”. A voltage
electrode, M , measures the potential with respect to a reference electrode,
N , which is located far enough from the current source to be assumed at
118 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.50: Idealized electrode configuration for the normal tool.

zero potential.
In practice, N is located on the bridle (the insulation-covered lower end
of the logging cable), and B is the cable armor above N . B was moved
downhole to implement the simultaneous measurement of SP. In order to
eliminate interference with the SP measurement, which is carried out at
DC, a source current of approximately 20 Hz is used. The AC source also
helps to avoid electrode polarization. Electrodes A and M are on the sonde.
The distance AM is called the spacing (16 inches for the short normal and 64
inches for the long normal). The depth at which the log is recorded, called
the measure point, is at O. The measure point is located midway between
A and M because of symmetry imposed by reciprocity.
In a homogeneous isotropic medium of infinite extent, the equipontential
surfaces surrounding the current-emitting electrode A are concentric spheres.
From Poisson’s equation (see Section 4.1.5 on laterolog modeling), the po-
tential VM in an homogeneous medium created by the current I at electrode
3.4. ELECTRODE (LATEROLOG) TOOLS 119

M is
Rt
VM = I, (3.25)
4πrM
where Rt is the formation resistivity (in ohm-m) and rM is the radial dis-
tance between A and M (in meters). The formation resistivity in the region
between spheres passing through M and N can then be expressed as
VM − V N
Rt = K . (3.26)
I
The tool coefficient K is a constant which scales a tool’s current and voltage
readings to the formation resistivity. For normal tools, K equals 4πrM and
depends only on the distance between the current source A and the potential
electrode M . The voltage at N , VN , is used as a reference potential. Since
N for normal tools is located at a relatively large distance from the current
source (over 30 feet), VN is close to zero volts.
In a heterogeneous medium, such as invaded, thin beds traversed by a
borehole, the apparent resistivity measured by a normal tool, Ra , is approx-
imated using the relationship in Equation (3.26) as

VM − V N
Ra = K . (3.27)
I
Ra will be a good estimate of Rt only under the most favorable conditions.
Therefore, the determination of Rt in geological beds from normal logs has
historically required the use of departure curves (described in Chapter 5),
as well as additional measurements by other tools.
The previous discussion is illustrated by the computed current patterns
and equipotential surfaces for a normal tool shown in Figure 3.51. The lo-
cations of the A and M electrodes for a 16-inch normal are indicated on the
figure. (The current patterns for a 64-inch normal tool would be identical,
with the voltage measurement at M moved to 64 inches above zero.) The ge-
ometry modeled is a thick, uninvaded bed penetrated by a vertical borehole.
The formation resistivity is 10 times the mud resistivity (Rt /Rm = 10). The
results shown were generated using a 2D semi-analytic code[24, 25]. The
presence of the conductive borehole disturbs the straight-line radiation of
current. Current lines emanating from A travel inside the borehole and are
eventually diffracted into the formation; they become essential radial at a
distance of several borehole diameters into the formation. The equipotential
surfaces are quasispherical only at large distances from the current source.
120 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.51: Computed current patterns (dark lines with arrows) and equipo-
tential surfaces (orthogonal light lines) for a 16-inch normal in a borehole
penetrating a thick uninvaded bed. The formation is axisymmetric, and the
figure represents a plane of revolution about the center of the borehole. The
borehole wall is indicated by a dashed line. The positions of electrodes A
and M are indicated on the tool mandrel along the vertical axis at the left.

(Effects of bed boundaries and invasion on current radiation are illustrated


in Section 3.7.8.)
Figure 3.52 shows 16-inch and 64-inch normal computed logs in the
benchmark formation. These logs were generated using a 2D finite element
code [274, 119]. The shallower 16-inch normal was designed to characterize
the invaded zone, since half of the potential drop in a homogeneous medium
occurs within a radius of 2AM , or 32 inches (see Equation (3.25)). The
deeper 64-inch normal was designed to read closer to Rt . In Figure 3.52,
neither of the tools reads Rt , even in the uninvaded beds. Note that in the
resistive bed between 80 and 83 feet, the 64-inch normal log actually deflects
3.4. ELECTRODE (LATEROLOG) TOOLS 121

Figure 3.52: 16-inch and 64-inch normal logs.

downward with small spurious peaks above and below the bed. The distance
between the peaks is equal to the bed thickness plus the AM spacing. This
feature is typical of the response of long normal tools in resistive beds thinner
than the AM spacing, and is one of their main disadvantages [227].
In the conductive uninvaded bed between 47 and 57 feet, the 64 inch
normal log indicates an apparent bed thickness significantly greater than
10 feet. Likewise, in the resistive uninvaded bed between 93 and 103 feet,
it indicates an apparent bed thickness less than 10 feet. In both cases the
difference between the true and apparent bed thickness is equal to the AM
spacing of the normal [227]. This same difference exists for the 16 inch
normal logs, but it is not as noticeable because of the smaller AM spacing.
In the conductive invaded bed between 27 and 37 feet, the shallower 16-
122 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

10000
85 mm SONDE
HOLE SIZE: 6”
LATERAL
5000
8”

10”

R18’8” 12”
Rm
R16”
1000 Rm
HOLE SIZE: 6”

500 8” NORMAL

Ra
10”
Rm
12”

100

50

10
10 50 100 Rt 500 1000 5000 10000
Rm

Figure 3.53: Borehole correction chart for the 16-inch normal and 18-foot
8-inch lateral.

inch normal reads near Rxo and the deeper 64-inch normal reads closer to
Rt , as might be expected. However, in the resistive bed between 113 and 123
feet, the 16-inch normal actually reads closer to Rt than the 64-inch normal
because extreme shoulder effect has lowered the 64-inch normal log. Note
also how long it takes the 64-inch normal log to approach Rt in the shoulder
beds at zero and 150 feet because of shoulder effect.
In order to remove the influence of the conductive borehole (Rm modeled
in Figure 3.52 is 0.1 ohm-m), both logs have been borehole corrected. The
borehole correction algorithm uses the computed tool response for known
borehole sizes and for Rt /Rm contrasts from 1 to 10000. The borehole cor-
rection chart for the 16-in normal and the 18-foot 8-inch lateral is shown
in Figure 3.53 [228]. (Borehole effect for the 64-inch normal is small for
moderate Rt /Rm contrasts, and borehole correction was historically seldom
performed for this tool.) To use the chart, the raw log resistivities are en-
tered on the left and the borehole corrected resistivities are then read across
3.4. ELECTRODE (LATEROLOG) TOOLS 123

Figure 3.54: 16-inch and 64-inch normal logs in 40 foot invaded (left) and
uninvaded (right) beds.

the bottom. Although borehole correction for early tools was performed
manually using charts such as Figure 3.53, present-day modeling of early
tool response uses software implementations of interpolation algorithms or
curve fits.
The response characteristics of normal tools can be better understood
by examining computed logs in thick beds. Figure 3.54 shows computed 16-
inch normal and 64-inch normal response in 40 foot invaded and uninvaded
beds. In both the resistive and conductive invaded beds, the 64-inch normal
log now reads closer to Rt . In the uninvaded beds, both tools approach Rt
but do not read Rt exactly because the borehole correction is based on tool
response in infinitely thick beds. In the uninvaded resistive bed between
95 and 135 feet, the 64-inch normal log still has not leveled off because 10
124 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.55: Idealized electrode configuration for the lateral tool.

percent of the potential drop still comes from outside a radius of 10AM , or
53 feet (see Equation (3.25)).

3.4.2 The Lateral

The lateral device was designed to provide a deeper resistivity measurement


than the normal tools, while at the same time improving the detection of
thin beds. As shown in Figure 3.55 [220], a constant current, I, is emitted by
electrode A and returns to electrode B. The potential gradient is measured
between electrodes M and N , which are very close relative to their distances
from A (in contrast to the normal tools). Thus the lateral can be viewed
as a differential measurement. In the most common version of the tool, the
distance between A and the measure point O is 18 feet 8 inches, while M
and N are 32 inches apart. The radius of investigation is approximately
equal to the distance AO. B is usually placed downhole far above A, as is
done for the normal tools.
The difference of the potentials VM −VN is proportional to the resistivity
3.4. ELECTRODE (LATEROLOG) TOOLS 125

Figure 3.56: 18-foot 8-inch lateral log.

of the surrounding medium and yields an apparent resistivity, Ra , given by

VM − V N
Ra = KLAT , (3.28)
I
where
rM · r N
KLAT = 4π , (3.29)
rM − r N
with rM and rN being the distances between A and M , and A and N ,
respectively.
In an alternate version of the lateral, the positions of the current and
voltage electrodes are interchanged, that is, A and B are moved to M and
N , and N and M are moved to B and A. This tool is called the “inverse,”
and it records the same resistivity values as the lateral by reciprocity. The
126 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.57: 18-foot 8-inch lateral logs in 40 foot invaded (left) and uninvaded
(right) beds.

inverse arrangement made it more practical to record measurements by the


two normals and the lateral simultaneously.
Figure 3.56 shows a 18-foot 8-inch lateral computed log in the bench-
mark formation. The unsymmetrical nature of the lateral response is very
apparent. The lateral logs are characterized by a fairly sharp peak near the
lower boundary of each bed which can be used to identify the presence of the
bed. However, the 10 and 3 foot beds in this formation are too thin for the
lateral to be able to resolve individual bed resistivities. Because the volume
of investigation of the lateral in this formation includes up to 6 beds, it is
difficult to interpret the effect of the two invaded zones. Note that in the
resistive invaded bed between 113 and 123 feet, the lateral actually reads
closer to Rt than in the uninvaded bed between 93 and 103 feet.
3.4. ELECTRODE (LATEROLOG) TOOLS 127

The chart in Figure 1.3 illustrates changes in the characteristic shape of


lateral logs in thin beds of various sizes. It can be seen from this chart that
lateral interpretation becomes extremely difficult in sequences of thin beds.
Figure 1.3 indicates that lateral interpretation becomes “straightforward”
only as the bed thickness approaches the distance 2AO, or approximately
40 feet.
Figure 3.57 shows computed 18-foot 8-inch lateral response in 40 foot
invaded and uninvaded beds. The effect of the shallow invasion on the center-
bed apparent resistivity readings is very small; invasion merely deflects the
log in the direction of Rxo near bed boundaries. Even in the relatively thick
40 foot uninvaded beds, the only place that the log approaches Rt is directly
above each bed boundary. Lateral tools have “anomalous zones” extending
a distance of approximately AO below each bed boundary which complicate
interpretation [227]. In order to correctly interpret lateral logs, a knowledge
of typical curve shapes obtained from modeling is extremely helpful.

3.4.3 Russian BKZ tools

Although focused laterolog tools are available in the former Soviet Union,
the most commonly run resistivity devices by far are unfocused normals
and laterals. Perhaps 80 percent of all resistivity logs are of the unfocused
type [145]. Western companies are finding it necessary to learn more about
the forgotten art of interpreting unfocused electrode tool response in order
to make enlightened investment decisions.
A brief history of well logging in the former Soviet Union helps to un-
derstand why the older normal and lateral tools have remained popular for
so many years. Conrad and Marcel Schlumberger’s company, Societé de
Prospection Électrique, introduced wireline electrical logging in the Soviet
Union. The first Russian well log was recorded by Schlumberger engineer
Raymond Sauvage in 1929 in the Grozney field north of the Caucasus moun-
tains [107]. Geologic conditions there as well as in the subsequently explored
Baku fields were well-suited to the early normal and lateral sondes, i.e., thick
sandstone beds interspersed with high contrast correlatable shales [252]. The
Soviet oil industry was very receptive to the emerging logging technology be-
cause it helped create prosperity for their new political system. The Schlum-
berger brothers were granted government contracts to train engineers and
manufacture logging equipment in Russia. Logging flourished there until
1937, when the Schlumberger crews were permanently expelled because of
128 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

the perceived political threat of foreigners [130]. This expulsion began the
near isolation of the Russian wireline industry that lasted for more than half
a century.
In the years that followed, Russian engineers concentrated their efforts
on improving the performance of multiple spaced lateral measurements (L.
M. Alpin and V. N. Dakhnov were the most notable pioneers). Laterals
were preferred over normals because the vertical resolution of a lateral tool
deteriorates less when spacings are increased to obtain deeper readings. The
generic name for Russian lateral measurements is BKZ (BKZ), which trans-
lates from Cyrillic as “lateral logging sounding” [252]. (Schlumberger influ-
ence prevails in the Russian word for logging, “karotazh”, which was derived
from the French word for coring, “carottage.”)
The present-day BKZ suite of measurements consists of up to five laterals,
one inverted lateral and one normal. Electrode spacings vary depending on
local conditions and needs. Table 3.10 shows the electrode configurations for
five of the most common BKZ laterals [145]. The two most common normal
sondes run in combination with the laterals have AM spacings of 10 inches
(0.25 meters) and 16 inches (0.4 meters).

EL04 EL10 EL22 EL42 EL85


A -18 -41 -89 -167 -335
M -2 -2 -10 -10 -20
O 0 0 0 0 0
N 2 2 10 10 20

Table 3.10: Electrode locations (in inches) for the most common BKZ later-
als.

Figure 3.58 shows computed BKZ lateral response in the benchmark for-
mation. Logs for the five most common sondes are on the left, and “top” and
“bottom” versions of the shallowest lateral are on the right. The standard
or bottom lateral has the paired potential electrodes positioned below the
current source. The inverted or top lateral has the paired potential elec-
trodes positioned above the current source. Running both top and bottom
versions in combination gives a clearer identification of bed boundaries, as
Figure 3.58 illustrates. The suite of five laterals delineates the beds more
clearly than the single 18-foot 8-inch lateral (see Figure 3.56) because most
of the BKZ measurements have shorter M N spacings. In the conductive in-
3.4. ELECTRODE (LATEROLOG) TOOLS 129

Figure 3.58: Logs of the five most common BKZ laterals (left) and the
shallowest top and bottom lateral (right) in the benchmark formation.

vaded bed between 27 and 37 feet, the separation of the lateral curves gives
a slight indication that invasion is present. However, in the resistive invaded
bed between 113 and 123 feet, shoulder effect masks the invasion.
Figure 3.59 shows computed BKZ lateral response in 40 foot invaded and
uninvaded beds. The presence of invasion is much more apparent in these
thicker beds, as shown by the systematic separation of the curves on the
left. However, the unsymmetrical nature of the lateral curves makes them
extremely difficult to interpret. In order to interpret BKZ logs visually,
Russian logs analysts have developed complex systems of rules [145] which
employ charts of computed departure curves to account for borehole effect,
130 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.59: BKZ lateral logs in 40 foot invaded (left) and uninvaded (right)
beds.

invasion and bed thickness. Manual interpretation methods such as this


are very time intensive and not readily applicable to long sections of logs.
Chart based interpretation has recently been computer automated both in
the former Soviet Union and in the West. Other recently developed methods
of estimating Rt and Rxo from BKZ logs include iterative forward modeling
and joint inversion of multiple curves.

3.4.4 Laterolog 7 (LL7)

From the previous description of normal and lateral measurements, it is clear


that these tools did not always give an accurate estimate of the formation
3.4. ELECTRODE (LATEROLOG) TOOLS 131

Figure 3.60: Idealized electrode con- Figure 3.61: Computed current pat-
figuration for LL7, showing the use terns and equipotential surfaces for
of an insulated bridle. LL7 in a thick uninvaded bed.

resistivity. In particular, these early devices were affected by large contrasts


between the borehole mud and the formation resistivity and by invasion,
and could not resolve thin beds. It became evident that a means to “focus”
the survey current was needed. The introduction of focusing for electrode
tools is attributed to Henri Doll [95]. The first commercial focused tool
was the Laterolog 7 or LL7 [91] (“laterolog” because the survey current
was focused laterally, and “7” because there were seven electrodes). The
electrode configuration and computed current patterns for LL7 are depicted
in Figure 3.60 and Figure 3.61, respectively. Note the horizontal trajectory of
the LL7 focused survey current in comparison to the normal survey current
shown in Figure 3.51. Table 3.11 shows the positions of the LL7 electrodes.
The LL7 is comprised of three current electrodes, A0 , A1U and A1L ,
and four voltage electrodes, M1U , M2U , M1L and M2L , called monitoring
132 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

electrodes. (“U” and “L” denote upper and lower. The notation A1 and
A 1 is sometimes used instead.) Electrodes A1U and A1L drive currents
into the formation which focus the current beam emitted by the center A0
electrode. The degree of survey current focusing can be altered by varying
the strength of these auxiliary currents, often referred to as bucking currents.
If the bucking currents are too small, the beam will be under focused and
very little will be gained over a normal device. At the other extreme, if the
bucking currents are too large, very little current will be emitted from the
survey current electrode A0 and the beam will be over focused.

Position
A1U -40
M2U -20
M1U -12
A0 0
M1L 12
M2L 20
A1L 40

Table 3.11: LL7 electrode locations (in inches).

In practice, the strength of the bucking currents is controlled by a feed-


back loop with sufficient gain to ensure that the potential gradient measured
between a pair of monitoring electrodes is null. The monitoring conditions
for LL7 are

VM1U − VM2L = 0, (3.30)


VM2U − VM1L = 0. (3.31)

These conditions impose that no currents are flowing in the vertical di-
rection in the vicinity of the monitoring electrodes. Thus the A0 survey
current enters the formation horizontally within the area bounded by the
monitor electrodes. Using the above conditions which involve monitoring
electrodes on either side of A0 allows separate control of both bucking cur-
rents and produces a relatively symmetrical response when the tool crosses a
bed boundary. (An earlier version of LL7 imposed a null potential gradient
between monitoring electrode pairs on the same side of A0 .) For a given
tool length, Doll found that the optimum focusing condition was obtained
3.4. ELECTRODE (LATEROLOG) TOOLS 133

Figure 3.62: Borehole correction chart for LL7.

when the spread ratio computed from the distances between electrodes has
the value of
A0
A1L + A0
A1U
 2.5 , (3.32)
A0M1U + A0M2U
where A 
0 Aj and A0 Mi denote distance.
Ra for LL7 is computed from the ohmic drop to the current emitted from
A0 between equipotential surfaces passing through the monitoring electrodes
and the reference potential electrode N ,
(VM1U + VM2U )/2 − VN
Ra = KLL7 . (3.33)
IA0

Since the N electrode is located relatively far from A0 , VN is usually assumed


to be negligible.
KLL7 is normally taken to be 1.38 [224]. The borehole correction chart
for LL7 is shown in Figure 3.62 [229]. For LL7 and most other laterolog tools,
the value of K is adjusted so that the tool response in an 8 inch borehole
crosses the Rcorr /RLL = 1 line between RLL /Rm = 10 and 100 (i.e., the
correction is unity). As a result, borehole correction only becomes necessary
for large boreholes and high RLL /Rm contrasts.
An examination of the conditions which must be met to compute LL7
response serves to illustrate the general method for calculating the response
of all focused laterolog tools. There are a total of seven electrodes, three
are current electrodes which emit (or receive) current and four are voltage
electrodes which are supposed to be passive or connected to infinite input
134 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

impedance electronics. Ten conditions are therefore necessary to solve for


the ten unknowns. In order to simplify calculations, the concept of transfer
impedance is introduced. If we designate Vi as the potential on electrode i
with respect to infinity, and Ij as the current emitted by electrode j, then
the transfer impedance Zij between electrodes i and j is defined as
Vi = Zij Ij . (3.34)
Thus Zij can be thought of as the ratio of the voltage on electrode i to the
current from electrode j which generated that voltage. The computation
of focused tool response can also be conceptualized as forming a particular
combination of normals [132].
The linear system of equations describing the LL7 operating conditions
can be written in the matrix form as
IA0 0 0 0 0 0 0 0 0 0 −1
0 0 0 0 0 0 VM1U 0 0
−VM2L 0
0 0 0 0 0 0 0 VM2U −VM1L 0 0
ZA0 A0 IA0 ZA0 A1U IA1U ZA0 A1L IA1L −VA00 0 0 0 0 0 0
ZA1U A0 IA0 ZA1U A1U IA1U ZA1U A1L IA1L 0 −VA1U 0 0 0 0 0 0
ZA1L A0 IA0 ZA1L A1U IA1U ZA1L A1L IA1L 0 0 −VA1L 0 0 0 0 0
ZM1U A0 IA0 ZM1U A1U IA1U ZM1U A1L IA1L 0 0 0 −VM1U 0 0 0 0
ZM2U A0 IA0 ZM2U A1U IA1U ZM2U A1L IA1L 0 0 0 0 −VM2U 0 0 0
ZM1L A0 IA0 ZM1L A1U IA1U ZM1L A1L IA1L 0 0 0 0 0 −VM1L 0 0
ZM2L A0 IA0 ZM2L A1U IA1U ZM2L A1L IA1L 0 0 0 0 0 0 −VM2L 0
= 0. (3.35)
The above linear system is generated from two sub-systems. The first three
rows of the matrix describe how the device is operated. A unit current
is emitted by electrode A0 (IA0 = 1). There is a short circuit between
electrode M1U and M2L (VM1U − VM2L = 0). There is also a short circuit
between electrode M2U and M1L (VM2U − VM1L = 0).
The last seven rows of the matrix relate the potential of each electrode to
the currents through their respective transfer impedances. In practice, any
2D or 3D modeling code which solves the basic Poisson equation can be used
to compute the transfer impedance for each current–voltage electrode pair-
ing. Focused tool response is calculated by solving the above system of linear
equations for the unknown current (Ij ) and voltage (Vi ) values. The appro-
priate current and voltage values are then substituted in Equation (3.33) to
generate the apparent resistivity read by the tool.
3.4. ELECTRODE (LATEROLOG) TOOLS 135

Figure 3.63: LL7 log.

Instead of the above dynamic hardware focusing, software focusing is


sometimes implemented for modern tools. In software focusing the transfer
impedances Zij are first measured sequentially by energizing each electrode
with a current of unit strength. Then the numerical values are digitized and
sent to the surface for processing. Finally Ra is computed from an expression
which is only a function of the transfer impedances.
Figure 3.63 shows a LL7 computed log in the benchmark formation.
Because there is very little borehole effect on LL7 in 8 inch holes (see Fig-
ure 3.62), the log was not borehole corrected. The vertical resolution of the
LL7 log is greatly improved compared to the unfocused normal (Figure 3.52)
and lateral (Figure 3.56) logs. In the uninvaded beds, LL7 reads near Rt in
all but the 3 foot resistive bed. In general, LL7 has very little shoulder effect
in beds which are thicker than the distance between the midpoint of the up-
136 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.64: Idealized electrode con-


figuration for LL3. G are guard elec- Figure 3.65: Computed current pat-
trodes, M is the A0 measure elec- terns and equipotential surfaces for
trode and S is a low resistance shunt. LL3 in a thick uninvaded bed.

per and lower measure electrode pairs (32 inches) [91]. In both the resistive
and conductive invaded beds, LL7 reads closer to Rxo than to Rt . Thus of
all the usual environmental effects, only invasion significantly influences LL7
response.

3.4.5 Laterolog 3 (LL3)

The basic LL3 design concept was first proposed in the 1920’s by Conrad
Schlumberger. However, it took until 1950 for a commercial version of the
tool to be developed by Henri Doll [94]. The LL3 consists of a short central
electrode which emits the survey current, surrounded by two larger symmet-
rical electrodes called guard electrodes. For this reason, LL3 is sometimes
3.4. ELECTRODE (LATEROLOG) TOOLS 137

Figure 3.66: Borehole correction chart for LL3.

called the ‘Guard Log’. The guard electrodes range from 3 to 5 feet in length,
with 5 feet being the most common size. These long metallic guard electrodes
are not without inconvenience; they are prone to surface impedance, prevent
the recording of SP and cannot be interlaced with induction measurements.
The LL3 electrode configuration is shown in Figure 3.64 [91]. The three
electrodes are short-circuited together and connected to a current source
whose return is located on a remote uphole electrode. In effect, the three
separate electrodes become a single current-emitting cylinder with region
near the tool maintained at a quasi-constant potential. This type of focusing
is termed passive focusing. In actuality, the potential between electrodes is
usually monitored and maintained by adjusting the measure current in order
to avoid problems with electrode surface impedance. Thus passive focusing
is not as passive as the term indicates. Ra is obtained by taking the ratio of
the voltage measured close to the A0 electrode and the current emitted by
A0 ,
VA
Ra = KLL3 0 . (3.36)
IA0

KLL3 is normally taken to be 0.50 [224]. The borehole correction chart


for LL3 is shown in Figure 3.66 [229].
Computed LL3 current patterns and equipotential surfaces are depicted
in Figure 3.65. Note that the region adjacent to the three electrodes is at
approximately the same potential. The central survey current beam enters
the formation laterally in comparison to the normal tool (Figure 3.51), but it
is not as sharply focused as LL7 (Figure 3.61). For this reason, passively fo-
138 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.67: LL3 log.

cused tools generally have slightly greater borehole effect and slightly poorer
vertical resolution than actively focused tools. Table 3.12 shows the positions
of the LL3 electrodes for the most common version of the tool.
Figure 3.67 shows a LL3 computed log in the benchmark formation.
Borehole correction cannot be neglected for LL3, and the log in this figure
has been borehole corrected using a software algorithm that performs inter-
polation using the data plotted in Figure 3.66. A comparison of Figure 3.67
and the corresponding log for LL7 (Figure 3.63) shows that shoulder bed
effect is greater for LL3 in both the resistive and conductive beds. This is
is a consequence of the fanning out of the passively focused LL3 A0 survey
current. Even though LL3 and LL7 have similar depths of investigation (see
Figure 3.68), LL3 reads closer to Rxo than LL7 in the invaded beds because
of shoulder bed effect.
3.4. ELECTRODE (LATEROLOG) TOOLS 139

Position
A1U -60 to -8
A0 -4 to 4
A1L 8 to 60

Table 3.12: LL3 electrode positions (in inches).

Figure 3.68: Pseudo-geometrical factors showing relative depths of investi-


gation of the obsolete laterolog tools LL7, LL3 and LL8.

3.4.6 Laterolog 8 (LL8)

Section 3.2.3 describes how the Dual Induction tool was developed for use
in fresh drilling muds, but required an additional shallow electrode-type
measurement which was physically interlaced on the induction mandrel. The
first focused resistivity device combined with the Dual Induction tool was
named Laterolog 8 (LL8) [97] because it had a total of eight electrodes. In
order to minimize the influence of the metallic electrodes on the induction
measurement, the LL8 used thin ring electrodes which were placed in the
low sensitivity region of the induction sonde. The LL8 design was a modified
version of LL7 with two current returns on the sonde body above the main
electrode array as shown in Figure 3.69. The LL8 currents are noticeably
unsymmetrical because the current returns are located fairly close to A0 .
The reference potential electrode is located on the insulated bridle above
the tool (see Figure 3.60), relatively far from A0 .
140 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.69: LL8 electrode configuration, with current patterns indicated.

The monitoring conditions for LL8 are

IA0 = IA0 return , (3.37)


IA1U + IA1L = IA1 return , (3.38)
VM1U − VM2L = 0, (3.39)
VM2U − VM1L = 0. (3.40)

Like LL7, an earlier version of LL8 imposed a null potential gradient between
monitoring electrode pairs on the same side of A0 , as shown in Figure 3.69.
Figure 3.69 also gives the locations of the LL8 electrodes (in inches). Ra for
LL8 is
(VM1U + VM2U )/2
Ra = KLL8 . (3.41)
IA0
3.4. ELECTRODE (LATEROLOG) TOOLS 141

Figure 3.70: Borehole correction chart for LL8.

Figure 3.71: LL8 log.


142 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.72: Schematic diagram of the Dual Laterolog electrode configura-


tion and current patterns. (The sonde is split for illustrative purposes only;
LLS and LLD currents are axisymmetric.) io denotes survey current and ia
bucking current. The MicroSFL is located on the lower (A2 ) electrode.

KLL8 is normally taken to be 0.87 [224]. The borehole correction chart


for LL8 is shown in Figure 3.70 [229].
Figure 3.71 shows a LL8 computed log in the benchmark formation. Note
that the LL8 log is slightly unsymmetrical. One of the major modifications
made to LL7 in designing LL8 was to decrease the spacings between elec-
trodes in order to provide a shallow measurement of the invaded zone for the
Dual Induction tool. As a result, the shallower LL8 log reads closer to Rxo
than the LL7 log (Figure 3.63) in the two invaded beds. The shorter LL8
spacings also cause shoulder bed effect to be less for LL8 than for LL7, as
is evident in the thin resistive bed between 80 and 83 feet. Because LL8 is
such a shallow tool, its readings are influenced by borehole effect, and LL8
logs should be borehole corrected when hole diameters are larger than 10
inches.
3.4. ELECTRODE (LATEROLOG) TOOLS 143

3.4.7 The Dual Laterolog tool (DLT)

The ultimate purpose of resistivity logging is to determine hydrocarbon sat-


uration from the true formation resistivity, Rt . Unfortunately, it has been
impossible (so far) to design a single deep-reading measurement of Rt which
is entirely free of the effects of the invaded zone. In the 1960’s, the solution
to the problem of invasion effect for laterlogs was the same as the solu-
tion for induction tools described in Section 3.2.3, namely to design three
separate tools with different depths of investigation. With three measure-
ments one could then solve for the formation and invasion resistivities (Rt
and Rxo ), and the invasion diameter (di ), assuming negligible or easily cor-
rectable shoulder-bed effect. The descriptions of early electrode tools have
shown that there is an additional consideration for laterologs; borehole effect
cannot be neglected or corrected simply as in the case of induction.
As a result, the strategy in designing the Dual Laterolog tool [148, 235,
245] or DLT was to devise a tool combination which had: (1) little bore-
hole effect, (2) good vertical resolution and (3) three well-distributed radial
depths of investigation. A microdevice on a pad, the MicroSFL (described
in Section 3.5) was used to give an accurate estimate of Rxo and to delineate
bed boundaries. With Rxo known, a dual depth of investigation laterolog
tool was then optimized to determine the remaining two unknowns, di and
Rt . The electrode configuration for the two DLT arrrays is shown in Fig-
ure 3.72 [245]. There are two independent measurements: a shallow depth of
investigation measurement LLS and a deep measurement LLD. Both mea-
surements use the same electrodes and have the same survey current beam
thicknesses, but different focusing methods are used to provide two different
depths of investigation.
Figure 3.73 shows computed current patterns for LLD and LLS in a
borehole penetrating an infinetely thick homogeneous formation. The addi-
tional orthogonal equipotential surfaces aid in the visualization of similarities
and differences between the shallow and deep focusing modes. Table 3.13
shows the positions of the DLT electrodes for the most recent version of the
Schlumberger tool, called DLT-E. Although laterologs are considered to be
DC tools, zero frequency measurements are impractical, mainly because of
noise rejection problems. Therefore, the DLT operates at a small but finite
frequency: 35 Hz for LLD and 280 Hz for LLS.
In order to achieve a deeper depth of investigation than previous tools,
the deep mode uses dynamic beam focusing introduced with the LL7 com-
144 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Position
A2U -174.5 to -69.5
M3U -39.3
A1U -38.8 to -24.5
M2U -16.0
M1U -10.0
A0 -4.5 to 4.5
M1L 10.0
M2L 16.0
A1L 24.5 to 38.8
M3L 39.3
A2L 69.5 to 174.5

Table 3.13: DLT electrode positions (in inches). The measure electrodes
(M ) are thin rings.

bined with the LL3 concept of long guard electrodes. As can be seen in
Figure 3.73, the current electrodes A1U , A1L , A2U and A2L are set at almost
the same potential and emit bucking currents into the formation surrounding
the tool. The magnitude of the A0 survey current is controlled in such a way
as to ensure that the average vertical potential gradient measured between
the monitoring electrodes M1U , M1L , M2U and M1L is null. This condition
forces the survey current beam to be well-focused into any bed adjacent to
the A0 electrode. The nominal thickness of the survey current beam is 2
feet. The equations enforcing the above LLD monitoring conditions are

VA1U = VA1L , (3.42)


VA2U = VA2L , (3.43)
VM1U + VM1L = VM2U + VM2L , (3.44)
VA2U + VA2L = VM3U + VM3L . (3.45)

The additional M3U , M3L electrode pair is used to further ensure that a
uniform potential gradient is maintained between the bucking electrodes
when there are high resistivity contrasts between beds. Ra for LLD is
(VM1U + VM1L )/2
Ra = KLLD . (3.46)
IA0
with KLLD equal to 0.89 [224].
3.4. ELECTRODE (LATEROLOG) TOOLS 145

Figure 3.73: Computed current patterns and equipotential surfaces for LLD
(a) and LLS (b) in a thick uninvaded bed. (Electrode locations and currents
are drawn to scale.)

All emitted currents are returned to a B electrode located at the surface.


The reference potential electrode, N , is located about 80 feet above the
sonde at the top of the insulated bridle which is used to support the weight
of the tool and provide electrical communications between the downhole
and surface electronics. The current patterns for LLD are quite similar
to those for LL3 (Figure 3.65). The LLD equipotential surfaces remain
fairly cylindrical for a long distance up and down the borehole because the
total length of the DLT is 28 feet, which is more than double that of LL3.
Electrically connecting the LLD A1 and A2 electrodes results in a pair of 13
foot “guard” electrodes.
In the shallow mode, the only difference consists of using electrodes A2U
and A2L as returns for currents emitted from electrodes A0 , A1U and A1L .
This type of focusing is called pseudo-laterolog because the current is re-
146 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

turned to nearby electrodes instead of to a remote electrode. Since the


current returns are so close, the survey current diverges quickly when it en-
ters the formation, resulting in a shallow depth of investigation. There is
not a great deal of borehole effect, as can be seen in Figure 3.73, because
the A2 current return electrodes are quite long. (Figure 5.5 in Chapter
5 shows the borehole correction chart for LLS.) The magnitude of the A0
survey current is controlled so that the average vertical potential gradient
measured between the two M1 , M2 monitoring electrode pairs is null, in the
same manner as LLD. As a result, the LLS survey current beam thickness is
approximately the same as that of LLD, as can be seen in Figure 3.72 and
Figure 3.73. For this reason, LLS and LLD have approximately the same
vertical resolution. The equations enforcing the LLS monitoring conditions
are

VA1U = VA1L , (3.47)


VA2U = VA2L , (3.48)
VM1U + VM1L = VM2U + VM2L , (3.49)
IA0 + IA1U + IA1L + IA2U + IA2L = 0. (3.50)

Ra for LLS is
(VM1U + VM1L )/2
Ra = KLLS . (3.51)
IA0
with KLLS equal to 1.45 [224]. LLS uses the same N reference potential
electrode as LLD.
In order to achieve satisfactory accuracy in both very high and very low
resistivity formations where the DLT was designed to run, a constant power
measurement system was developed. Previous laterolog tools held the survey
current constant and detected variations in voltage. This approach is most
accurate in high-resistivity formations. In the DLT system, both the A0
survey current and the voltage at A0 are measured and the product of the
two (i.e., power) is held constant. This allows the DLT to have a response
range of 0.1 to 40,000 ohm-m [245], much wider than the ranges of previous
tools.
Figure 3.74 shows LLD and LLS computed logs in the benchmark forma-
tion. Also shown is a computed MicroSFL log, which closely follows the Rxo
curve. The LLS and LLD logs were not borehole corrected since correction
is only necessary for large hole sizes. In the uninvaded bed between 47 and
57 feet, LLD departs from Rt because the survey current flows preferentially
3.4. ELECTRODE (LATEROLOG) TOOLS 147

Figure 3.74: Dual Laterolog log.

in the conductive bed (squeeze effect), while much of the bucking current
remains in the resistive shoulders as it flows to the remote return.
Both LLS and LLD logs have significantly better vertical resolution than
LL7 (Figure 3.63). Nevertheless, it is evident that it is still possible to make
substantial errors estimating Rt in invaded thin beds. Even though LLD
was designed to be deeper than existing tools, it reads only slightly closer to
Rt than LL7 (Figure 3.63) or LL3 (Figure 3.67) in the invaded beds. The
LLD deep focusing was optimized to produce a large separation between
LLS and LLD curves in infinitely thick beds. However, when shoulder bed
effect occurs in combination with invasion, it can cause LLD to read much
closer to Rxo in thin beds than in the infinitely thick bed limit.
Figure 3.75 shows pseudo-geometrical factors illustrating how much shal-
148 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.75: Pseudo-geometrical factors for LLS and LLD in an infinitely


thick bed and in a 10 foot bed showing how the relative depth of investigation
of LLD changes.

lower LLD effectively becomes in a 10-foot bed. In infinitely thick beds the
LLS and LLD curves are widely separated for all invasion radii, while the
curves only separate between invasion radii of 20 to 80 inches in a 10-foot
bed. Although interpretation in thin invaded beds can be performed by se-
quentially applying charts for bed thickness correction (see Figure 5.8) and
invasion correction (see Figure 5.10), chartbook methods are obviously inac-
curate because invasion correction is based on the infinitely thick bed limit.
Iterative forward modeling or carefully constrained inversion can provide a
more accurate value for Rt in cases such as Figure 3.74.
Two log interpretation anomalies often mentioned in connection with
LLD are Delaware effect and Groningen effect. They are caused by a break-
down of the assumptions that “the current return is considered to be at
infinity” and “the reference electrode is assumed to be at zero potential,”
under extreme conditions.
Early LLD logs had erroneously high readings below very resistive beds.
This error was called Delaware effect, named after the west Texas formation
where it was first observed. The close proximity of the B return electrode
to the N reference electrode on the bridle in the first version of the DLT
(see Figure 3.60) caused a negative potential to be generated at N when the
returning currents were confined within the borehole. When VN cannot be
3.4. ELECTRODE (LATEROLOG) TOOLS 149

Figure 3.76: LLD log with in- Figure 3.77: LLD AC current patterns
creasing resistivity gradient in a in the presence of Groningen effect; the
conductive bed below a resistive current flows to the tool via the armored
bed, characteristic of Groningen cable and returns to the surface through
effect. the casing.

neglected, Equation (3.46) becomes


(VM1U + VM1L )/2 − VN
Ra = KLLD . (3.52)
IA0
The spurious negative reference potential caused an increase in the above
numerator, which in turn increased the recorded resistivity. Moving B to the
surface widened the distance between B and N and significantly reduced the
problem under normal logging conditions. In addition, the LLD frequency,
which was initially the same as LLS (280 Hz), was lowered to 35 Hz to
decrease AC effects caused by the longer distance between B and N .
Groningen effect is named after the large Dutch gas field where it was first
observed. Even after correcting for Delaware effect, an unexpected increase
150 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

in LLD apparent resistivity still occurred in low-resistivity (around 1 ohm-m)


reservoirs below very resistive beds. LLD logs would systematically increase
for a distance of 100 to 200 feet as the tool approached the lower boundary
of the resistive bed. Comparisons with LLS and MicroSFL showed that very
little or no invasion was present. Figure 3.76 [220] shows a log with typical
Groningen effect error. The effect was very pronounced when casing was set
inside the resistive bed, with errors sometimes reaching 10 ohm-m. When
there was no casing present, errors were less than 1 ohm-m. The overly
high LLD readings caused unjustified optimism in estimating the amount of
hydrocarbons present [270].
Theoretical work was carried out in the late 1970’s which clarified the
physical causes of Groningen effect [166]; the major factor was found to be
skin effect. With B at the surface, the LLD currents often had to travel 1
to 2 miles to reach the current return. At 35 Hz and 1 ohm-m, the skin
depth, δ, is 280 feet, which is small in comparison. Because of skin effect,
the current returning to B is constrained to remain within a cylinder of
radius of δ around the cable carrying current down to the tool, effectively
forming a coax. This confinement of the current around the cable creates an
additional AC impedance which in turn generates a negative potential at N
and distorts the apparent resistivity reading.
The presence of casing in the high-resistivity shoulder increases the mag-
nitude of Groningen effect. The casing acts as an electromagnetic screen
around the armored cable, and the current is channeled to the surface inside
the casing. Figure 3.77 [166] shows the path that the current takes. Near the
casing shoe (lower extremity), the current reaches the outside of the casing
over the characteristic length of the casing [166], Lcasing (950 feet for steel
casing in a 10 ohm-m formation). It then flows down to the casing shoe, pen-
etrating to only a distance of δcasing (typically 0.1 inch), where it reverses
direction and flows upward on the inside of the casing within δcasing . Thus
the casing enhances the negative potential that affects N . The gradient on
the resistivity log is caused by the changing distance between N and the
casing shoe as the tool moves up the borehole.
A method was devised to detect and correct Groningen effect [167]. It
is based on the fact that the potential has in-phase and out-of-phase com-
ponents. The apparent resistivity is derived from the in-phase potential.
In the presence of Groningen effect, the out-of-phase potential shifts from
its base-line value. This phase shift is monitored and used to correct the
apparent resistivity.
3.4. ELECTRODE (LATEROLOG) TOOLS 151

Figure 3.78: Changes made to DLT in designing HALS (only the top portion
of the symmetrical sonde is shown). On the left is the tool scaled in half
proportionally. On the right is the tool after optimization.

Groningen effect cannot be modeled with the DC laterolog codes that are
normally used. In the late 1980’s, an AC finite element code was constructed
especially for modeling Groningen effect [177].
In the early 1990’s, a modified version of the DLT was introduced with
LLD and LLS arrays that were half the length of the traditional DLT. The
tool was called the High Resolution Azimuthal Laterolog Sonde (HALS) [236]
and was part of a new tool combination called Platform Express. The shorter
HALS was easier to combine with other tools than the cumbersome 28-foot
DLT. It also provided higher vertical resolution and made it possible to
negotiate horizontal wells with small radius of curvature. In addition, the
152 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.79: HALS log.

HALS included an azimuthal resistivity measurement, the ARI (see Section


3.8.6) on the A0 electrode.
In designing HALS, the DLT electrode positions were first scaled down
by a factor of 0.5 and then optimized as shown in Figure 3.78. The environ-
mental effects that were improved by optimization are noted on the figure
because this information is useful for the general design of laterolog tools.
The depth of investigation of HLLD is shallower than that of LLD because
of the shorter length of the HALS tool. The shorter length also makes HLLD
more sensitive to shallow invasion than LLD.
Figure 3.79 shows a HALS log in the benchmark formation. In the con-
ductive and resistive 3-foot beds, the apparent bed thicknesses indicated by
the HALS logs are greater than those of the DLT logs (Figure 3.74), illus-
3.4. ELECTRODE (LATEROLOG) TOOLS 153

trating the improved vertical resolution of HALS. In the invaded beds, the
HALS logs give the same shallow and deep resistivity values as the DLT logs.
This is not surprising, since the deep depth of investigation provided by the
long over-all length of LLD decreases in thin beds (see Figure 3.75).

3.4.8 The Spherically Focused Log (SFL)

The Spherically Focused Log (SFL) [231, 244, 233] was designed in the early
1970’s as a replacement for 16-inch normal and LL8 devices. LL8 was used
for almost 20 years to provide the Dual Induction tool with an additional
shallow resistivity measurement for the interpretation of invasion, and to
supplement the limitations of the induction tools in delineating beds thinner
than 4 feet. However, LL8 had two major operational problems: (1) borehole
effect was large for hole sizes greater than 10 inches, and (2) like the DLT,
LL8 required an 80-foot bridle, which was awkward to use and prone to
reference electrode effects.
In order to overcome these problems, the focusing system for the SFL
was designed to be different [231] than the focusing used by previous elec-
trode devices. While LL7 and other laterolog-type systems attempt to focus
the survey current in the shape of a planar disc (see Figure 3.61), the SFL
monitoring conditions establish quasispherical equipotential shells around
the A0 survey current electrode. The SFL electrode configuration and ide-
alized current patterns are shown in Figure 3.80 [233]. Table 3.14 shows the
positions of the SFL electrodes.

Position
M2U -56.5
M1U -46.5
A1U -15.0
M0U -9.0
A0 0.0
M0L 9.0
A1L 15.0
M1L 46.5
M2L 56.5

Table 3.14: SFL electrode locations (in inches).


154 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.80: Schematic diagram of the SFL electrode configuration and cur-
rent patterns. io (dashed) is the survey current and ia (solid) is the bucking
current. The measure voltage is the drop between the equipotential surfaces
B and C.

The tool is comprised of nine electrodes on the sonde and a current return
located on the armored cablehead over 20 feet above the tool. This current
return can be assumed to be at infinity for all practical purposes. There are
three current electrodes, A0 , A1U and A1L , two measure potential electrodes,
M0U and M0L , and four monitoring potential electrodes, M1U , M2U , M1L
and M2L .
The current emitted from A0 consists of two parts: the survey current
which travels through the formation to the return on the cablehead, and
the bucking current which returns to A1U and A1L (the two bucking cur-
rent returns are short circuited together). The magnitude of the currents is
controlled by a feedback loop that imposes a null between the monitoring
electrode pairs M1U and M1L , and between M1U and M1L . This bucking cur-
rent system serves to block the flow of the survey current within the borehole
and establishes equipotential spheres. The measure voltage is the difference
3.4. ELECTRODE (LATEROLOG) TOOLS 155

Figure 3.81: Computed current Figure 3.82: Computed current pat-


patterns and equipotential surfaces terns and equipotential surfaces for re-
for SFL in a thick uninvaded bed. ciprocal SFL in a thick uninvaded bed.

between the average potential at the inner M0 electrode pair and the average
potential at the outer two M1 and M2 electrode pairs. In Figure 3.80 this is
shown as the voltage drop between the equipotential spheres passing through
the electrodes. In addition, a constant potential is maintained between these
two spherical surfaces. Since the voltage drop is constant, the intensity to
the survey current is proportional to the conductivity of the volume of the
formation between the two spherical surfaces.
The equations enforcing the above monitoring conditions are

VA1U = VA1L , (3.53)


VM1U + VM1L = VM2U + VM2L , (3.54)
200IM0U + VM0U = VM0L . (3.55)
156 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

The apparent resistivity, Ra , for SFL is

(VM0U + VM0L )/2 − (VM2U + VM1U + VM1L + VM2L )/4


Ra = KSF L , (3.56)
IA0 + IA1U + IA1L

with KSF L equal to 2.13.


The SFL design provides a good example of the principle of reciprocity as
applied to electrode tools. Although SFL focusing and current distribution
is different from that of laterologs, the response of both types of tools is
quite similar because laterologs and SFL-type tools are reciprocals of each
other. The principle of reciprocity states that the role of current and voltage
electrodes can be exchanged. Thus the focusing current is the reciprocal of
the monitor voltage and conversely, and the measure current is the reciprocal
of the measure potential and conversely. This also implies that the formation
transfer impedance matrix is symmetrical, that is, Zij = Zji . Like LLS, SFL
borehole correction is only necessary for large boreholes and high Rt /Rm
contrasts.
Figure 3.81 shows computed current patterns for SFL and Figure 3.82
shows computed current patterns a reciprocal SFL. The reciprocal SFL is
similar to a LLS (see Figure 3.73) where the center electrode is split and
the monitoring implemented in a slightly different manner. Therefore it
is not surprising that similar performance can be achieved with these two
seemingly different devices.
Figure 3.83 shows a SFL log in the benchmark formation. The vertical
resolution and depth of investigation of SFL are much like that of the LL8
(see Figure 3.71), which it replaced. The SFL response to individual beds is
quite symmetrical in comparison to LL8.

Demonstrating electrode tool physics with SFL current patterns

For the SFL and for all electrode tools in the benchmark formation, the
apparent thickness of conductive beds is greater than the apparent thickness
of resistive beds. Because currents often travel large distances from their
source to return electrodes, adjacent beds can have a considerable effect on
the apparent resistivity reading in a particular bed of interest. Figure 3.84
illustrates the manner in which current and voltage patterns are affected by
the presence of conductive or resistive shoulders for the 10-foot uninvaded
beds in Figure 3.83. When the surrounding shoulders are more conductive
3.4. ELECTRODE (LATEROLOG) TOOLS 157

Figure 3.83: SFL log.

than the bed of interest as in Figure 3.84 (a), the survey current beam be-
comes “defocused” and spreads out, with a significant amount of current
flowing into the shoulder beds. This is known as an anti-squeeze configura-
tion. Defocusing reduces a tool’s depth of investigation which makes it more
subject to borehole and invasion effects. The preferential flow of current in
the conductive shoulders causes resistive beds to appear thinner than they
are in actuality.
When the surrounding shoulders are more resistive than the bed of in-
terest as in Figure 3.84 (b), the survey current beam becomes more focused
inside the bed. This is known as a squeeze configuration. In extreme cases,
squeezing of the survey current inside a conductive bed can result in a deeper
measurement than in a homogeneous formation. If electrode tools with dif-
ferent focusing are run together, squeeze effect can create resistivity curve
158 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.84: Computed current patterns and equipotential surfaces for SFL
centered in (a) a thin resistive bed and (b) a thin conductive bed.

separations that imitate an invasion profile; apparent resistivity readings are


shifted because the survey and bucking currents are affected differently de-
pending on the proximity of individual electrodes to the conductive bed (see
also Figure 3.74). The preferential flow of current in conductive beds causes
them to appear thicker than they are in actuality.
Complex formation geometries such as thin invaded beds weaken focusing
to an even greater degree since invaded zones provide an additional pathway
for diverting the survey current away from the deeper noninvaded forma-
tion. This causes log measurements to depart from Rt and to approach Rxo .
Figure 3.85 illustrates the manner in which current and voltage patterns
are affected by the presence of both conductive or resistive shoulders and
conductive or resistive invasion. Four cases are modeled, each with differ-
ent resistivity contrasts between the uninvaded formation (Rt ), invaded zone
(Rxo ) and shoulder beds (Rsh ). Figure 3.85 (a) corresponds to the invaded
3.4. ELECTRODE (LATEROLOG) TOOLS 159

Figure 3.85: Computed current patterns and equipotential surfaces for SFL
centered in thin invaded beds. Four different resistivity contrasts are mod-
eled, with the relative resistivities indicated at the top of each panel.
160 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

bed between 113 and 123 feet in Figure 3.83. Figure 3.85 (b) corresponds to
the invaded bed between 27 and 37 feet. The other two cases are included
for completeness.
In Figure 3.85 (a), it is evident that much of the survey current is diverted
into the conductive shoulder beds through the invaded zone, which is less
resistive that the uninvaded formation. In Figure 3.85 (b), more of the survey
current reaches the uninvaded formation than in (a) because the invaded zone
is less conductive, but the current still spreads into the conductive shoulder
beds to a large extent. In Figure 3.85 (c), the resistive shoulder beds channel
the survey current into the conductive invaded zone and subsequently into
the still more conductive uninvaded formation. In Figure 3.85 (d), some
of the survey current flows back into the borehole through the conductive
invaded zone since it is prevented from circulating far from the tool by the
resistive uninvaded formation and the even more resistive shoulder beds.
Computed current patterns and equipotential surfaces such as these help to
understand how the complexity of the formation influences the dynamics of
tool focusing. They also demonstrate how difficult it is for electrode tools
to read Rt in complex formations and why inversion is often necessary.

3.4.9 High Resolution Laterolog Array (HRLA)

Up to this point, the evolution of electrode devices has consisted of modi-


fying tool design based on modeling, and improving interpretation through
increasingly complex charts and data processing algorithms. The main goals
have been to sharpen vertical resolution and to eliminate parasitic effects.
However, in spite of recent improvements to the modern Dual Laterolog–
Microresistivity tool, it still can give ambiguous values for the true formation
resistivity, Rt , in laminated reservoirs, in thin invaded beds and in deviated
or horizontal wells.
In the 1980’s, it was recognized that inversion techniques which take into
account the true 2D or 3D formation structure [271, 273, 268, 267] were useful
for improving the estimation of formation resistivity from electrode tool mea-
surements. However, inversion still did not eliminate all ambiguities because
of the inadequate information content of the tools of that time. Resistivity
log inversion temporarily fell out of favor and iterative forward modeling
became popular because it allowed high-resolution information from other
measurements (including geological surveys and core analysis) to be incor-
porated in the solution. Even though data integration often had to be done
3.4. ELECTRODE (LATEROLOG) TOOLS 161

mode 1 mode 3 mode 5


mode 0 mode 2 mode 4
0V
0V
0V
0V
0V
0V
24 ft
A0 xV xV xV xV xV xV

0V
0V
0V
0V
0V
0V

position

potential (V)

Figure 3.86: Schematic representation of the HRLA electrode array.

manually, the added confidence in the accuracy of the iterative modeling


solution was deemed to be worth the additional effort.
In the early 1990’s, the concept of multiple-spacing arrays used for in-
duction tools [151] was applied to electrode arrays in order to increase their
intrinsic information content and to design a tool more suited to inversion.
Studies performed by Halliburton [259, 260] and Baker-Atlas [141, 152] in
collaboration with Shell concentrated on using a single current injection
electrode and multiple potential monitoring electrodes which make normal
or lateral measurements. Although these arrays are easy to build and to
model, they suffer from the disadvantages associated with normals and later-
als: strong borehole effect and shoulder effect extending over large distances.
To decrease these effects, measurements can be combined to simulate focused
tool response. However, the combination of measurements made at different
logging depths is prone to errors caused by imperfect control of sonde motion
and sonde position in the borehole.
In an attempt to achieve a more accurate measurement, Schlumberger
162 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Array 1 Array 2 Array 3 Array 4 Array 5

Figure 3.87: Computed current patterns for the five HRLA arrays (RLA1
through RLA5) showing the flow of the survey current (dark) and buck-
ing current (light) for progressively deeper measurements, left to right. A
homogeneous 1 ohm-m formation is modeled.

developed a focused array tool based on the LL3 principle [237, 128]. Since
the tool is robustly focused, borehole and shoulder effects are minimized
at the measurement stage. Signals from all arrays are measured at the
same time and logging position. This ensures that all measurements are
exactly depth-aligned and avoids the generation of artifacts on logs caused
by irregular tool motion.
A schematic representation of this tool, called the High Resolution Lat-
erolog Array (HRLA), is shown in Figure 3.86 [237]. The tool is symmetrical
with six focused measurement modes yielding six different depths of inves-
tigation. A central current electrode (A0 ) emits the survey current. It is
surrounded by six segmented bucking current electrodes on each side, plus a
total of twelve monitoring electrodes. The bucking current focuses the sur-
vey current into the formation at variable depths of investigation as shown in
Figure 3.87 [128]. Progressively deeper measurements are created by main-
taining additional bucking electrodes around A0 at the same potential as
shown in Figure 3.86. The remaining outer electrodes are set at zero po-
3.4. ELECTRODE (LATEROLOG) TOOLS 163

tential and act as current returns. The monitoring electrodes are used to
maintain the accuracy of the equipotential conditions close to the center of
the tool.
In order to achieve a reasonably deep depth of investigation without
making the tool impractically long, the conductive housings of the tools im-
mediately above and below the laterolog device are used as part of the array.
By having all currents return to the tool body rather than to the surface,
voltage reference effects are eliminated. In addition, there is no longer a need
for the cumbersome insulating bridle. Thus Groningen effect and drill-pipe
effects that encourage current flow inside the borehole in horizontal wells are
not a problem.
The HRLA acquires its six measurements simultaneously at frequencies
ranging from 75 to 270 Hertz. The six measurement modes are focused (i.e.,
equipotential conditions are enforced) by a combination of hardware and
software focusing [128]. The hardware injects the currents in a way that is
as close to focused as possible. However, hardware focusing is subject to
physical limitations which can result in slight voltage imbalances in the dy-
namic logging environment. Software focusing by means of the mathematical
superposition of signals is used to ensure that the focusing conditions are
respected by correcting any imperfections.
The result is six focused measurements with varying depths of investiga-
tion that are intrinsically resolution matched. The shallowest mode, RLA0,
is sensitive primarily to the borehole environment and is used to estimate the
mud resistivity. The apparent resistivities RLA1 through RLA5 are sensitive
to the formation at progressively deeper depths of investigation. The appar-
ent resistivity measurements for each of the five HRLA arrays are obtained
by first dividing the potential at A0 with respect to the cable armor by the
magnitude of the survey current, and then multiplying by the appropriate
tool constant, or
VA i
Ra i = KLAi 0 , (3.57)
IA0 i

where i denotes the array sequence (from 1 to 5) and KLA is the tool
constant.
The six depths of investigation offer a better differentiated set of measure-
ments for inversion than the Dual Laterolog. A real-time 1D inversion [237]
is available for providing Rt at the wellsite. The algorithm is a three param-
eter inversion that assumes step profile invasion and uses the five formation
164 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.88: HRLA log.

measurements at each logging station. Prior to inversion, the logs RLA1


through RLA5 are corrected for borehole effect (including eccentricity) us-
ing RLA0 to determine the mud resistivity.
There is also a slower 2D axisymmetric inversion [128] available which
simultaneously accounts for both radial and axial resistivity variations. Bed
boundary locations are first determined by inflection point segmentation.
Initial formation parameters (Rt , Rxo and di ) are derived from the input
resistivity logs. A fast 2D finite element forward model is then used to com-
pute tool response to this initial forward model. The computed response is
compared to the actual logs, and formation parameters are adjusted based
on the computed sensitivities. The process of tool response modeling and
3.4. ELECTRODE (LATEROLOG) TOOLS 165

Figure 3.89: 2D inversion results for the HRLA log shown in Figure 3.88.

formation model refinement is iterated until a good match is obtained be-


tween the computed and actual logs. By inverting for the combined effects
of invasion and shoulder beds in two dimensions, their interdependent effects
are accounted for more exactly than by 1D inversion. A fully 3D inversion
is also under development.
Figure 3.88 shows a HRLA log in the benchmark formation. In the
uninvaded beds, the optimal electrode arrangement and more robust focusing
of the HRLA provide better vertical resolution than LL3 (Figure 3.67). In
the invaded beds, the depth of investigation of the RLA3 curve is closest
to that of LL3, which has bucking electrodes nearly the same length as the
RLA3 mode. There is a uniform spread between the five HRLA curves in the
invaded beds. However, even though the invasion is not extremely deep, all of
166 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

the HRLA curves cluster near Rxo . In the resistive invaded bed between 113
and 123 feet (which represents a typical oil-bearing zone), the low apparent
resistivities could easily lead to a low interpreted value for Rt . This, in turn,
would cause an underestimation of calculated reserves (see Section 5.1), or
perhaps even a missed production opportunity. Because of the channeling of
laterolog currents by invasion, an electrode tool cannot yield a measurement
of Rt in invaded thin beds which is as deep as AIT (Figure 3.32) or even
DIT (Figure 3.26) without processing or inversion.
The five HRLA curves do, however, provide more detailed information
about the invaded zone than the Dual Laterolog (Figure 3.74). Although
the separation between the HRLA curves in the invaded beds in Figure 3.88
may seem small, this information is crucial for quantifying invasion effect
and Rt by means of inversion. Figure 3.89 shows the results obtained from
the 2D inversion of the logs in Figure 3.88. The inversion clearly indicates
invasion in the two beds between 27 and 37 feet, and between 113 and 123
feet. Although Rxo , Rt and Ri are slightly lower than the actual values,
a reservoir evaluation using Rt from the inversion would certainly be more
accurate than an evaluation based on the logs in Figure 3.88 before inversion.
A new Rxo pad tool, the MicroCylindrically Focused Log (MCFL) is
normally run in the field with the HRLA. The MCFL uses both longitudi-
nal and azimuthal focusing [108] to provide a Rxo measurement with high
(1 inch) vertical resolution, which is useful for validating the inversion of
field logs [237].

3.5 Microresistivity tools

Microresistivity tools provide a measurement of the resistivity of the invaded


zone, Rxo , with high vertical resolution. A direct measurement of Rxo is use-
ful for correcting deep resistivity measurements in order to better determine
Rt . The first microresistivity tools were introduced in the late 1940’s. To ac-
curately measure Rxo , a tool must have a very shallow depth of investigation
because the invaded zone may extend only a few inches beyond the borehole
wall. Since shallow mandrel tools with short spacings are subject to severe
borehole effect, a sidewall pad design was adopted for microresistivity tools.
Pad devices consist of a short-spaced tool mounted on a curved surface,
such as the Microlog tool shown in Figure 3.90. The pad is mechanically
pressed against the formation to reduce the short-circuiting effect of the
3.5. MICRORESISTIVITY TOOLS 167

Figure 3.90: The Microlog pad, front (left) and side (right) views, showing
the electrode configuration.

mud. Microresistivity tool response is, however, affected by the presence of


the mudcake (see Section 2.1.1 and Figure 2.1), which can build up against
the borehole wall in permeable formations. The amount of mudcake effect
depends on the mudcake thickness, hmc , and the mudcake resistivity, Rmc .
Moreover, mudcakes can be anisotropic, with the mudcake resistivity parallel
to the borehole wall less than that across the mudcake. Mudcake anisotropy
increases the mudcake effect so that the effective (or electrical) mudcake
thickness is greater than the thickness indicated by the caliper which samples
the borehole size.
The evolution of microresistivity tools parallels that of laterolog devices.
During the 1950’s through the 1970’s, pad tool equivalents of the normal,
lateral, LL3, LL7 and SFL were developed for measuring Rxo . The vertical
resolution of these tools ranges from two to six inches. During the 1950’s
and 1960’s, microresistivity tool response was modeled experimentally using
a test tank. In the 1970’s, 2D tool response was computed by assuming
the pad was sufficiently small to neglect borehole curvature and treating
the mudcake as a thin planar layer [119]. Starting in the mid-1980’s, tool
168 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

response was simulated using full 3D modeling codes [127]. Some of the older
microresistivity tools described in this section (i.e., the MicroLaterolog and
the Proximity log) are no longer in use.

3.5.1 The Microlog

The first microresistivity device introduced was the Microlog [90]. The tool
configuration is shown in Figure 3.90 [90]. A button electrode, A, mounted
on an insulated pad, emits a current into the formation which returns at a
distance sufficiently large to be considered at infinity. The potential is mea-
sured at two monitoring electrodes, M1 and M2 , located in vertical alignment
with the current electrode at distances of 1 inch and 2 inches, respectively.
The tool provides two independent measurements with different depths of
investigation. There is a 2 inch micronormal resistivity, R2 , with

VM 2
R2 = K2 , (3.58)
IA0

and a 1 inch microinverse resistivity, R1 ×1 , with

VM 1 − V M 2
R1 ×1 = K1 ×1 . (3.59)
IA0

The values of K that are used vary with borehole size. For an 8-inch bore-
hole, K1 ×1 is 0.32 and the ratio K2 /K1 ×1 is 1.3. Given the spacings,
the 2 inch micronormal has a greater depth of investigation than the mi-
croinverse, and both measurements can resolve beds which are a few inches
thick.
The two measurements can also be used to delineate permeable beds by
identifying the presence of a mudcake. When invasion occurs, a mudcake
builds up against the borehole wall, causing the micronormal and microin-
verse curves separate. Usually the resistivity of the mudcake is considerably
lower than that of the invaded zone, which causes the shallower microinverse
to read a lower resistivity value than the micronormal. Correction charts,
such as the one shown in Figure 3.91 [230], can be used to derive an estimate
of Rxo /Rmc and hmc by entering the tool readings on the left and bottom
axes (assuming that Rmc is known from direct measurements). Although the
separation of the micronormal and microinverse curves can be used to flag
permeable zones, quantitative inferences of permeability are not possible.
3.5. MICRORESISTIVITY TOOLS 169

20

15

10
9
8

1.5

1
1 1.5 2 3 4 5 6 7 8 9 10 15 20

Figure 3.91: The Microlog interpretation chart.

One of the most successful uses of the Microlog has been in evaluat-
ing thinly bedded shaly sand reservoirs. No curve separation is seen in
front of impermeable shale beds (as long as good pad contact is achieved).
Estimating the sand reservoir footage by “counting the sands” can easily
be done by cumulating the zones on the log where the micronormal and
microinverse curves separate.
In tight formations where invasion is usually negligible, both curves read
similar values. The main disadvantage of the Microlog is that it lacks res-
olution for large values of Rxo /Rmc , as shown by the closeness of the high
contrast curves in Figure 3.91. In cases where the invasion depth is less
170 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.92: The MicroLaterolog pad showing a front view of the electrode
configuration (left) and a side view of current patterns (right).

than four inches, Microlog readings are affected by Rt . Being unfocused, the
micronormal and microinverse suffer from the same limitations as unfocused
mandrel electrode tools in all high contrast situations.

3.5.2 The MicroLaterolog

The MicroLaterolog [89] was introduced as a replacement for the Microlog


in high contrast environments. It is a focused pad tool which operates on
the same principle as the mandrel LL7. Circular electrodes are used in order
to focus the survey current in both the vertical and azimuthal directions,
resulting in a cone-shaped narrow beam. The MicroLaterolog electrode con-
figuration and current patterns are shown in Figure 3.92 [89]. There is a
central A0 current electrode surrounded by concentric M1 , M2 and A1 cir-
cular ring electrodes. The distance between successive rings ranges from one
half to one inch. Each of the ring electrodes is composed of a series of small
circular buttons which are short-circuited together.
3.5. MICRORESISTIVITY TOOLS 171

Figure 3.93: Pseudo-geometrical factors comparing the depths of investi-


gation of the MicroLaterolog (MLL), Proximity log (PL) and MicroSFL
(MSFL) for Rxo < Rt (left) and Rxo > Rt (right).

A constant current of known intensity is emitted from A0 . A bucking


current is emitted from the outer A1 electrode. The magnitude of the bucking
current is controlled by a feedback loop which maintains the M1 and M2
monitoring electrode rings at the same potential. Currents return to an
electrode located on the bridle. The MicroLaterolog apparent resistivity is
VM 1
Ra = KM LL . (3.60)
IA0

The value of KM LL that is used varies depending on borehole conditions.


The MicroLaterolog has a shallow depth of investigation of only a few
inches as illustrated by the pseudo-geometrical factor curves in Figure 3.93.
When the mudcake thickness is greater than 3/8 inch, MicroLaterolog re-
sponse should be corrected for mudcake effect using the chart shown in
Figure 3.94 [230]. Similar charts are available for correcting the response of
other microresistivity tools for mudcake effect.

3.5.3 The Proximity log

The Proximity log was introduced in 1960. It was designed to be less sen-
sitive to thick mudcakes than the MicroLaterolog. It uses the same concept
172 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.94: The MicroLaterolog mudcake correction chart.

of passive guard focusing as the LL3 mandrel laterolog tool. The Proximity
log electrode configuration is shown in Figure 3.95. There is a central rect-
angular A0 survey current electrode and a large A1 guard electrode located
on the outer edge of the pad. Both current electrodes are short-circuited
together and maintained at a constant potential. The thin M0 electrode lo-
cated between A0 and A1 measures the potential with respect to a reference
electrode located on the bridle. This location of M0 makes the potential
measurement insensitive to electrode impedance variations and to the effect
of current flow. The Proximity log apparent resistivity is

VM 0
Ra = KP L . (3.61)
IA0

The value of KP L that is used varies depending on borehole conditions.


Since the Proximity log has a significantly deeper depth of investiga-
tion than the Microlog and the MicroLaterolog, mudcake correction is only
necessary when the mudcake thickness is greater than 3/4 inch. However,
the depth of investigation of the Proximity log was found to be too deep in
practice; if invasion is shallow or moderate, the effect of Rt is noticeable,
especially when Rxo is larger than Rt (see Figure 3.93). In this case, it be-
comes necessary to use other measurements along with the Proximity log
in order to solve simultaneously for Rxo , Rt and invasion diameter. The
“Grand Slam” interpretation method [99] was introduced for this purpose.
This method uses Dual Induction measurements in addition to the Proximity
log. The vertical resolution of the Proximity log is approximately 6 inches.
3.5. MICRORESISTIVITY TOOLS 173

A1 guard

M0 measure

A0 survey

Figure 3.95: The Proximity log electrode configuration.

3.5.4 The MicroSpherically Focused Log (MSFL)

The MicroSFL (MSFL) was introduced to provide a Rxo measurement for


the Dual Laterolog tool [245]. It is mounted on the lower A2 guard elec-
trode of the tool as shown in Figure 3.72. The MSFL was designed to have
two distinct advantages over earlier microresistivity devices: (1) a much
shallower depth of investigation than the Proximity log which makes it less
subject to the effect of Rt , and (2) less sensitivity to mudcake effect than
the MicroLaterolog.
The MSFL concept was adapted from the mandrel SFL tool. The elec-
trode configuration is shown in Figure 3.96 [245]. The tool design was op-
timized in a test tank during the early 1970’s and required several years of
laboratory experimentation (no 3D codes were available at that time). There
are two current electrodes, A0 and A1 , a measure potential electrode, M0 ,
and two monitoring potential electrodes, M1 and M2 . A survey current is
emitted from A0 and is returned to the back of the pad, which for practical
purposes can be assumed to be at infinity. Because the current return is on
the pad itself, the MSFL can be easily combined with other tools, and it
is sometimes run with the Dual Induction tool. A bucking current is also
emitted from A0 which returns to A1 . The magnitude of the bucking cur-
rent is controlled to ensure that a null is maintained between the M1 and
M2 monitoring electrodes. The measure voltage is taken between the M0
174 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Figure 3.96: The MicroSFL current distribution (left) and electrode configu-
ration (right). The i0 curve demotes the survey current and ia is the bucking
current. The shaded area denotes the volume of investigation.

and M1 potential electrodes.


Two external focusing conditions are specified: the potential difference
between the monitor electrodes is null (VM2 − VM1 = 0), and the measure
potential is maintained at a preset value Vref , with Vref = VM0 − 12 (VM1 +
VM2 ). The apparent resistivity is
Vref
Ra = KM SF L . (3.62)
IA0

As illustrated in Figure 3.93 the depth of investigation of the MSFL is


only a few inches, which is significantly less than that of the Proximity log.
Mudcake correction is only required when the mudcake thickness is greater
than 5/8 inch. Mudcake effect is small because the SFL type focusing forces
the survey current to flow directly into the formation. The rectangular
shape of the electrodes also contributes to reducing mudcake effect while
preserving the shallow depth of investigation. 3D computer modeling of
MSFL response in the 1980’s confirmed that the rectangular electrodes were
able to accurately maintain the focusing conditions although they are not
natural equipotential shapes.
3.6. IMAGING TOOLS 175

Synthetic Microlog curves can be computed from the partial MSFL re-
sponse. Since the survey current sees primarily the invaded zone and the
bucking current sees primarily the mudcake, it is possible to mathematically
construct micronormal and microinverse curves.

3.6 Imaging tools

Imaging tools evolved jointly from microresistivity tools and from the dip-
meter tool. Dipmeter measurements have been used since the 1950’s [83, 60]
to determine formation structure, rather than resistivity. A dipmeter tool
consists of four conducting pads positioned at 90◦ to one another [6]. Each
pad contains a small button electrode which emits a low-frequency current.
All pads and buttons are held at a constant potential relative to a return
electrode located in an upper section of the tool string. Thus the dipme-
ter is a passively focused tool, like the Proximity log and LL3. The button
current measured on each pad is proportional to local resistivity variations
occurring directly in front of it. Formation dip is calculated by correlat-
ing small-scale bedding features which cross the pads. To ensure that dips
are oriented properly with respect to one another over the entire length of
the borehole, the azimuthal position of the tool is recorded with a mag-
netometer. Modern dipmeter tools include a second button on each pad
for cross-correlation. The single button pad is called the High-resolution
Dipmeter Tool (HDT), and the two button pad is called the Stratigraphic
High-resolution Dipmeter Tool (SHDT). (Formation MicroScanner imaging
tools also make dipmeter measurements, and SHDT buttons are shown in
Figure 3.97.) Like other electrode tools, conventional dipmeters can only be
run in water-based muds. An induction dipmeter, call the Oil-Based mud
Dipmeter Tool (OBDT) [104, 160] uses the voltage difference generated by
a transmitter at two receivers to measure dip in oil-based muds.

3.6.1 Formation MicroScanner (FMS)

Improvements in data acquisition and transmission that occurred during the


1980’s led to the development of pad tools containing 2D arrays of micro-
electrodes which record both vertical and azimuthal variations in resistivity.
Currents measured by these electrode arrays are scaled and signal processed
to generate electrical images of the borehole wall which resemble core pho-
176 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Inclinometer

2.8 in.
Preamplification
Cartridge

27 buttons
Hydraulics 0.2 in. diameter
0.4 in. 50% overlap
0.1 in.
Side-by-side
SHDT buttons
4 arm sonde

Figure 3.97: The four-arm FMS tool (left) and a close-up view of a single
pad (right).

tographs. The first electrical imaging tool was Schlumberger’s Formation


MicroScanner (FMS) [109], which was introduced in 1985. The FMS tool is
shown in Figure 3.97 [109].
3D finite element modeling [127] was used to optimize the size and overlap
of the FMS button electrodes, and to determine the optimum placement
of the electrode array in order to avoid the defocusing of button currents
that occurs near pad edges. The spatial resolution of the FMS is 0.2 inch
in depth and in azimuth. Depth of investigation is one to three inches.
Figure 3.98 [234] compares a core photo to images from adjacent FMS pads
in the same formation. Small variations in current intensity of less than 1 cm
in size can be seen on the FMS images in the darker clay matrix.
Because there are large gaps between the four pads, FMS images covers
only about 20% of the surface of an 8 inch diameter borehole. To address
this problem, an improved version of the tool, called the Fullbore Formation
MicroImager (FMI) [217], was introduced in 1991. The FMI has an auxiliary
pad (called a “flap”) attached to each of the four main pads. The new tool
3.6. IMAGING TOOLS 177

FMS FMS
image core image
pad 3 photo pad 4

5113.2 m
The lighter zones on
the core (white on the
images) are very fine
grain limestone

The darker, more


broken looking zones
are limestone clasts
in a clay matrix

5113.8 m

7 cm

Figure 3.98: Comparison of a core photo (center) and images from two FMS
pads (left and right) in a limestone formation. Dark areas are conductive
media and light areas are resistive media.

has a total of 192 electrodes on eight pads, giving 80% coverage in an 8 inch
borehole.
The FMI was so successful that electrical imaging soon rivaled coring as
one of the primary methods used for structural analysis in complex reser-
voirs. Imaging logs were widely used not only by petrophysicists, by geolo-
gists, geophysicists and petroleum engineers as well. The success of the FMI
created a demand for imaging tools tailored to specific applications.
178 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Vm

M3 M3

Ii ∆V=0 ∆Vi

Ic
M4 M4

Figure 3.99: ARI electrode configuration showing the main imaging mea-
surement (left) and a shallow auxiliary measurement (right) which is used
to correct the main measurement for borehole and eccentricity effects.

3.6.2 Azimuthal Resistivity Imager (ARI)

Because FMI images are generated from signal processed measurements of


relative current intensity, they are difficult to calibrate to formation resis-
tivity. The Azimuthal Resistivity Imager (ARI) [79, 236] was introduced to
provide a quantitative image of the formation resistivity beyond the invaded
zone. The ARI is a mandrel tool that is mounted on the A2 electrode of the
Dual Laterolog (see Figure 3.72). The ARI electrode configuration is shown
in Figure 3.99 [79].
Low-frequency currents are emitted from twelve azimuthal electrodes
which encircle the mandrel. The electrode voltages are held at the same
potential as two surrounding monitoring ring electrodes (M3 and M4 in Fig-
ure 3.99), providing active lateral focusing. Azimuthal focusing is passive
by means of adjacent currents. Twelve azimuthal resistivity values, Ri , are
computed from the Ii electrode currents, with
Vm
Ri = k , (3.63)
Ii
where Vm is the mean potential of the M3 and M4 electrodes relative to
the cable armor, and k is a tool constant. A shallow auxiliary measurement
is also performed for correcting the azimuthal resistivities for borehole and
3.6. IMAGING TOOLS 179

eccentricity effects. The auxiliary measurement currents are returned to the


A2 electrode as shown in Figure 3.99, which forces them to remain primarily
within the borehole.
The vertical resolution of the ARI is approximately 8 inches, and az-
imuthal resolution is 60◦ . ARI resolution is poorer than that of FMI be-
cause ARI electrodes are much larger, and because the ARI is not in direct
contact with the borehole wall, which allows electrode currents to fan out in
the mud. However, the ARI resistivity measurements have higher vertical
resolution than the 2 to 3 foot resolution of the Dual Laterolog. And unlike
the FMI, ARI measurements can be used as apparent resistivities in satu-
ration calculations. The depth of investigation of the ARI is between that
of LLS and LLD. ARI images are useful for obtaining a general overview of
bedding structure, for locating fractures, and for correcting dip effect on the
Dual Laterolog logs.

3.6.3 Resistivity-At-the-Bit tool (RAB)

The Resistivity-At-the-Bit tool (RAB) [57, 212, 179] was introduced to pro-
vide resistivity images in the logging while drilling environment. The tool is
designed to operate under laterolog conditions (salty muds, high formation
resistivity and conductive invasion). It is either attached to the drill bit or
located higher on the drill string near the mud motor. Figure 3.100 [57]
shows the tool configuration.
A toroidal transmitter drives a low-frequency axial current along the
drill pipe. This induces a voltage on the drill collar which causes currents to
flow down the collar out into the formation, and then return to the collar.
Toroidal receivers measure the axial currents traveling along the collar, while
ring and button electrodes measure the currents leaving the tool. The mag-
nitudes of all the currents are determined by the resistivity of the formation.
The tool makes five independent resistivity measurements. Two mea-
surements, one at the bit and one at the ring (see Figure 3.100), are non-
azimuthal. Quantitative azimuthal image measurements with three different
depths of investigation are made by three button electrodes mounted on
the side of the drill collar. As the drill string rotates, the buttons scan the
borehole wall, producing 360◦ images. An azimuthal scan typically consists
of 56 azimuthal samples. Vertical resolution is approximately 2 inches and
azimuthal resolution is approximately 15◦ .
180 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Batteries

Upper transmitter and


wireless telemetry

Azimuthal
electrodes

Ring electrode and


current monitor

Azimuthal Gamma Ray

Stabilizer

Lower transmitter and


current monitor

Float valve bore

Figure 3.100: RAB electrode configuration.

The average depths of investigation of the three button images are 5, 8


and 10 inches. The three measurements are used to evaluate shallow early-
time invasion. By combining the button and ring measurements, Rt can be
determined in the presence of invasion.
Since RAB measurements are taken close to the bit, they can be used for
geosteering and for selecting casing and coring points. There is also a gamma
ray sensor located between the RAB electrodes. The sensor is eccentered,
and its count rates are binned in quadrants. This data is used to produce
low-resolution images of formation density.

3.6.4 Oil-Base MicroImager tool (OBMI)

The Oil-Base MicroImager tool (OBMI) [66] was introduced in 2000 to pro-
vide imaging capability in synthetic and oil-based muds. These muds have
3.6. IMAGING TOOLS 181

Figure 3.101: Side view (left) of the OBMI tool showing the measurement
principle, and front view (right) showing the electrode arrangement.

recently grown in popularity because they allow improved borehole stability


and increased drilling efficiency, especially in deep-water environments.
The OBMI uses the four-terminal method of measuring resistivity. Fig-
ure 3.101 [66] shows the tool configuration. An alternating current, I, is
injected into the formation between electrodes A and B. The potential dif-
ference δV is measured between a pair of small voltage sensors at C and
D. The apparent resistivity in the interval of the formation opposite these
sensors is given by
δV
Ra = k , (3.64)
I
where k is a tool constant. The depth of investigation of the OBMI is three
to four inches, and the tool is considered to be a Rxo measurement.
There are five pairs of potential measuring electrodes located at the mid-
dle of the pad as shown in Figure 3.101. The OBMI has four pads mounted
on arms positioned at 90◦ around the borehole. The five measurements from
the four pads are displayed as an image of the formation resistivity around
the borehole wall that is 20 pixels wide. The voltage sensors are spaced
0.4 inch apart both vertically and horizontally, giving the pixels a nominal
size of 0.4 × 0.4 inch. The vertical resolution of the tool is 1.2 inches. Pad
182 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

coverage is 32% in an 8 inch diameter borehole. The maximum standoff that


the tool can tolerate increases with the formation resistivity and decreases
with mud resistivity, ranging from 0.2 to 0.5 inch. At this time, the OBMI
is the only imaging tool that can make measurements in oil-based muds.

3.7 Resistivity through casing

The first patents for resistivity measurements through casing were filed in
the 1930’s [13]. Although the basic measurement principle (described below)
is relatively simple, the orders of magnitude of precision needed to make an
accurate measurement were unachievable at that time. The subject was re-
visited by Kaufman [155, 156] and Vail [258] in the late 1980’s. The first
commercial logging tools were introduced in the late 1990’s by Baker-Atlas
and Schlumberger. The Baker-Atlas tool is called the Through Casing Re-
sistivity tool (TCR) [184]. The Schlumberger tool is call the Cased Hole
Formation Resistivity (CHFR) tool [55, 134, 52].
The CHFR tool configuration is shown in Figure 3.102 [55]. Current is
sent via the wireline cable to the tool in the same manner as it is for open
hole laterolog tools. The measure current path is indicated by the dark solid
lines in Figure 3.102. Injection electrodes send current into the casing, where
it flows in both directions to return to the surface. Most of the current flows
inside the casing, but some leakage into the formation occurs.
Three voltage electrodes (A, B and C in Figure 3.102) are applied to
the casing to measure the formation current that leaks out of it. The three
voltage measurements give an estimate of the first derivative of the axial
current, which is proportional to the formation conductivity.
Since the tool measures the casing current by means of its voltage drop
in the casing segment resistance, the measurement must be calibrated to
account for any difference in the two sections. The output is proportional to
both the formation current ∆I and also to the casing resistance difference
∆Rc . When the tool is switched to “calibrate” mode (dotted path in Fig-
ure 3.102), current is injected using a downhole current source, with a small
distance between injection and return. In this case, the formation current
∆I is null and ∆Rc is measured directly.
The formation current is derived by combining the results of the “mea-
sure” and “calibrate” acquisition steps (the tool cannot move between the
3.7. RESISTIVITY THROUGH CASING 183

(solid)

(dotted)

Rt = K . Vo / ∆I, where ∆I = (V1 - V2) / Rc

Figure 3.102: CHFR tool configuration and measurement principle. Dark


solid lines indicate the “measure” mode, and dotted lines indicate the “cal-
ibrate” mode.

two steps). Formation resistivity is calculated using the tool voltage with
respect to the surface and a tool constant (K), in the same manner as other
electrical tools.
The tool uses a low-frequency alternating current because direct current
polarizes and drifts under these circumstances. Skin effect in the casing
(δ = 5 mm at 5 hertz) limits the tool frequency to a few hertz.
Typical formations have resistivities about a billion times that of a steel
casing. Because currents are sensitive to the geometry of the materials they
travel through, the large volume of the formation allows the ratio between the
formation current and the injected current to (fortunately) be in the range
of 10−4 instead of 10−9 . Since the formation current is measured through a
drop in casing resistance (around a few tenths of a micro-ohm), the actual
tool measurements are in the nanovolt range. The ability to handle these
small measurements under logging conditions only became possible in the
late 1990’s.
184 CHAPTER 3. ELECTRICAL WELL-LOGGING MEASUREMENTS

Because CHFR tool physics is much different than that of open-hole


tools, analytical and numerical modeling was used to simulate tool response.
This modeling showed that environmental effects, such as the effect of the
cement behind casing, are very limited. Field tests have demonstrated good
agreement between CHFR and open-hole induction and laterolog measure-
ments in the same well. The depth of investigation of the CHFR ranges from
6 to 30 feet, depending on the level of the formation conductivity.
Prior to the introduction of cased-hole resistivity tools, formation evalu-
ation through casing was performed primarily using nuclear measurements.
Now full saturation calculations behind casing are possible using both nu-
clear and resistivity logs. Cased-hole resistivity measurements are particu-
larly useful for monitoring hydrocarbon movement and water floods in ma-
ture reservoirs where all wells are cased. They also are useful when unstable
well conditions have prevented the acquisition of open-hole logs, and for
identifying bypassed hydrocarbons.
Chapter 4

Modeling of tool response

Summary: This chapter describes the most common analytical and numerical
methods used to construct computer codes for modeling resistivity logging tool re-
sponse. The purpose of the chapter is to provide a modeling overview with sufficient
background so that students or interested readers can select the appropriate meth-
ods for writing their own forward modeling codes. The analysis necessary to go
from Maxwell’s equations to synthetic tool response is outlined, with emphasis on
techniques that have proved to be the most computationally efficient for resistivity
tools. To avoid becoming a cookbook, many details have been omitted. Additional
information can be found in the references listed at the end of this chapter. These
references are admittedly biased toward Schlumberger authors for two reasons: the
author of this thesis has worked with these people over the years to develop the
modeling software described in this chapter, and in many cases these were the first
papers published on the subject.

4.1 Analytical methods

Analytical methods are defined as methods which use exact mathematical


solutions to simulate tool response [31]. Although analytical methods may
involve some numerical analysis, such as series representations of Bessel func-
tions or the numerical evaluation of integrals, solutions are still cast in terms
of the partial differential equation being solved. When analytical methods
are applied to 2D geometries, equations can become extremely complex and
186 CHAPTER 4. MODELING OF TOOL RESPONSE

tedious to translate into computer code. For this reason, the use of ana-
lytical methods is usually limited to modeling simple 1D geometries such
as invasion with no layering or layering with no invasion. Since analytical
codes run rapidly and do not require large computer memories, they were
used exclusively for modeling the response of early resistivity tools.
When more powerful computers became available in the late 1970’s, nu-
merical methods began to replace analytical methods. Numerical methods,
such as finite element and finite difference techniques, cast solutions to differ-
ential equations in terms of a large number of simultaneous linear equations,
which are solved using matrix methods. Numerical methods are more eas-
ily adapted to complex geometries than analytical methods, which makes
them well-suited for modeling 2D and 3D logging environments. Although
numerical methods have become very efficient, the simplicity of analytical
methods still makes them useful for understanding the basic physical prin-
ciples governing tool response.
The most commonly used analytical solutions for modeling induction and
laterolog tool response are described in the following section. The emphasis is
on analytical methods for induction tools; laterolog tools ordinarily require
the use of numerical methods to accurately model electrode and borehole
dimensions, which must be represented exactly. Solutions for homogeneous
media are described in detail to demonstrate how tool response is derived
from Maxwell’s equations, and to allow interested readers to construct their
own codes. Solutions for heterogeneous media are summarized briefly, with
additional information provided in the references indicated by the Glossary
of computer codes in Section 4.4. Because this chapter is a review of general
methods, solutions are for isotropic media unless otherwise noted.

4.1.1 Doll’s induction geometrical factor theory

Section 3.2.1 showed how Doll’s geometrical factor theory was used as a
computational aid for optimizing induction coil configurations. The devel-
opment of geometrical factor theory is examined here in greater detail. As a
first order approximation, Doll assumed that the fields generated by a point
dipole transmitter in a wellbore were essentially the same as the fields in a
vacuum. Thus the voltage at a receiver would be the sum of contributions
from an infinite number of eddy current loops. Using the Biot-Savart law,
Doll showed [88] that the contribution of a single loop having a unit cross
sectional area, with a radius ρ and at a distance z from the midpoint of the
4.1. ANALYTICAL METHODS 187

two coils, is
∆ VR = K g(ρ, z) σ(ρ, z) dρ dz . (4.1)
The coil configuration is the same as that of Figure 3.5. σ is the conductiv-
ity of the formation within each loop and K is a tool constant containing
information about dimensions, given by
µ0 2 ω 2 NT AT NR AR IT
K= . (4.2)
4πL
ω is the angular frequency (2π·frequency), µ0 is the magnetic permeability of
free space, AT and AR are the transmitter and receiver cross sectional areas,
NT and NR are the respective coil turns, IT is the transmitter current and
L is the coil spacing. g(ρ, z) is the geometrical factor, or weighting function
associated with each loop, defined as
L ρ3
g(ρ, z) = 3 , (4.3)
2 rT3 rR
wherer is the distance between the loop and the transmitter or receiver coil
(r = ρ2 + (z + L/2)2 ). The total receiver voltage is the integration over
ρ and z of an infinite number of such loops,
 ∞  ∞  ∞  ∞
L ρ3
VR = K 3 σ(ρ, z) dρ dz = K g(ρ, z) σ(ρ, z) dρ dz .
−∞ 0 2 rT3 rR −∞ 0
(4.4)
The 2D geometrical factor g(ρ, z) is shown in Figure 3.6 for a two-coil sonde.
The real apparent conductivity signal, σR , can be expressed in terms of
the receiver voltage by dividing the voltage by the tool constant K. The
apparent conductivity signal is thus the sum of the formation conductivity
elements weighted by g(ρ, z), i.e.,
 ∞  ∞
σR = g(ρ, z) σ(ρ, z) dρ dz . (4.5)
−∞ 0

As a historical note, until the late 1960’s induction response in invaded


thin beds was approximated using tables of precomputed g(ρ, z) values. Fig-
ure 4.1 shows such a table for 6FF40. To compute tool response, g(ρ, z) in
each box was multiplied by the appropriate σ(ρ, z) at that location. The
“integration” was performed by adding the partial responses on a desk cal-
culator. One of this author’s first assignments at Schlumberger was to write
a computer program to automate this task.
188 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.1: Table giving values of g(ρ, z) for 6FF40 at specified depths and
radii.
4.1. ANALYTICAL METHODS 189

Figure 4.2: Vertical geometrical factors for a two-coil sonde and 6FF40.

It is often informative to isolate the contributions to the total signal


from either a thin horizontal slice of the formation at depth z (the vertical
geometrical factor), or a thin cylindrical shell at arbitrary radius ρ (the radial
geometrical factor). Integrating Equation (4.4) over ρ yields the vertical
geometrical factor gz (z), which is defined [88] as
1 L
gz (z) = , for |z| ≤
2L 2
L L
gz (z) = 2 , for |z| > . (4.6)
8z 2
Figure 4.2 shows a comparison of gz (z) for a two-coil sonde and 6FF40.
Integrating Equation (4.4) over z yields the radial geometrical factor
gρ (ρ) [88], which is expressed in the most computationally efficient form [121]
in terms of elliptic integrals as
   
κ 2
κ E(κ) + κ K(κ) − E(κ) ,
2
gρ (ρ) = (4.7)
L
where K and E are complete elliptic integrals of the first and second kind,
and the modulus, κ, and the complementary modulus, κ , are defined as,
L 

κ=  and κ = 1 − κ2 . (4.8)
L2 + 4ρ2
Figure 4.3 shows a comparison of gρ (ρ) for a two-coil sonde and 6FF40.
190 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.3: Radial geometrical factors for a two-coil sonde and 6FF40.

The integrated radial geometrical factor, Gρ (ρ), gives the contribution to


the total signal from a cylinder of radius ρ. The expression for the integrated
radial geometrical factor in terms of elliptic integrals is

ρκ  
Gρ (ρ) = 1 − κE(κ) + K(κ) − E(κ) . (4.9)
L

Figure 4.4 shows a comparison of Gρ (ρ) for a two-coil sonde and 6FF40.
Doll’s geometrical factor theory is still occasionally used as a rapid means
of comparing the response characteristics of induction tools. The geometrical
factor approximation becomes more accurate as a tool’s operating frequency
approaches zero. However, if the frequency of commercial tools is lowered
to a point where geometrical factor theory is accurate enough to provide
a linear solution to the inverse problem, the tool signal becomes too small
to be measured. Applications of the Born approximation to the inverse
problem for induction [192] have shown that Doll’s geometrical factor theory
is equivalent to a first order Born approximation.

4.1.2 Induction skin effect in homogeneous media

Geometrical factor theory was used to successfully approximate the response


of early induction tools such as 5FF27 and 5FF40 because their shallow mea-
surements were influenced very little by skin effect. However, geometrical
4.1. ANALYTICAL METHODS 191

Figure 4.4: Integrated radial geometrical factors for a two-coil sonde and
6FF40.

factor theory proved to be less than adequate for modeling the Dual In-
duction tool, which was designed to read deeper into the formation and
was therefore more susceptible to reduction in signal level in conductive
formations due to skin effect (see Section 3.2.1). Moran and Kunz [194]
derived the exact formulation for modeling the response of induction tools
from Maxwell’s equations in the late 1950’s. In addition to studying tool
sensitivities to skin effect in homogeneous media, they also solved the 1D
planar layered and cylindrically layered problems. Similar derivations were
also published by Duesterhoeft [102, 103], and by deWitte and Lowitz [87],
and in the former Soviet Union by Kaufman [157] and Nikitina [198].
From Equations (2.17)–(2.18) and Equations (2.24)–(2.26), Maxwell’s
equations for induction tools in isotropic media are

∇ × H + iω∗E = J e , (4.10)
∇ × E − iωµH
H = −KK e, (4.11)

where ∗ is the complex permittivity equal to  + iσ/ω.


For most analytical modeling problems, induction transmitter and re-
ceiver coils are assumed to be small dipole current loops around the z-axis
in a rotationally symmetric configuration, as shown in Figure 4.5. The coils
have an infinitesimally small radius, a. The solution for the electromagnetic
fields is found for the given source distribution J (rr). In the coordinate sys-
192 CHAPTER 4. MODELING OF TOOL RESPONSE

r
Receiver

L r
θ
y
ρ
φ
a x
0
Transmitter
IT

Figure 4.5: Small loop antennas in an unbounded medium, showing the


coordinate system.

tem shown in Figure 4.5, r refers to an observation point, while r  and other
primed coordinates refer to a source point.
If ∇ × H is evaluated by means of Equation (4.11) and eliminated from
Equation (4.10), we obtain
∇ × ∇ × E (rr) − k 2E (rr) = iωµJ
J (rr), (4.12)
Where ω 2 µ∗ is replaced by k 2 = iωµσ + ω 2 µ. (Applying the appropriate
vector identities to Equation (4.12) gives the wave equation in a source-free
region
∇2E + k 2E = 0, (4.13)
which is also known as the Helmholtz equation, and is often referred to in
categorizing the induction problem.)
To determine E in terms of a given source J , the use of dyadic Green
functions is introduced. The method of Kong [161] is summarized here. A
more traditional approach uses a vector potential, and is described in Moran
and Kunz [194]. A Green function is the response due to a point source and
is useful for expressing a field in terms of its source. Since both E (rr) and
J (rr) are vectors, we can write the volume integral
  
E (rr) = iωµ dV  G(rr, r  ) · J (rr ), (4.14)
4.1. ANALYTICAL METHODS 193

where G(rr, r ) is the dyadic Green function that enables one to determine
the electric field E from a given source distribution J . (A dyad can be
defined in terms of two vectors; in index notation the ijth component of a
dyad D is Dij = Ai Bj .)
The right hand side of Equation (4.12) can be cast in a form similar to
Equation (4.14) by using the three-dimensional delta function δ(rr −rr ), such
that   
J (rr) = dV  δ(rr − r ) I · J (rr ), (4.15)

where I is a unit dyad represented by a unit diagonal matrix. Substituting


Equation (4.14) and Equation (4.15) into Equation (4.12) gives a differential
equation for the dyadic Green function G(rr, r  )
∇ × ∇ × G(rr, r  ) − k 2G(rr, r  ) = I δ(rr − r  ). (4.16)
Because the observation point r is always assumed to be outside the source
distribution, the differential operator and the volume integal can be inter-
changed. The dyadic Green function can in turn be expressed in terms of a
scalar Green function g(rr, r  )
 
1
G(rr, r  ) = I + ∇∇ g(rr, r  ), (4.17)
k2
making use of the dyadic operator ∇∇.
General solutions to Equation (4.13) can be written in terms of eikr , and
the scalar Green function is determined to be

eik|r −r |
g (rr, r  ) = , (4.18)
4π|rr − r  |
where |rr −rr | is the distance from the observation point r to the source point
r .
For a small current loop with an infinitesimally small radius a as shown
in Figure 4.5, the current density takes the form
J (rr ) = iφ IT δ(ρ − a) δ(z  ). (4.19)
In cylindrical coordinates the electric field vector due to the current loop is
calculated from
  2π a ∞ 
1   J (rr )eik|r −r |
E (rr) = iωµ I + 2 ∇∇ · dφ dρ ρ dz  (4.20)
k 4π|rr − r  |
0 0 −∞
194 CHAPTER 4. MODELING OF TOOL RESPONSE

To evaluate the integral, the radial vectors r and r are expressed in terms
of their Cartesian components

r = ix r sin θ cos φ + iy r sin θ sin φ + iz r cos θ


r  = ix a cos φ + iy a sin φ . (4.21)

The current loop is in the x-y plane and φ = π/2 for the radial vector r  .
The distance |rr − r  | is

|rr − r  | = |iix (r sin θ cos φ − a cos φ ) + iy (r sin θ sin φ − a sin φ ) + iz r cos θ|



a2 2a
= r 1+ − sin θ cos(φ − φ ). (4.22)
r2 r

The scalar Green function is then expanded in the form of a MacLauren


series for a/r → 0. Taking the first two terms gives

 

eik|r−r | eikr a d eik|r−r |
≈ +
4π|rr − r  | 4πr r d(a/r) 4π|rr − r  | a/r→0
eikr a eikr
= + (−ikr + 1) sin θ cos(φ − φ ) . (4.23)
4πr r 4πr

Noting that iφ = −iix sin φ + iy cos φ , the first integral in Equation (4.20) is
evaluated by substituting Equation (4.23) to give

2π 
IT eik|r −r |
a dφ (−iix sin φ + iy cos φ )
4π|rr − r  |
0
πa2 IT eikr
= (−iix sin φ + iy cos φ) (1 − ikr) sin θ
4πr2
πa2 IT eikr
= iφ (1 − ikr) sin θ. (4.24)
4πr2

Substituting the above result into Equation (4.20) and noting that Equa-
tion (4.24) is independent of φ, we see that the ∇∇ operator does not con-
tribute. The electric field vector then becomes
ikr  
i
2e
E (rr) = iφ ωµkIT πa 1+ sin θ. (4.25)
4πr kr
4.1. ANALYTICAL METHODS 195

The magnetic field vector is


1
H (rr) = ∇ × E (rr)
iωµ
   2 
eikr i i
= −k 2 IT πa2 ir + 2 cos θ
4πr kr kr
  2  
i i
+iiθ 1 + + sin θ . (4.26)
kr kr

Evaluating Equation (4.25) and Equation (4.26) for a receiver, R, located


at L on the z-axis (θ = 0) gives
E (rrR ) = E (0, L) = 0 (4.27)
  2 
eikL i i
Hz (rrR ) = H (0, L) = −2 k 2 IT AT + , (4.28)
4πL kL kL
where AT is the transmitter area πa2 . After simplification, we obtain
NT AT IT 2(1 − ikL) ikL
Hz = − e , (4.29)
4π L3
for a transmitter with NT turns. A similar result would be obtained by
modeling the small current loop as a magnetic dipole, that is,
J e = 0,
K e = M δ(rr − r T ). (4.30)

The voltage induced in a receiver coil with NR turns and an area of NR


is
V = iωµ NR AR Hz , (4.31)
which is
NT AT NR AR IT 2iωµ
V = (1 − ikL)eikL . (4.32)
4π L3
The voltage is calibrated to formation conductivity using the same factor,
K, defined in Equation (4.2), such that
2i (1 − ikL)eikL
V =K . (4.33)
ωµ L2
Dividing the voltage V by K yields the complex conductivity signal, σR +
iσX .
2i (1 − ikL)eikL
σR + iσX = . (4.34)
ωµ L2
196 CHAPTER 4. MODELING OF TOOL RESPONSE

The evaluation of Equation (4.34) gives the response of an induction sonde


in homogeneous media.
In order to assess skin effect, and to compare the above exact solution to
geometrical factor theory, it is useful to expand Equation (4.34) in powers
of L/δ, where δ is the skin depth defined as δ = 2/ωµσ = (1 + i)/k. This
expansion [194] yields

2i 2L
σR + iσX = σ + 2
− σ(1 + i)... . (4.35)
ωµL 3 δ

The leading (real) term in Equation (4.35) is σ, which is the tool response
without skin effect according to geometrical theory. The second (imaginary)
term is in quadrature with the transmitter current and is independent of
conductivity. This term is the mutual inductance that exists between the
transmitter and receiver. The third term represents the conductivity de-
pendent skin effect ignored in geometrical factor theory. Note that the real
and imaginary parts of this term are equivalent. This means that after the
mutual term is removed, the X-signal provides a first order approximation
of skin effect. From Equation (4.35), the apparent conductivity signal of a
two coil sonde in a homogeneous medium can be approximated by
 2 L
σR  σ 1 − , (4.36)
3 δ
where L is the coil spacing, σ is the conductivity of the medium and δ is
the skin depth. For a multi-coil sonde, the effective length Le (see Equation
(3.12)) is used in place of L.
The mutual inductance can mask the formation dependent contribution
to σX . For example, the mutual inductance for a 40 inch coil spacing is
12.27 S/m, while formation conductivities range from from approximately
0.001 S/m to slightly greater than 5 S/m. For this reason the mutual term
is sometimes subtracted from the total response, replacing σX by σXF and
yielding an alternate form for Equation (4.34),

2i (1 − ikL)eikL − 1
σR + iσXF = . (4.37)
ωµ L2

The real and imaginary parts of Equation (4.37) are more nearly the same
order of magnitude than those of Equation (4.34). Therefore this alternate
form is often used in computer codes in order to avoid numerical inaccuracies
4.1. ANALYTICAL METHODS 197

Figure 4.6: Homogeneous medium response of two-coil sondes operating at


20 kHz with spacings of 10 inches (left) and 100 inches (right).

when modeling tool response in complex geometries. For field tools, the
cancellation of the mutual is accomplished by adjusting the turns or location
of an auxiliary bucking coil.
Figure 4.6 compares the homogeneous medium response of two-coil son-
des with spacings of 10 inches and 100 inches operating at 20 kHz. The
curves were generated using Equation (4.37). Note that the R-signal curves
for the 10 inch spacing are practically linear up to formation conductivities
of 10 S/m, while the 100 inch curves become nonlinear around 0.1 S/m, in-
dicating the dependence of skin effect on transmitter-receiver spacing. This
is to be expected, since the 100 inch sonde is significantly deeper than the 10
inch sonde, and the loss due to skin effect gets progressively greater as the
signal penetrates further into the formation. The 10 inch sonde, although
relatively free from skin effect, would necessarily be an extremely shallow
measurement of the near-borehole environment.

4.1.3 Induction real axis, spectral integration

Because analytical solutions for limiting cases are very useful for benchmark-
ing more complex numerical codes, they are described here in a fair amount
of detail. In order to arrive at solutions for modeling induction response in
the limiting 1D cases of cylindrical or planar boundaries in axisymmetric
media, Maxwell’s equations in Equation (4.10) and (4.11) are first rewritten
198 CHAPTER 4. MODELING OF TOOL RESPONSE

in the form of a partial differential equation. The induction transmitter and


receiver coils are assumed to be small dipole current loops around the z-axis
in a rotationally symmetric configuration, as shown in Figure 4.5. The coils
have an infinitesimally small radius, a.
The current density is expressed in terms of delta functions, which relate
the singular source behavior to an integral of a well-behaved function. The
current density of an infinitesimally small transmitter loop is
δ(ρ − a)
J e = IT δ(z) iφ , (4.38)
2πρ
where IT is the transmitter current and the delta functions represent the
source in cylindrical coordinates.
Because the electromagnetic fields are rotationally symmetric, ∂φ ≡ 0.
From Equations (2.47)–(2.48) we have

−∂z Hφ + iω∗ Eρ = 0 

δ(ρ − a) 
∂z Hρ − ∂ρ Hz + iω∗ Eφ = IT 2πρ δ(z)  , (4.39)
1 ∗

ρ ∂ρ (ρHφ ) + iω Ez = 0

−∂z Eφ − iωµHρ = 0  
∂z Eρ − ∂ρ Ez − iωµHφ = 0 . (4.40)
1 

ρ ∂ρ (ρEφ ) − iωµHz = 0
Fields with {Eρ , Ez , Hφ } and {Hρ , Hz , Eφ } are not coupled, and fields with
{E ρ , Ez , Hφ } are source free. Consequently we take Eρ = Ez = Hφ = 0.
The remaining equations for the field {Hρ , Hz , Eφ }
= 0 are then

δ(ρ − a) 
∂z Hρ − ∂ρ Hz + iω∗ Eφ = IT 2πρ δ(z) 

−∂z Eφ − iωµHρ = 0 
. (4.41)
1 ∂ (ρE ) − iωµH = 0 

ρ ρ φ z

Substituting gives

1 δ(ρ − a)
∂ρ ∂ρ (ρEφ ) + ∂z2 Eφ + k 2 Eφ = −iωµIT δ(z). (4.42)
ρ 2πρ
Equation (4.42) is the basic differential equation for computing induction
response in axisymmetric isotropic media. A more general form of Equa-
tion (4.42) is
1 1 δ(ρ − a)
∂ρ (ρ ∂ρ Eφ ) − 2 Eφ + ∂z2 Eφ + k 2 Eφ = −iωµIT δ(z). (4.43)
ρ ρ 2πρ
4.1. ANALYTICAL METHODS 199

σ2

σ1
R
h

T
0

Figure 4.7: Configuration for modeling planar boundaries.

The solutions for computing tool response in the presence of cylindrical


and planar boundaries are next examined in detail for the degenerate case
of a single boundary sepatating two media having different conductivities.
For these simple geometries and at induction frequencies, the solutions are
in terms of integrals that can be evaluated by real-axis integration, that is,
no contour deformation into the complex domain is necessary.

Planar boundaries

The configuration for modeling induction tool response to a single planar


boundary is shown in Figure 4.7. The collinear transmitter and receiver
are centered on the z-axis. The coils perpendicularly traverse the boundary
between two media with conductivities σ1 and σ2 . The transmitter is located
at z = 0, and the receiver is located at a distance of +z from the transmitter.
The distance from the transmitter to the boundary is denoted as h.
We define Ẽφ as the Fourier-Bessel transform of Eφ . To obtain the
spectral representation, the Fourier-Bessel transform is applied, which gives

∞
Ẽφ (kρ , z) = J1 (kρ ρ)Eφ (ρ, z)ρ dρ, (4.44)
0
200 CHAPTER 4. MODELING OF TOOL RESPONSE

∞
Eφ (ρ, z) = J1 (kρ ρ)Ẽφ (kρ , z)kρ dkρ . (4.45)
0

Referring to Equation (4.43), in an unbounded medium we then have


iωµ
∂z2 Ẽφ + kz2 Ẽφ = − J1 (kρ a)δ(z), (4.46)

with
1
kz = (k 2 − kρ2 ) 2 , (kz ) > 0. (4.47)
The corresponding Green function is

∂z2 G̃ + kz2 G̃ = −δ(z), (4.48)

eikz |z|
G̃(kρ ; z, z  ) = . (4.49)
2ikz
Consequently the E-field of a current loop in an unbounded medium is
∞
ωµ eikz |z|
Eφ (ρ, z) = J1 (kρ ρ)J1 (kρ a) kρ dkρ . (4.50)
4π kz
0

The reflected and transmitted fields satisfy

∂z2 Ẽφ + kz2 Ẽφ = 0, (4.51)

with upgoing solutions in medium 2 in the form of eikz z and downgoing


solutions in medium 1 in the form of e−ikz z . The reflected field then becomes

ẼφR (kρ , z) = R(kρ )e−ikz z , −∞ < z < h. (4.52)

And the transmitted field is

ẼφT (kρ , z) = T (kρ )eikz z , h < z < ∞. (4.53)

The reflection (R) and transmission (T ) coefficients are determined from


boundary conditions at the interface z = h. The boundary conditions to be
satisfied are the continuity of the electric field Eφ and ∂z Eφ at the boundary.
Hz is obtained from Eφ using the relationship

1 1 1
Hz = (∇ × E)z = ∂ρ (ρEφ ). (4.54)
iωµ iωµ ρ
4.1. ANALYTICAL METHODS 201

Figure 4.8: Response of 20-inch and 60-inch two-coil sondes crossing an


interface between beds with conductivities σ1 = 0.01 S/m and σ2 = 0.5
S/m.

For a transmitter and receiver located in region 1 with the receiver situated
at a distance of +z above the receiver, as shown in Figure 4.7,
∞
i kρ3  ik1z z k1z − k2z ik1z (2h−z) 
Hz = J0 (kρ ρ) e + e dkρ . (4.55)
4π k1z k1z + k2z
0

When the transmitter is in region 1 and the receiver is in region 2,


∞
i 2
Hz = J0 (kρ ρ) kρ3 eik1z h eik2z (z−h) dkρ . (4.56)
4π k1z + k2z
0

Similar expressions can be derived for Hz when the receiver is situated at


a distance of −z below the transmitter. The complex conductivity signal,
σR + iσX , for a two-coil sonde with a spacing L is computed using the
relationship
−4πiLIT
σR + iσX = Hz . (4.57)
ωµ
The response of a multi-coil sonde is computed using Equation (3.11) to
combine the responses of individual coil pairs.
To illustrate how modeling the response of a tool crossing a planar bound-
ary helps to study vertical resolution, Figure 4.8 compares computed logs
202 CHAPTER 4. MODELING OF TOOL RESPONSE

b
σ1 σ2

0 T

Figure 4.9: Configuration for modeling cylindrical boundaries.

for 20-inch and 60-inch two-coil sondes. These logs were generated by eval-
uating Equations (4.55) through through (4.57). In Figure 4.8, the planar
boundary is situated at a depth of 0.0 inches. σ1 = 0.01 S/m and σ2 = 0.5.
The the vertical resolution of the shorter 20-inch sonde is much sharper than
that of the 60-inch sonde.
The above solution has been generalized for an arbitrary number of hori-
zontal plane boundaries by Anderson and Gianzero [29], and for an arbitrary
number of dipping plane boundaries by Anderson et al. [32]. The problem of
planar boundaries was first given a rigorous treatment by Sommerfeld [239]
in his solution to the radio telegraphy problem.

Cylindrical boundaries

The configuration for modeling induction tool response to a single cylin-


drical boundary is shown in Figure 4.9. The collinear transmitter and re-
ceiver are centered on the z-axis, which is located at the center of the cylin-
der. The medium inside the cylinder has a conductivity of σ1 and and the
exterior region has a conductivity of σ2 . The radius of the cylinder is denoted
as b. This configuration can be used to study borehole effect, or invasion
effect in cases where borehole effect is negligible.
The method for solving the cylindrical boundary problem parallels the
4.1. ANALYTICAL METHODS 203

previous solution for a planar boundary. Applying a Fourier transform to


Eφ gives
∞
Ẽφ (ρ, kz ) = e−ikz z Eφ (ρ, z) dz, (4.58)
−∞
∞
1
Eφ (ρ, z) = eikz z Ẽφ (kz , z) dkz . (4.59)

−∞

Referring to Equation (4.43), in an unbounded medium we then have

1 1 δ(ρ − a)
∂ρ (ρ ∂ρ Ẽφ ) − 2 Ẽφ + kρ2 Ẽφ = −iωµIT , (4.60)
ρ ρ 2πρ
with
1
kρ = (kz2 − k 2 ) 2 , (kρ ) < 0. (4.61)
The corresponding Green function is
1 1 δ(ρ − a)
∂ρ (ρ ∂ρ G̃) − 2 G̃ + kρ2 G̃ = − , (4.62)
ρ ρ ρ
 (1)
π H1 (kρ a) J1 (kρ ρ), ρ<a
G̃(kz ; ρ, a) = − (1) . (4.63)
2i J1 (kρ a) H1 (kρ ρ), ρ>a
When a → 0,
π (1)
G̃(kz ; ρ, 0) = −
H (kρ ρ). (4.64)
2i 0
Consequently the E-field of a current loop in an unbounded medium is
∞ 
1 iωµ −π (1)
Eφ (ρ, z) = eikz z J1 (kρ a)H1 (kρ ρ) dkz . (4.65)
2π 2π 2i
−∞

The reflected and transmitted fields satisfy


1 1
∂ρ (ρ ∂ρ G̃) − 2 G̃ + kρ2 G̃ = 0, (4.66)
ρ ρ
(1)
with outgoing solutions in medium 2 in the form of H1 (kρ ρ) and incoming
(2)
solutions in medium 1 in the form of H1 (kρ ρ). The reflected field then
becomes
(2)
ẼφR (kρ , z) = R(kρ )H1 (kρ ρ), 0 < ρ < b. (4.67)
204 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.10: Response of 20-inch and 60-inch two-coil sondes in an invaded


bed with σ1 = 0.5 S/m and σ2 = 0.01 S/m plotted as a function of increasing
invasion radius.

And the transmitted field is


(1)
ẼφT (kρ , z) = T (kρ )H1 (kρ ρ), b < ρ < ∞. (4.68)

The reflection (R) and transmission (T ) coefficients are determined from


boundary conditions at the cylindrical interface ρ = b.
Because of the oscillatory behavior of the J and H Bessel functions, the
Green function is often cast in terms of the modified Bessel functions I and
K and a cosine integral, which is more efficient for performing numerical
integration. After applying the appropriate Bessel identities, the expression
for computing the complex conductivity signal for a two-coil sonde with a
spacing L becomes

−iLIT 2
σR + iσX = (1 − ik1 L)eik1 L
ωµ L3
 
2 ∞ 2 k1ρ K1 (k2ρ b) K0 (k1ρ b) − k2ρ K0 (k2ρ b) K1 (k1ρ b)
+ k1ρ cos(kz L) dkz ,
π 0 k1ρ K1 (k2ρ b) I0 (k1ρ b) + k2ρ K0 (k2ρ b) I1 (k1ρ b)
(4.69)

where the integral for the source term is written in closed-form. The response
of a multi-coil sonde is computed using Equation (3.11) to combine the re-
sponses of individual coil pairs. The numerical integration in Equation (4.69)
4.1. ANALYTICAL METHODS 205

can be efficiently performed by using the method of Filon’s weights [80] to


evaluate the cosine integral.
To illustrate how modeling tool response to a cylindrical boundary helps
to study depth of investigation, Figure 4.10 compares the response of 20-inch
and 60-inch two-coil sondes generated by evaluating Equation (4.69). In this
case, σ1 = 0.5 S/m and σ2 = 0.01, and the invasion radius is varied. The
shallower 20-inch sonde reads closer to σ1 , while the deeper 60-inch sonde is
affected more by σ2 . Both sondes read below σ1 for deep invasion because
of skin effect.
The above solution has been generalized for an arbitrary number of cylin-
drical boundaries by Gianzero and Anderson [122].

4.1.4 The induction Born response function

Section 3.2.1 showed how Born response functions were used to evaluate the
amount of signal coming from each part of the formation, but provided no
information on how these functions were derived. Born response functions
will be examined in greater detail in this section. The most rigorous deriva-
tion of the Born response function in the open literature is given in Habashy
and Anderson [135], Habashy et al. [139] and Spies and Habashy [240] and
is summarized in this section for the convenience of the reader. The mea-
sured signal is formulated in terms of a convolution-type integral over a
transformed conductivity distribution of the formation. The kernel of this
integral operator is computed beyond traditional geometrical factor theory.
Starting from Maxwell’s equations
E = J e,
∇ × H + (iω − σ)E (4.70)
∇ × E − iωµ0H = −K K e, (4.71)
the wave equation for the electric field in isotropic media is cast as
∇ × ∇ × E (rr) − kb2 E (rr) = iωµ0 δσ(rr) E (rr) + iωµ0 J e (rr) − ∇ × K e (rr), (4.72)
where
kb2 = iωµ0 σb , (4.73)
δσ(rr) = σ(rr) − σb . (4.74)
r is the observation point, with r being the source point. J e (rr) and K e (rr)
are the electric and magnetic impressed current sources, respectively. σ(rr) is
206 CHAPTER 4. MODELING OF TOOL RESPONSE

r
rR Receiver

ε(r),
r σ(r),
r µ0

Vs
εb, σb, µ0
r’
r
rT Transmitter

Figure 4.11: Configuration for modeling Born response functions.

the actual complex conductivity distribution and σb is the constant complex


conductivity of the background medium. The configuration is shown in
Figure 4.11.
The solution to Equation (4.72) is represented in terms of the dyadic
Green function Gb (rr, r  ) as

E (rr) = E b (rr) + iωµ0 drr δσ(rr ) Gb (rr, r  ) · E (rr ), (4.75)
Vs

where Vs is the support of δσ(rr). E b (rr) is the response of the source in the
background medium which satisfies the wave equation

∇ × ∇ × E b (rr) − kb2 E b (rr) = iωµ0 J e (rr) − ∇ × K e (rr). (4.76)

The dyadic Green function is governed by the equation

∇ × ∇ × Gb (rr, r ) − kb2 Gb (rr, r ) = I δ(rr − r ). (4.77)

The solution to Equation (4.77) can be represented in the form


 
1
Gb (rr, r ) = I + ∇∇ gb (rr, r ), (4.78)
kb2
4.1. ANALYTICAL METHODS 207

where gb (rr, r ) is the scalar Green function given by the expression



eikb |r −r |
gb (rr, r ) = . (4.79)
4π|rr − r  |

Under the Born approximation, the electric field inside the scatterer Vs is
approximated by the electric field of the background medium. In addition,
the electric field at observation points outside Vs is given in terms of the
dyadic Green function by the approximate expression

E (rr) ≈ E b (rr) + iωµ0 drr δσ(rr ) Gb (rr, r  ) · E b (rr ). (4.80)
Vs

The corresponding magnetic field is


  
1
H (rr) = ∇ × E (rr) ≈ H b (rr) + drr δσ(rr ) ∇gb (rr, r ) × E b (rr ). (4.81)
iωµ0
Vs

For an infinitesimally small solenoid transmitter at the position r T directed


along the z-axis, carrying a current IT with NT turns and having a cross-
sectional area AT ,

J e (rr) = 0, (4.82)
K e (rr) = −iiz iωµ0 NT IT AT δ(rr − r T ). (4.83)

The electric field in the background medium is given by



E b (rr) = iωµ0 drr Gb (rr, r  ) · (∇
∇ × K e)
VT
 
= −iωµ0 NT IT AT iz × ∇gb (rr, r T ) . (4.84)

The voltage measured at a small solenoid receiver with NR turns at the


position rR whose cross-sectional area is AR and whose axis is collinear with
that of the transmitter (on the z-axis) is

VR = iωµ0 NR AR iz · H R (rrR ). (4.85)

Substituting from Equations (4.81) and (4.84) in Equation (4.85) gives, after
some algebraic manipulation,
208 CHAPTER 4. MODELING OF TOOL RESPONSE


VR = iωµ0 NR AR iz · H b (rrR )
    
+iωµ0 NT AT IT drr δσ(rr) ∇s gb (rr, r R ) · ∇s gb (rr, r T )
Vs

ωµ0 IT
= NR AR NT AT iF (|rrR − rT |)

  
ωµ0
− dρ dz δσ(ρ, z) ρ F (|rr − rR |) F (|rr − rT |) ,
3
(4.86)
4
where ∇s is the transverse Laplacian operator
∂ ∂
∇s = ix + iy , (4.87)
∂x ∂y
and
eikb R
F (R) = (1 − ikb R) . (4.88)
R3
The receiver response in Equation (4.86) is the expression for the Born re-
sponse function in the axially symmetric case. These results are obtained un-
der the first-order Born approximation, and are valid for small contrasts and
small scatterers. An expanded version of this approximation, the extended
Born approximation, makes it possible to calculate responses at higher con-
ductivity and permittivity contrasts and larger scattering sizes. See Habashy
et al. [139] for a description of the extended Born approximation.
At induction frequencies, the Born response function is in numerical
agreement with the response functions proposed by Gianzero and Ander-
son [121], and by Moran [192]. The Doll geometrical factor can be obtained
from the Born response function by setting σ equal to zero.
Because the Born response function accurately accounts for skin effect
at moderately high frequencies, it has been applied to compute response
functions for the 2-MHz Compensated Dual Resistivity (CDR) tool. The
CDR tool broadcasts an electromagnetic wave and measures its phase shift
and attenuation between two receivers (see Section 3.3). The phase shift and
attenuation measurements are each calibrated to resistivity and displayed
as two separate curves. The attenuation measurement characterizes the
strong radial radiation of the CDR’s vertical magnetic dipole antennas by
reading deeper than the phase shift. The Born response functions are useful
4.1. ANALYTICAL METHODS 209

for visualizing how the formation conductivity level affects the volumes of
investigation of the two measurements.
Figure 4.12 and Figure 4.13 [135] show Born response functions for the
CDR tool computed in a 2 ohm-m and a 10 ohm-m formation, respectively.
Note the consistantly deeper depth of investigation of the attenuation mea-
surement in both cases, and the shallower depth of investigation of both
measurements in the more conductive 2 ohm-m formation.
The apparent conductivity signal can also be derived from Equation (4.86).
In a homogeneous formation, the receiver response is given by
iωµ0
VR = NR AR NT AT IT F (L)

iωµ0 (1 − ikb L)eikb L
= NR AR NT AT IT (4.89)
2π L3
where L = |rrR − rT | is the transmitter–receiver spacing. This is precisely
Equation (4.32). In the low frequency limit
 
iωµ0 NR AR NT AT IT iωµ0 L2
VR = 1 + σb . (4.90)
2πL3 2
From Equation (4.90) it can be seen that the quadrature (imaginary) compo-
nent is independent of the formation conductivity, while the in-phase (real)
component depends linearily on the formation conductivity. Therefore, the
receiver voltage is related to the formation conductivity by
ω 2 µ20 NR AR NT AT IT
VR = − σb , (4.91)
4πL
which is the same tool constant given in Equation (4.2) (in logging, a positive
voltage is usually assumed). From Equation (4.91), the apparent conductiv-
ity is therefore given by
4πL
σ a + i σx = − VR . (4.92)
ω 2 µ20 NR AR NT AT IT
Thus the complex apparent conductivity signal corresponding to the receiver
response of Equation (4.86) is
eikb L
σa + i σx = 2 σb (1 − ikb L)
kb2 L2
 
|rrR − r T |
+ dρ dz δσ(ρ, z) ρ3 F (|rr − r R |) F (|rr − r T |). (4.93)
2
210 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.12: Born response functions for CDR in a 2 ohm-m formation.


4.1. ANALYTICAL METHODS 211

Figure 4.13: Born response functions for CDR in a 10 ohm-m formation.


212 CHAPTER 4. MODELING OF TOOL RESPONSE

4.1.5 Laterolog response

Although electrode tools can be modeled more efficiently using numerical


methods, simple analytical expressions describing tool response are useful for
understanding basic laterolog physics, and for comparing laterolog response
with induction response in limiting situations. At low frequencies, Maxwell’s
equations for laterolog tools in isotropic media follow from Equations (2.17),
(2.18) and (2.20) as
E + J e,
∇ × H = σE (4.94)
∇ × E = 0, (4.95)
∇ · µH
H = 0, (4.96)
∇ · σE
E = −∇∇ · J e. (4.97)

In the magnetic field formulation, we find

∇ × (σ −1∇ × H ) = M e , (4.98)

where the source term is given by

M e = ∇ × (σ −1J e ). (4.99)

The magnetic field formulation allows one to model frequency effects on


laterologs, such as Groningen effect [177]. Alternatively, in the potential
formulation we use
E = −∇∇Φ, (4.100)
where Φ in the scalar potential, and find

∇ · (σ ∇Φ) = ∇ · J e . (4.101)

Equation (4.101) is known generically as Poisson’s equation. The potential


formulation is customarily used in DC laterolog modeling. All of the source
conditions are represented by boundary terms at the surface of the tool. If
the formation is axisymmetric, we assume that Φ is only a function of ρ and
z. From Equation (2.44), the differential equation for computing laterolog
response in cylindrical coordinates is then
1
∂ρ (σ ρ ∂ρ Φ) + ∂z (σ ∂z Φ) = 0. (4.102)
ρ
Laterologs consist of electrodes at fixed potentials, Vi , separated from one
another by insulating sections. The electrodes are mounted on the surface of
4.1. ANALYTICAL METHODS 213

Figure 4.14: Comparison of the potential distribution generated by a point


electrode and a finite electrode 4 inches in length and 4 inches in diameter
in a homogeneous medium with Rt = 100 ohm-m.

a cylindrical mandrel which is approximately 4 inches in diameter. Electrode


lengths range from 0.5 inches for SFL, to almost 9 feet for the A2 guard
electrode of the Dual Laterolog (see Table 3.4). Φ = Vi on the electrodes, and
the normal derivative ∂ν Φ = 0 on the insulating sections between electrodes.
In addition, the normal component of the current density is only present on
the electrode surface and vanishes at the insulating surface of the mandrel.
Because of these mixed boundary conditions, there is no closed-form so-
lution even for the canonical problem of laterolog response in a homogeneous
medium. The most common approach for obtaining an analytical solution
has been to assume that an electrode is an infinitely thin ring or a point
source and to ignore the presence of the mandrel. This approach is suffi-
ciently accurate (except in the case of extremely high Rt /Rm contrasts) for
tools whose electrode lengths are smaller than the mandrel radius, such as
normals, laterals and SFL. For tools with long electrodes, such as the Dual
Laterolog, LL3 or the High Resolution Laterolog Array, finite element or
finite difference solutions are more appropriate.
In a homogeneous medium of conductivity σ, the current I emitted from
a point electrode radiates uniformly in all directions, and the equipotential
surfaces are concentric spheres centered at the current source. The electric
214 CHAPTER 4. MODELING OF TOOL RESPONSE

field excited is [153]


I ir
E (rr) = . (4.103)
σ 4π|rr|2
∇Φ, the potential at a distance r from the current source is
Since E = −∇
I IR
Φ= = , (4.104)
4πσr 4πr
where R is the resistivity of the medium.
Figure 4.14 compares the potential created in a homogeneous medium by
a point source (Equation (4.104)) with the potential created by an electrode
on an insulating mandrel which 4 inches in diameter and 4 inches in length.
The solution for the 4-inch electrode was generated using the finite element
code LATER [119, 274]. The potentials are plotted as a function of distance
from the electrode measured along the surface of a tool. The point source
approximates the potential distribution of the 4-inch electrode at moder-
ate to large distances from the electrode, but it is not accurate enough for
modeling tools with closely spaced electrodes or long electrodes.
Analytical solutions to Equation (4.102) for thin ring electrodes with-
out a mandrel in the presence of either cylindrical or planar boundaries can
be obtained in a manner analogous to the solution of Equation (4.42) for
induction. Long electrodes are approximated by a series of thin rings that
are held at the same potential. The analytical solution for the cylindrical
boundary problem [165] has been used to study borehole and invasion effect
and anisotropy. However, the analytical solution for planar boundaries is
seldom used because laterologs are subject to a significant amount of bore-
hole effect, and therefore the presence of the borehole must be taken into
account in the model. Because of this requirement, a 2D analog computer
called the resistor network [131] (see Figure 1.4) was introduced in 1950 to
simulate laterolog response in azimuthally symmetric formations. The re-
sistor network was replaced in the early 1980’s by finite element modeling
codes [119, 177].

4.2 Numerical methods

Mixed boundary conditions cause analytical codes to become extremely com-


plicated, resulting in significantly longer computer run-times. Thus the an-
alytical methods described in the previous section are only practical for
4.2. NUMERICAL METHODS 215

modeling 1D geometries. Numerical methods are more efficient for study-


ing 2D and 3D problems, such as the combined effects of invasion and bed
boundaries in vertical or deviated wells.
Numerical methods are defined as methods which recast solutions to
partial differential equations in terms of a large number of simultaneous
linear equations. These equations are solved by matrix methods to yield the
values of electromagnetic fields at discrete points in space. The generality of
this approach makes it extremely efficient for modeling complex 2D and 3D
logging geometries. The most common numerical methods used in resistivity
modeling are finite element and finite difference techniques.
In the finite element method, Maxwell’s equations are written as inte-
gral equations which are then discretized based on the variational principle
of minimizing the total energy at points on a gridded network. The grid
is usually selected to conform to the geometry being modeled. The dis-
cretization process results in a set of large matrix equations, with non-zero
elements spread throughout the matrix because of the complex organization
of the discretization. Since the matrix equations are relatively unstructured,
their solution can be computationally slow. The main advantage of the fi-
nite element method is its ability to conform grid structures to complex
surfaces [27].
The finite difference method uses discretization based on a direct differ-
ence approximation of the differential operator form of Maxwell’s equations.
This leads to a Cartesian grid configuration. Although this grid usually does
not conform to the formation geometry, it can be made to approximate any
geometry through the use of material averaging techniques [82, 196]. The
matrix equations resulting from the discretization are usually well-structured
because of the regularity of the Cartesian grid and always sparse because
the derivatives are local operators. Thus the matrix equations can be eas-
ily solved by fast, specialized computational methods, which is the main
advantage of the finite difference method.
To illustrate the use of the finite element and finite difference methods in
resistivity modeling, two induction modeling problems that the author of this
thesis was involved in solving are used as illustrative examples. These ex-
amples are: the 2D axisymmetric finite element modeling of Dual Induction
response, and the 3D finite difference modeling of Array Induction response.
216 CHAPTER 4. MODELING OF TOOL RESPONSE

4.2.1 The finite element method

One of the first uses of the finite element method in electromagnetic model-
ing [63] was in the power industry in the late 1960’s for the solution of com-
plicated small scale problems involving generators and transformers. As the
memory capacity of computers increased during the 1970’s, the method was
extended to larger scale problems in geophysics. In resistivity logging, the
finite element method was first successfully used to model laterolog response
in 2D axisymmetric formations [119]. Shortly after this, it was applied to
compute 2D axisymmetric induction response [62, 22]. The AC induction
problem was more difficult to solve because of the additional memory re-
quired by complex matrices, and the several orders of magnitude difference
between the real and imaginary parts of the solution.
The steps involved in using the finite element method to model resistivity
tool response are outlined below. The 2D modeling of induction response
described in Chang and Anderson [62, 22] is used as an example of the
application of each step.

1. Define the problem in terms of the partial differential equation being


solved, the boundary conditions and domains of interest.
The model is an axisymmetric system in which bed boundaries are
perpendicular to the borehole and the sonde is centered on the borehole
axis. The configuration is shown in Figure 4.15. Equation (4.42) is
solved in terms of the magnetic vector potential Aφ , such that

Eφ = iωµAφ , (4.105)

and
1
Hz = Aφ + ∂ρ Aφ , (4.106)
ρ
to give 
1
∂z2 Aφ + ∂ρ ∂ρ (ρAφ ) + k 2 Aφ = 0, (4.107)
ρ
where k 2 = iωµσ (neglecting displacement current). A difference po-
tential formulation is used to overcome the numerical problems of the
singular behavior of the dipole source and the large mutual coupling
between transmitter–receiver pairs at induction frequencies. In this
formulation, AIφ denotes the incident potential in an infinite homoge-
neous medium having the conductivity of the borehole mud. Aφ is
4.2. NUMERICAL METHODS 217

z
φ

Figure 4.15: Cylindrical coordinate system.

redefined as the difference potential such that

Aφ = ATφ − AIφ , (4.108)

where ATφ is the total vector potential. The incident field accounts for
the source and mutual terms, and is expressed analytically [194] as

NT IT AT ρ
AIφ = (1 − ikm r)eikm r , (4.109)
4π r3
where km is the propagation constant of the borehole mud, and the
remaining variables are as defined at the beginning of this chapter
in Equations (4.2) and (4.3). The partial differential equation to be
solved then becomes

1
∂z2 Aφ + ∂ρ ∂ρ (ρAφ ) + k 2 Aφ − (k 2 − km
2
)AIφ = 0. (4.110)
ρ
Boundary conditions are Dirichlet.
218 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.16: Circular ring element.

2. Derive the variational integral for the problem.


The variational integral [195] derived from Equation (4.110) is
  
1
I= ρ(∂z2 Aφ )2 + (∂ρ (ρAφ ))2 − k 2 ρA2φ
ρ

+ (k −2 2
km )2ρAφ AIφ dρ dz. (4.111)

The unknown function Aφ is a solution of the differential equation if


it makes the integral I stationary.
3. Divide the space into elements.
The 3D space is divided into circular ring elements with rectangular
cross-sections as shown in Figure 4.16. To eliminate the complication
of inputting grid nodes, an automatic grid generation algorithm was
designed which is based on the estimated discretization error. Bivari-
ate linear rectangular elements are used because they conform well to
the 2D boundaries, and linear interpolation is relatively free from ar-
tifacts when applied to the oscillating and decaying induction fields.
These elements have a discretization error [242], E, which in the radial
direction is proportional to the second derivative of Aφ as

Eρ ≈ ∂ρ2 Aφ (∆ρ)2 , (4.112)

where ∆ρ denotes the radial grid size. In the axial direction, the error
and grid generation are similar to those for the radial grid, except
that ρ is replaced by z in Equations (4.112) and (4.113). The error in
Equation (4.112) is estimated analytically from AIφ (Equation (4.109))
because Aφ is unknown. A regular grid with 0.5 inches between nodes
is used in the region surrounding the tool. Away from the tool, the
4.2. NUMERICAL METHODS 219

50

40

30

20

10
z (inches)

-10

-20

-30

-40

-50
0 10 20 30 40 50 60 70
ρ (inches)

Figure 4.17: Typical finite element grid.

grid is expanded, with the new grid size ∆ρn (or ∆zn ) obtained from
the relationship

ρn−1 (dρ2 Aφ )ρn−1 (∆ρn−1 )2 = ρn (dρ2 Aφ )ρn (∆ρn )2 , (4.113)

using the previous grid points ρn−1 and ρn and previous grid size
∆ρn−1 . A typical grid is shown in Figure 4.17. Only the portion
of the grid closest to a transmitter located at z = 0 is illustrated. The
220 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.18: Element formulation.

entire grid extends several hundred feet into the formation both radi-
ally and axially. The grid generation is automatically terminated at
the point where Aφ in the formation is 15 orders of magnitude less that
Aφ in the borehole, effectively setting the exterior boundary conditions
to zero.
To model invasion boundaries without introducing additional elements,
radially inhomogeneous elements are used (the axial grid is fitted to
each bed boundary). From Equations (4.105) and (4.106), both Aφ and
∂ρ Aφ are continuous at radial boundaries. Therefore, a single element
may contain more than one material constant [242]. The element for-
mulations of inhomogeneous elements are the same as those of regular
elements, except that a step-discontinuous conductivity is introduced
when evaluating the variational integral in Equation (4.111).
4. Interpolate the unknown function in each element through the nodal
values.
The function Aφ is approximated by an interpolating polynomial pass-
ing through the nodal values in each rectangular element. The local
coordinates for the ξth element are shown in Figure 4.18. The lowest
order polynomial which can interpolate Aφ in a rectangular element is
4.2. NUMERICAL METHODS 221

a bivariate linear function. The interpolation can be written in terms


of a linear combination of basis functions as

4
Aφ (ρ̃, z̃) = Aξj Nξj (ρ̃, z̃). (4.114)
j=1

The basis functions are defined as


   
1 2ρ̃ 2z̃
Nξj (ρ̃, z̃) = 1 + (−1)int((j+1)/2) × 1 + (−1)j . (4.115)
4 ∆ρξ ∆zξ

Nξj is equal to one at node ξj and zero at all of the other three nodes.
Substituting Aφ of Equation (4.114) into the integral Equation (4.111)
and integrating over the area of the rectangle gives the element integral
Iξ , with


∆z/2 
∆ρ/2   4 2  2
 1 
4
Iξ = ρ Aξj ∂z Nξj  +  Aξj ∂ρ (ρNξj )
j=1
ρ j=1
−∆z/2 −∆ρ/2
 2 

4 
4
−k ρ 
2
Aξj Nξj  + (k 2 − km
2
)2ρAIφ Aξj Nξj dρ̃ dz̃. (4.116)
j=1 j=1

5. Perform the variational integral in each element using the interpolated


function.
The element integral of Equation (4.116) can be carried out analyti-
cally. The explicit expression is quite lengthy and is given in Chang
and Anderson [62]. The result is a quadratic function of the nodal
values of Aξj .
6. Add the variational integral over all the elements.
The overall variational integral is the sum of the element integrals as
follows 
I= Iξ . (4.117)
ξ

To introduce constraints, Equation (4.117) can also be written [242] in


matrix form as
222 CHAPTER 4. MODELING OF TOOL RESPONSE

  
A1
   
 Ke 
 A2 

I= A1 A2 A3 A4   
  A3 
ξ
A4
 
E1
  
 E2 
−2 A1 A2 A3 A4   dρ dz, (4.118)
 E3 
E4

where Ke is the element stiffness matrix assembled from the basis


functions and E are known predetermined quantities, which in this
case are used to modify boundary conditions along the borehole wall
in order to accomodate the difference potential solution.
7. Take the derivative of the variational integral with respect to every
nodal value and set it to zero, resulting in a linear matrix equation.
The objective is to find the unknown vector potential such that the
integral I is stationary. To accomplish this, the first derivatives of I
with respect to the nodal values are set to zero for all the nodes with
unknown Aφ , or
∂Aξj I = 0. (4.119)
This yields a system of linear equations in the form of

KA = E. (4.120)

These are the equations to be solved for A, and therefore all we need
to know is the stiffness matrix K and the vector E . The process of
minimizing the potential energy automatically finds the values of A
which are solutions of the partial differential equation.
8. Solve the system of linear equations after making adjustments for any
boundary conditions not incorporporated in the variational integral.
For most induction problems the dimensions of the complex matrix
equations, Nρ by Nz are of the order 40 by 2000 (the large number of
axial grid points are needed for the numerous transmitter and receiver
stations occurring over the depth span of a log). Such matrices are
often solved by sparse matrix algorithms. Because the regular rect-
angular grid generates a banded matrix, we used a block Gaussian
4.2. NUMERICAL METHODS 223

elimination algorithm to solve the system. The size of each subma-


trix to be solved depends only on the number of radial nodes and is
typically approximately 40 by 40. (Although the memory capacity
of today’s computers no longer requires such a simplification for 2D
problems, block Gaussian elimination is still sometimes used for 3D
problems [265]).
Submatrices are assembled along each set of nodes extending radially
into the formation, relating each node to the nodes directly above
and below it. The linear equations can then be written in a block
tridiagonal form as
    
B1 C1 ϕ1 E1
 D    
 2 B2 C2  ϕ2   E2 
    
 . . .  . = . .
    
 D Nz −1 B Nz −1 C Nz −1  .   . 
D Nz B Nz ϕ Nz E Nz
(4.121)
The submatrix blocks B , C and D each have dimensions of Nρ by Nρ .
The vector ϕ is the unknown vector potential such that
 
Ai, 1
 Ai, 2 
 
 
ϕi =  . . (4.122)
 
 . 
Ai, Nρ

E contains the source excitation information for the appropriate nodes


in variational form, which is used to modify the boundary conditions
for the difference potential solution. E has dimensions Nρ by NS (num-
ber of source locations). Since the left hand side of Equation (4.121)
only contains information about the formation geometry, it only needs
to be assembled once. Equation (4.121) is solved sequentially from the
bottom to the top of the formation for ϕ at each set of axial nodes using
a recursive algorithm which is described in Chang and Anderson [62].
9. Post-process the results to obtain tool response at the desired locations.
The vector potential at each receiver location is calculated by linear
interpolation of the solution at neighboring nodes. The complex appar-
ent conductivity signal is then obtained directly from Aφ by multiply-
ing by the appropriate scaling constants [194], after adding the closed
224 CHAPTER 4. MODELING OF TOOL RESPONSE

ID - Apparent resistivity

Resistivity (ohm-m)

ri = 40”

Semi-analytic solution
Finite element method

Depth (feet)

ID - X-signal
Resistivity (ohm-m)

ri = 40”

Semi-analytic solution
Finite element method

Depth (feet)

Figure 4.19: Comparison of ID response in an invaded bed computed with


the finite element method and the semi-analytic method for the apparent
resistivity signal (top) and the X-signal (bottom).

form expression for AIφ to the difference potential solution. To compute


a series of logging points, the computer software automatically moves
the sonde along the borehole axis between specified starting and stop-
ping points on the grid. Since the formation parameters do not change
for the length of the log, the coefficients to the linear equations only
4.2. NUMERICAL METHODS 225

need to be computed once. A comparison of finite element results with


results generated by the semi-analytic (mode matching) code described
in Section 4.3.1 is shown in Figure 4.19 [23] for the ID (deep) induction
tool. The agreement between the two solutions is excellent.

4.2.2 The finite difference method

The first use of the finite difference method in hydrocarbon exploration was
in the area of seismic modeling in the 1970’s. Soon afterwards, it was success-
fully employed to simulate the response of borehole acoustic logging tools.
In electromagnetics, finite difference techniques were applied to the time-
domain modeling of systems used in surface mineral prospecting in 2D [202]
and 3D [266] geometries in the 1980’s.
In the area of borehole resistivity modeling, the finite element method
is generally used more often than the finite difference method. There were
some efforts to model 2D laterolog response using the finite difference method
during the 1980’s [273, 256]. However, today’s most commonly used 2D and
3D laterolog codes (listed in the Glossary of Codes at the end of this chapter
with associated references) employ the finite element method because it can
easily handle the complex boundary conditions on electrodes.
For induction tools, the semi-analytic method (described in Section 4.3.1)
became the method of choice for 2D modeling over purely numerical tech-
niques because of its speed and efficient management of computer memory.
Current 3D finite element induction modeling efforts have been plagued by
accuracy problems at high dip angles [41]. The most successful use of the
finite difference method in resistivity logging has been the 3D modeling of
induction response [263, 81, 18], using the spectral Lanczos decomposition
method (SLDM) of Druskin and Knizhnerman.
The Schlumberger version of the 3D induction SLDM code has been
used to analyze a number of complex cases [18] where one formerly had
to rely on intuition to understand the tool response. These cases include
invasion in horizontal and highly deviated wells, anisotropy effects in invaded
dipping beds, non-circular invasion profiles, gravity segregation of invasion
and drilling-induced vertical fracture systems.
The 3D modeling results for these cases indicate that induction logs in
complex formations still have geometrical interpretations, but these interpre-
tations are much different than one has been accustomed to in vertical wells.
226 CHAPTER 4. MODELING OF TOOL RESPONSE

Invasion in horizontal wells essentially adds (or subtracts) signals from the
larger signals produced by nearby beds. Invasion in thin beds in high angle
wells is very difficult to analyze without 3D modeling because it does not
yield to “eyeball” interpretation. And when anisotropy is present in addition
to invasion, full 3D modeling becomes an absolute necessity. Non-circular
invasion profiles do not affect induction logs a great deal, and can be ignored
to first order in invasion interpretation. Drilling-induced vertical fractures
can cause array induction response to mimic invasion profiles when oil-based
mud is used, but the fractures do not significantly affect response when they
are filled with conductive mud.
The procedure for applying the SLDM finite difference technique to
model array induction response is summarized as follows. Maxwell’s equa-
tions for induction tools in 3D again follow from Equations (2.17)–(2.18) and
Equations (2.24)–(2.26) to give

∇ × (σ − iω)−1∇ × H − iωµH
H = ∇ × (σ − iω)−1J e . (4.123)

The frequency of induction tools is in the 10’s of kiloHertz range, which is


low enough so that σ ω. Therefore we have

∇ × (σ −1∇ × H ) − iωµH
H = M e, (4.124)

where the source density M e is given by

M e = ∇ × (σ −1 J e ). (4.125)

The induction code uses the SLDM technique to solve Maxwell’s equa-
tions on a staggered Cartesian (x, y, z) grid. SLDM was initially proposed
as a general solver for 3D surface electromagnetic prospecting problems in
the time and frequency domains for inhomogeneous conductive media by
Druskin and Knizhnerman [100]. A more detailed mathematical description
can be found in a subsequent paper by the same authors [101].
The application of SLDM to solving Maxwell’s equations for induction
tools can be divided into three stages:

1. Maxwell’s nonstationary equations are approximated on a spatial finite


difference staggered Yee–Lebedev grid [272]. This yields a system of
ordinary differential equations with respect to time.
4.2. NUMERICAL METHODS 227

2. The solution of the system is represented as the product of functions


of its stiffness matrix and the vector describing the initial condition.
For problems in the frequency domain, this matrix function is the
standard matrix resolvent giving the time-independent magnetic field
strength. To eliminate the null-space of Maxwell’s operator and trun-
cate spurious modes, these matrix functions are multiplied by a spec-
tral window-like projection operator, which also can be considered as
a matrix function.
3. The resulting matrix functions are approximated numerically with
SLDM. SLDM can be considered as a natural extension of the con-
jugate gradient method (for solving linear algebraic systems) to the
computation of arbitrary matrix functions.

To summarize the SLDM algorithm, we first assume that one is applying


the conjugate gradient method (CGM) to solve a problem for the magnetic
field H in the frequency domain with the spatial finite difference operator A
and a source M , with the equation in the form

A − iωII )H
(A H = M. (4.126)

Applying the finite difference approximation for the staggered Cartesian


ˆ and ∇)
grids (∇ ˇ gives

1ˆ 1ˇ
AH = ∇ × ∇×H . (4.127)
µ σ
The solution is in the form of the matrix equation

A − iωII )−1M .
H = (A (4.128)

In the SLDM, Equation (4.128) is approximated numerically by applying


the Lanczos method for generating the eigenvectors of A. The solution is
represented as a projection into the Krylov subspace K m , where

K m = span{M
M , AM , ... , Am−1M }. (4.129)

The basis vectors q 1 , q 2 , ... , q m of K m after m iterations are computed by


the Gram-Schmidt orthogonalization of the frequency independent vectors
M , AM , ... , Am−1M . The orthogonalization is carried out by a three term
recurrence algorithm, with the coefficients forming the tridiagonal matrix T ,
in the form of T = QT AQ. The orthogonal basis vectors q i form the matrix
228 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.20: The grid and coordinate system used for the SLDM finite dif-
ference calculations. The “stairstep” shows how a dipping boundary is im-
plemented in the Cartesian system.

Q. The approximate solution for the magnetic field H at the nodal points
is then obtained by solving the system

T − iωII )−1QT M .
H = Q(T (4.130)

Most of the computational effort is related to generating the matrices Q and


T . An actual value for the frequency is only used in the computation of (TT−
−1
iωII ) , but this expresion can be solved directly. For different frequencies,
the CGM uses the same Krylov subspace K m but has to be implemented
repeatedly for each new frequency. In contrast, SLDM computes a basis in
K m just once and then uses it to produce results for all values of ω using the
same projection principle. Thus SLDM solutions for multiple frequencies are
obtained for approximately the same computational cost as a CGM solution
for a single frequency.
The rectangular finite difference grid makes it relatively simple to input
formation descriptions for vertical and horizontal wells. Borehole and inva-
sion surfaces are modeled by series of small rectangular prisms. In dipping
formations, bed boundaries are modeled in staircase fashion. Figure 4.20 [18]
shows the finite difference coordinate system. A bed boundary at 45◦ to the
grid is also illustrated in this figure. To implement boundaries such as this,
we use a homogenization technique [196] that does not require the finite
4.2. NUMERICAL METHODS 229

Figure 4.21: Comparison of AIT logs computed with the SLDM code and a
analytical solution for six uninvaded beds at 70◦ dip.

difference grid to conform to the interfaces of the inhomogeneities. This ho-


mogenization technique yields a finite difference approximation for complex
geometries which is of the same quality as the finite element method, but
with much smaller computational cost.
The size of a typical induction problem is approximately 50 x 50 x 60
grid points. The grid extends out to approximately 100 ft in each direction,
with the grid size increasing as the distance from the borehole increases.
A graphical interface is used to input formations containing complex
3D features such as faults and fracture systems. The interface maps the
geometry onto the 3D grid. There is also a text-based interface for simple
subsets of 3D geometries which have uniform dip and cylindrical/elliptical
invasion. Both the text-based interface and the finite difference code are
written in standard FORTRAN 77 in order to be machine independent.
The code performance on modern workstations or high-end PC’s is sim-
230 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.22: AIT logs in an invaded horizontal well at a series of positions


above and below a sand–shale interface.

ilar: from 10 to 20 seconds per logging station for an array induction tool
consisting of a single transmitter and multiple receivers operating at three
frequencies [151]. The complex apparent conductivity signal at each receiver
location is calculated by linearly interpolating the solution for Hz at neigh-
boring nodes and multiplying by the appropriate scaling constants [194]. In
order to accurately model the large mutual signal and account for any dis-
cretization errors, the zero frequency finite difference solution is subtracted
from the solution at each receiver, and the equivalent closed form solution
is added.
Comparisons with existing 1D and 2D codes have demonstrated that
the SLDM code gives accurate results in all the limiting cases tested. Fig-
ure 4.21 [18] shows a comparison of SLDM and analytical (ANAL) results for
the AIT tool in six uninvaded beds at 70◦ dip. The values of Rt modeled are
also indicated. The analytical solution was computed with the ANISBEDS
code [137, 35]. Even though SLDM code uses homogenization at 70◦ dip,
4.2. NUMERICAL METHODS 231

Figure 4.23: AIT logs in the invaded horizontal well configuration of Fig-
ure 4.22, except with anisotropy in the lower bed.

the agreement with the analytical code is excellent. Note that there is con-
siderable shoulder effect on all of the curves except for the shallow 10-inch
log in the thickest bed, where it overshoots Rt . The extreme thinness of the
beds and the high dip angle have made this a complicated case to interpret
even without invasion.
Figure 4.22 [18] shows AIT response to invasion in an oil bearing perme-
able sand bed below a cap shale interface. This is an example of a horizontal
well case that is practically impossible to interpret without 3D modeling.
The geometry is sketched at the top of Figure 4.22. To study the relative
effects of invasion and the cap shale in the resistive oil-bearing, the sonde
remains parallel to the bed boundary and the distance between the sonde
and the boundary is varied. This approximates snapshots in time as the hor-
izontal well crosses the boundary. Invasion exists only in the permeable sand
bed and not in the impermeable shale. As the tool crossed the interface, the
invaded zone is truncated as shown in Figure 2.8. In addition to the SLDM
232 CHAPTER 4. MODELING OF TOOL RESPONSE

results, the limiting analytical solutions of invasion at an infinite distance


from the boundary (invasion only limit) and the tool crossing the boundary
with no invasion present (beds only limit) are also shown in Figure 4.22. In
the permeable bed, the 90-inch curve tracks the beds only limit, and the 10-
inch curve reads consistently near Rxo , while the 20, 30 and 60-inch curves
fall in between.
These results show that for shallow to moderate invasion such as this, the
deepest curve can be used to infer Rt , while the shallowest curve indicates
Rxo . The relative separation between the intermediate curves can only be
used to estimate the depth of invasion at a considerable distance below the
interface (in this case over 10 feet) because the tool is quite sensitive to the
more conductive shale when it is near the interface.
There is also a version of the code (called MAXANIS) that models in-
duction response in anisotropic formations using a super-staggered Lebedev
grid for arbitrarily oriented tensors [82, 81]. This code was used to analyze
how anisotropy further complicates the already difficult interpretation of in-
vasion in a horizontal well. Figure 4.23 [18] shows the same configuration
as Figure 4.22, but with anisotropy added in the lower bed outside of the
invaded zone. The invaded zone remains isotropic. The limiting case of an
anisotropic bed without invasion (beds only limit) is also shown. The be-
havior of the five AIT curves in Figure 4.23 is similar to Figure 4.22; the
deepest curve tracks the beds only limit and the shallowest curve reads Rxo ,
while the three other curves fall in between. However, in Figure 4.23, the
beds only limit in the lower bed is now an average of Rh and Rv . If the lines
indication Rh and Rv were removed, it would be dificult to determine that
anisotropy was present based on the behavior of the resistivity curves alone.
In addition to providing valuable interpretation insights, the 3D finite
difference code has made it possible to validate new inversion algorithms
in a wide variety of complex logging environments. It has been used to
test Merlin processing [49], an automatic high angle inversion method for
layered media with invasion. The code has also been extended to 2-MHz
LWD resistivity tools [28]. Since displacement currents cannot be neglected
at 2-MHz, this code has a slower computation speed.
When faster approximate forward modeling codes are developed (such as
the extended Born approximation [255]), the speed and accuracy of this 3D
code will allow them to be benchmarked. However, with computer speed in-
creasing on a yearly basis, the routine use of full 3D finite difference modeling
4.3. HYBRID METHODS 233

in appropriate situations is also a real possibility.


The MAXANIS code also models arbitrarily oriented magnetic and elec-
tric dipoles in anisotropic media. This code allows the testing of new an-
tenna systems with sensitivity to both direction and anisotropy in realistic
formations. The next generation of resistivity tools will yield much more
information than Rt . They will also give information about directional con-
ductivity and formation geometry. Efficient 3D modeling will play a major
part in the design of these new tools.

4.3 Hybrid methods

Hybrid techniques typically break problems into two parts, one of which is
solved analytically and the other numerically. This leads to codes that can
quickly model tool response in fairly complicated geometries. The 2D ax-
isymmetric modeling of induction or laterolog response in a borehole through
horizontal invaded beds is an example of a problem that has been successfully
solved by a hybrid method. The partial differential equation that governs
the tool response is reduced to two 1D problems—one in the radial plane and
the other in the vertical plane. Then the radial distribution of the field is
treated numerically by the finite element technique, and vertical distribution
is treated analytically by modal analysis. The main advantages of hybrid
codes is that they typically run much faster than equivalent finite element
or finite difference calculations, and they do not require the solution of large
systems of linear equations.

4.3.1 Fast semi-analytic (mode matching)

During the 1980’s, the growing use of modeling in interactive log interpreta-
tion created a demand for fast 2D modeling codes. Although finite element
and finite difference codes were run routinely on mainframe computers at
research and engineering centers, these codes were extremely impractical
to run on the microcomputers commonly used at field log interpretation
centers. Integral equation solutions of that time [120, 276] ran rapidly for
the degenerate case of a borehole and two invaded beds, but computer run
times increased dramatically when more beds were added to the model (re-
cent work by Abubakar overcomes this problem using a conjugate gradient
fast Fourier transform technique [1].)
234 CHAPTER 4. MODELING OF TOOL RESPONSE

In 1984, the numerical mode matching method was shown by Chew,


et al. [70, 23] to be a computationaly efficient technique for modeling 2D
induction response on mid-sized to small computers. Mode matching was
also applied to the computation of laterolog response [25, 257, 173], but with
less success because some computational efficiency is lost in the simulation
of long electrodes.
In the mode matching method, a borehole passing through a bed bound-
ary is viewed as the junction of two open waveguides. For any wave incident
on the boundary, the reflected and transmitted waves are determined by the
requirement that the tangential electric and magnetic fields are continuous
across the plane of the bed boundary. This condition can be expressed as an
integral equation for the field in the plane of the bed boundary. Mode match-
ing is one of several ways that has been used to extract useful information
from this equation.
The most significant technical complication of the junction problem is
the treatment of the continuous spectrum. Pudensi and Ferriera [204] have
shown how the continuum modes, also known as radiation modes, can be
systemmatically approximated by a set of discrete modes using Hermite
functions which form a complete set on an infinite interval. In their work,
this technique is adapted to cylindrical structures and further developed by
using more general expansion functions. The discrete modes are obtained
by solving the differential equation first, eliminating the need to formulate
and solve an integral equation. The method applies to an inhomogeneous
waveguide of arbitrary profile.
Chew et al. [70] adapted the mode matching technique to model borehole
logging tool response. In addition to simulating the response of conventional
induction (vertical magnetic dipole) and laterolog tools, it has been used to
investigate vertical electric dipole tools [70]. Because of its computational
efficiency, mode matching has also been applied to calculate time-domain
response, and it has been used to investigate several theoretical transient
induction tools for borehole logging [26]. Only the diffraction of axially
symmetric waves is considered here, with the diffraction of non-axisymmetric
waves being a topic of additional research [174, 67].
The geometry under consideration is shown in Figure 4.24. The applica-
tion of mode matching to the modeling of induction response will be sum-
marized first. Then modifications necessary for modeling laterolog response
will be briefly outlined.
4.3. HYBRID METHODS 235

N
dN
2a N-1
4 d4
2b 3
d3
2
d2
1
d1
ρ’, z’ 0
ρ
d-1
-1 d-2
-2
-M+1 d-M
.
-M

Figure 4.24: Geometry and notation used for the induction numerical mode
matching code. The transmitter is a current loop antenna on a metallic
mandrel centered in a borehole which penetrates an arbitrary number of
invaded beds.

From Maxwell’s equations, the vector wave equation satisfied by the


induction magnetic field is

∇ × µ−1 ∇ × E − ω 2 µ∗ E = iωµJ


µ∇ J e, (4.131)

where ∗ is the complex permittivity equal to +iσ/ω. Under the assumption


of axial symmmetry, the electric field E satisfies the differential equation

1 1
ρµ ∂ρ∂ρ + µ ∂z ∂z + ω 2 µ∗ ρ Eφ = −iωµ I e ρ δ(ρ − ρ ) δ(z − z  ),
ρµ µ
(4.132)
for a current loop of radius ρ located at the depth z  . For the geometry
of Figure 4.24, we assume that  and µ are arbitrary functions of ρ while
they are independent of z in each region. Therefore it is more expedient to
236 CHAPTER 4. MODELING OF TOOL RESPONSE

find eigenfunctions which are the homogeneous solutions of Equation (4.132)


that satisfy 
1 2 2 ∗
ρµ ∂ρ ∂ρ + ∂z + ω µ F (ρ, z) = 0. (4.133)
ρµ
To find the solution of Equation (4.133), we apply separation of variables
and assume that

F (ρ, z) = fα (ρ)eikαz z aα , (4.134)
α

where aα is a constant independent of ρ and z. It then follows that



1
ρµ ∂ρ ∂ρ + ω 2 µ∗ − kαz
2
fα (ρ) = 0. (4.135)
ρµ

Equation (4.135) is a one dimensional Sturm-Liouville equation to be solved


2 the
numerically, with the eigenfunction fα (ρ) being the solution, and kαz
eigenvalue. The boundary conditions to be satisfied by fα (ρ) are

fα (ρ) = 0, when ρ → ∞
1
fα (ρ) = 0, when ρ = 0 . (4.136)
ρ

If a metallic mandrel of radius a is modeled, the tangential electric field is


required to be zero at the mandrel surface. Thus we have the additional
boundary condition &
&
1 &
fα (ρ)& = 0. (4.137)
ρ &
ρ=a

To obtain the solution to Equation (4.135), we first choose a set of basis


functions gn (ρ) which is complete over the interval (a, +∞), so that we can
expand


fα (ρ) = bαn gn (ρ). (4.138)
n=1

The boundary condition on gn (ρ) on the mandrel surface is


&
&
1 &
gn (ρ)& = 0. (4.139)
ρ &
ρ=a

The above boundary condition implies the boundary condition of Equa-


tion (4.137) if we are working with a finite summation in Equation (4.138).
4.3. HYBRID METHODS 237

Substituting the first N terms of Equation (4.138) into Equation (4.135)


gives
∞ 
1
bαn ρµ ∂ρ ∂ρ + ω 2 µ∗ − kαz
2
gn (ρ) = 0. (4.140)
n=1
ρµ

Multiplying by (ρµ)−1 gm (ρ) and integrating from a to ∞ eliminates the ρ


dependence in Equation (4.140). Defining the inner product as

' ( ∞
1
f, g = dρ f (ρ) g(ρ), (4.141)
ρµ
a

Equation (4.140) becomes



  
bαn Bm, n − kαz
2
Gm, n = 0, (4.142)
n=1

where ' ( ' (


1
Bn, m = gm , ρµ ∂ρ ∂ρ gn + ω 2 gm , µ∗ gn , (4.143)
ρµ
and ' (
Gn, m = gm , gn . (4.144)

Using integration by parts, it can be shown that


∞ ∞
1   µ∗
Bn, m = − dρ g (ρ) gm (ρ) + ω 2 dρ gn (ρ) gm (ρ), (4.145)
ρµ n ρµ
a a

and
∞
1
Gn, m = dρ gn (ρ) gm (ρ), (4.146)
ρµ
a

where the primes indicate derivatives with respect to the argument of the
functions. With the definition of the inner product in Equation (4.141),
Bn, m and Gn, m are symmetric tridiagonal matrices.
Triangular functions were chosen for gn (ρ) because they are computation-
ally efficient while giving extremely accurate results. The triangular element
configuration is shown in Figure 4.25 [70]. We also tested Gauss-Hermite
functions. Although fewer Gauss-Hermite functions are needed to approxi-
mate the radial fields, the integrations of Equations (4.145) and (4.146) must
238 CHAPTER 4. MODELING OF TOOL RESPONSE

fα(ρ)

a ρn-1 ρn ρn+1 ρ

Figure 4.25: Piecewise linear triangular functions for approxinating the ra-
dial field.

be performed numerically. Triangular functions are assumed to be a piece-


wise linear approximation of the fields in the radial direction. This allows
the integrations to be performed analytically, making it possible to compute
the matrix elements more efficiently. Since induction fields are exponen-
tially small a few skin depths from the borehole, around 40 basis functions
are adequate for approximating fields radially. A small step size is used for
the triangular functions close to the borehole. The step size is gradually
increased away from the borehole, in a manner similar to the radial finite
element grid shown in Figure 4.17.
The diagonal matrix elements Gn, n and Bn, n and the matrix elements
on either side of the diagonal are defined in closed form as follows,
 ) *  ) *
1 ρn−1 ρ2n−1 ρn 1 ρn+1 ρ2n+1 ρn
Gn, n = − + 2 ln − + + 2 ln ,
2 δn δn ρn−1 2 δn+1 δn+1 ρn+1
) * (4.147)
1 ρn+1 ρn+1 ρn ρn+1
Gn, n+1 = Gn+1, n = − + − 2 ln , (4.148)
2 δn+1 δn+1 ρn
 ) *
1 ρn−1 ρ2n−1 ρn
Bn, n = ω 2
µ∗n − + 2 ln
2 δn δn ρn−1
 ) *
1 ρn+1 ρ2n+1 ρn
−ω 2 µ∗n+1 + + 2 ln
2 δn+1 δn+1 ρn+1
 ) * ) *
1 ρn 1 ρn+1
− 2 ln + 2 ln , (4.149)
δn ρn−1 δn+1 ρn
4.3. HYBRID METHODS 239

 ) *
1 ρn+1 ρn+1 ρn ρn+1
Bn, n+1 = Bn+1, n = ω 2
µ∗n+1 − + − 2 ln
2 δn+1 δn+1 ρn
) *
1 ρn+1
+ 2 ln , (4.150)
δn+1 ρn

where ρn is the radial coordinate and δn is the step size of the nth triangular
element. These simple expressions can be evaluated rapidly. All of the
remaining matrix elements are zero.
2 are obtained by truncating Equation (4.142) and
The eigenvalues kαz
solving  
G−1 · B − kαz
2
I · bα = 0. (4.151)
The eigenvectors bα are also obtained from Equation (4.151), which are then
used to derive the eigenfunctions in Equation (4.138). If N basis functions
are used in Equation (4.138), Equation (4.151) will produce N eigenvalues
and N eigenvectors.
The eigenvectors obtained by solving Equation (4.142) satisfy G orthog-
onality, that is,
bTα · G · bβ = δαβ Dα . (4.152)
Having obtained the eigenvectors and eigenvalues, we can systematically
solve Equation (4.132), which yields

ωI e 
N
fα (ρ )fα (ρ) ikαz |z−z  |
ρEφ = − e (4.153)
2 α=1 kαz Dα

In the presence of bed boundaries, we match boundary conditions and de-


termine the reflection and transmission operators in matrix representation.
The generalized reflection operator is

Ri+1, i+2 · T i+1, i · e2ik i+1 z (di+2 −di+1 )


T i, i+1 · R̃
Ri, i+1 = R i, i+1 +
R̃ , (4.154)
I − R̃Ri+1, i+2 · R i+1, i · e2ik i+1 z (di+2 −di+1 )
with
RN, N +1 = 0,
R̃ (4.155)
R−M, −M +1 = 0,
R̃ (4.156)
and
D i · k i z − bi · b−1 bi · b−1
j · D j · k j z · (b j )
T
R i, j = , (4.157)
D i · k i z + bi · b−1 bi · b−1
j · D j · k j z · (b j )
T
240 CHAPTER 4. MODELING OF TOOL RESPONSE

T i, j = (bbi · b−1
j ) · (I
T
I + Ri, j ). (4.158)
The general formulation of the field (in region zero) is

ωI e T 
ρEφ = − f 0 (ρ) · eik 0 z |z−z |
2
 
eik 0 z (z−d−1 ) · R̃
R0, −1 · (eik 0 z (z −d−1 ) + eik 0 z (d1 −d−1 ) · R̃
R0, 1 · eik 0 z (d1 −z ) )
+
I − e−ik 0 z d−1 R̃ R0, −1 eik 0 z (d1 −d−1 ) R̃
R0, 1 eik 0 z d1
 

e−ik 0 z (z−d1 ) · R̃
R0, 1 · (eik 0 z (d1 −z ) + eik 0 z (d1 −d−1 ) · R̃
R0, −1 · eik 0 z (z −d−1 ) )
+
R0, 1 eik 0 z (d1 −d−1 ) R̃
I − eik 0 z d1 R̃ R0, −1 e−ik 0 z d1
· (kk 0 z · D 0 )−1 · f 0 (ρ ). (4.159)

In the above, bi is a column vector containing the eigenvectors bα , f i is a


column vector containing the eigenfunctions fα , k i z and D i are diagonal
matrices containing kαz and Dα , eik i z is a propagator and R i, j and T i, j
are reflection and transmission matrices. A more detailed derivation of the
expressions for computing the field in any bed is given in papers by Chew
et al. [69, 73].
The complex apparent conductivity signal is obtained from ρEφ by mul-
tiplying by the appropriate scaling constants [194]. For modeling induction
response, an additional step is necessary to accurately calculate both the
resistive and reactive parts of the complex tool signal because of the large
mutual contribution to the reactive signal. In this case, the matrix represen-
tation of the solution in any one bed is subtracted from the entire solution,
and a more accurate analytical solution for cylindrical media (such as real
axis integration) is added back in its place. This modification is not neces-
sary for propagation tools such as 2-MHz LWD tools, since both the real and
imaginary field components are approximately the same order of magnitude
because of their higher operating frequencies.
The excellent agreement of the semi-analytic results with results gen-
erated by the 2D finite element code shown in Figure 4.19 also serves to
validate the accuracy of the numerical mode matching method. The mode
matching codes for modeling induction and 2-Mhz tool response have been
used extensively in log interpretation [11]. These codes were used to gener-
ate all of the induction logs in the benchmark formation shown in Chapter
3. Run times on a PC or modern workstation are two to three minutes for
each log.
4.3. HYBRID METHODS 241

For electrode tools (laterologs), the potential Φ satisfies the differential


equation

1 1 I δ(ρ − ρ ) δ(z − z  )
∂ρ σρ ∂ρ + ∂z σ ∂z Φ = − , (4.160)
σρ σ 2πρσ(ρ )
where σ is the conductivity. Equation (4.160) is Poisson’s equation in an
inhomogeneous axisymmetric medium. The current source is modeled as
a thin magnetic current loop encircling an insulating mandrel. Nonzero
frequency effects are ignored. For the geometry of Figure 4.24, it is assumed
that σ is an arbitrary function of ρ while it is independent of z in each region.
Similar to induction tools, we find eigenfunctions which are the homogeneous
solutions of Equation (4.160) that satisfy

1 1
∂ρ σρ ∂ρ + ∂z σ ∂z F (ρ, z) = 0. (4.161)
σρ σ
As before, we apply separation of variables and assume that

F (ρ, z) = fα (ρ)eikαz z aα , (4.162)
α

where aα is a constant independent of ρ and z. It then follows that



1
∂ρ σρ ∂ρ − kαz
2
fα (ρ) = 0. (4.163)
σρ
The boundary conditions to be satisfied by fα (ρ) are

fα (ρ) = 0, when ρ → ∞
1
fα (ρ) = 0, when ρ = 0 . (4.164)
ρ
With an insulating mandrel of radius a, the current at the surface of the
mandrel is required to be zero. In this case, we have the additional boundary
condition &
&
&
∂ρ fα (ρ)& = 0. (4.165)
&
ρ=a

To obtain the solution to Equation (4.163), we choose a set of basis functions


gn (ρ) which can reasonably approximate fα (ρ), so that we can expand


fα (ρ) = bαn gn (ρ). (4.166)
n=1
242 CHAPTER 4. MODELING OF TOOL RESPONSE

The boundary condition on gn (ρ) on the mandrel surface is


&
&
&
∂ρ gn (ρ)& = 0. (4.167)
&
ρ=a

The above boundary condition implies the boundary condition of Equa-


tion (4.165) if we are working with a finite summation in Equation (4.166).
Substituting the first N terms of Equation (4.166) into Equation (4.163)
gives
∞ 
1
bαn ∂ρ σρ ∂ρ − kαz
2
gn (ρ) = 0. (4.168)
n=1
σρ
Multiplying by σρ gm (ρ) and integrating from a to ∞ eliminates the ρ de-
pendence in Equation (4.168). Defining the inner product as

' ( ∞
f, g = dρ σρ f (ρ) g(ρ), (4.169)
a

Equation (4.168) becomes



  
bαn Bm, n − kαz
2
Gm, n = 0, (4.170)
n=1

where ' (
1
Bn, m = gm , ∂ρ σρ ∂ρ gn , (4.171)
σρ
and ' (
Gn, m = gm , gn . (4.172)
Using integration by parts, it can be shown that
∞
Bn, m = − dρ σρ gn (ρ) gm

(ρ), (4.173)
a

and
∞
Gn, m = dρ σρ gn (ρ) gm (ρ). (4.174)
a

With the definition of the inner product in Equation (4.169), Bn, m and Gn, m
are symmetric tridiagonal matrices.
4.3. HYBRID METHODS 243

Triangular functions were also chosen for gn (ρ) for modeling electrode
tool response, allowing the integrations to be performed analytically. Around
45 basis functions are needed to accurately approximate the radial fields.
The diagonal matrix elements Gn, n and Bn, n and the matrix elements on
either side of the diagonal can be defined in closed form and are given in
Anderson and Chew [25]. All of the remaining matrix elements are zero.
2 are obtained by truncating Equation (4.170) and
The eigenvalues kαz
solving  
G−1 · B − kαz
2
I · bα = 0. (4.175)

The eigenvectors bα are also obtained from Equation (4.175), which are then
used to derive the eigenfunctions in Equation (4.166).
The eigenvectors obtained by solving Equation (4.170) satisfy G orthog-
onality, that is,
bTα · G · bβ = δαβ Dα . (4.176)
Having obtained the eigenvectors and eigenvalues, we can systematically
solve Equation (4.160), which yields

I  N
fα (ρ )fα (ρ) ikαz |z−z  |
Φ=− e (4.177)
4πi α=1 kαz Dα

Accuracy can be increased by taking more terms in the expansion of Equa-


tion (4.166). From this point onward, the treatment of the multiple bed
reflection and transmission operators for electrode tools is the same as is
given in Equations (4.154) through (4.159) for induction tools.
For a laterolog tool with multiple current and voltage electrodes, the
potentials generated by each current source at each voltage electrode are
inserted in a transfer impedance matrix. This matrix, in combination with
the monitoring and measurement conditions is then used to solve for the tool
response in the conventional manner as described in Section 3.7.4.
Figure 4.26 [173] shows a comparison of SFL response computed using
the numerical mode matching method with results generated by a 2D finite
element code [119, 274]. The excellent agreement serves to validate the
accuracy of mode matching for computing the response ef electrode tools
with thin ring electrodes. The mode matching program is eight times faster
than the finite element code for computing SFL response. For computing
the response of tools with log electrodes such as the Dual Laterolog, the
244 CHAPTER 4. MODELING OF TOOL RESPONSE

1000

SFL
Rt

100
Resistivity (ohm-m)

Finite element

Mode matching

Rxo
ri=20”
10 ri=12”

Rm=1 ohm-m
rb=4”

1
-32 -24 -16 -8 0 8 16 24 32
Depth (feet)

Figure 4.26: Comparison of finite element and mode matching results for
SFL in a four bed formation with invasion in the two cemtral beds.

run times for the two codes become equal because a series of thin rings are
needed to approximate the long electrodes.

4.3.2 With/without skin effect hybrid

Before the development of the numerical mode matching method, the hybrid
method of Kaufman [157] was sometimes used to model induction response
in 2D axisymmetric formations. This method is still used today when a fast
approximate solution is needed because the computer program is relatively
easy to implement. Kaufman’s hybrid method uses the planar layerd model
to treat skin effect in the far field where it is greatest, combining this with
Doll’s geometrical factor theory to compensate for the contributions of the
borehole and invaded zones.
To compute tool response in an invaded axisymmetric formation such as
the three beds shown in Figure 4.27, the planar layered solution with skin
effect (without borehole and invasion) is first calculated with a 1D spectral
integral code using σti for the bed conductivities. This response is then ad-
justed to account for the presence of the borehole and invasion by adding
4.3. HYBRID METHODS 245

σt1 rb

ri

σt2 σxo2 σm

σt3
ρ

Figure 4.27: Geometry and notation used for modeling induction response
with the hybrid method.

the geometrical factor contributions for each cylindrical region. Each geo-
metrical factor contribution is multiplied by the difference in conductivity
between the region itself and the adjacent exterior region, which systemat-
ically accounts for the volumetric change in conductivity. Using the three
beds in Figure 4.27 as an example, the procedure is given by
−∞
 rb
σHybrid = σM ulti−layer + (σm − σt1 ) g(ρ, z) dρ dz
d1 0
d1rb d1ri
+(σm − σxo2 ) g(ρ, z) dρ dz + (σxo2 − σt2 ) g(ρ, z) dρ dz
d2 0 d2 0
 rb
+∞
+(σm − σt3 ) g(ρ, z) dρ dz, (4.178)
d2 0

with g(ρ, z) as defined in Equation (4.3). It is only necessary to evaluate


the z integration numerically in the computer code; the ρ integration can be
performed analytically which saves a considerable amount of computer time.
246 CHAPTER 4. MODELING OF TOOL RESPONSE

Figure 4.28: 6FF60 logs computed with the hybrid method and the numerical
mode matching method.

The procedure in Equation (4.178) can easily be generalized for an arbitrary


number of beds and invaded zones.
A comparison of 6FF40 response computed with the Kaufman hybrid
method and the numerical mode matching method is shown in Figure 4.28 for
the invaded benchmark formation. The agreement between the two curves
is excellent in the lower resistive invaded bed, but the curves separate in
the upper conductive bed where there is significant skin effect. Thus the
hybrid method is limited to cases where the contribution of skin effect in the
invaded zone is negligible, i.e., shallow invasion of low conductivity. Since
geometrical factor theory only compensates the resistive signal, the reactive
signal is unchanged from the planar layered model.
4.4. GLOSSARY OF COMPUTER CODES 247

4.4 Glossary of computer codes

The following computer codes were either used to generate results shown in
this thesis, or were referred to in the text. References are provided in all
cases.

4.4.1 Induction codes

HSE An analytical code for computing induction response with skin effect
in homogeneous media. It is based on a closed form expression in Moran
and Kunz [194], which is also given in Equation (4.34).
VGF, RGF and IRGF Analytical codes for computing Doll’s vertical,
radial and integrated radial geometrical factors, respectively. Closed form
formulas are derived in Doll’s paper [88]. Because this paper contains many
typos, the corrected formulas used in the three codes are given in Equa-
tion (4.6), Equation (4.7) and Equation (4.9).
VGFSE, RGFSE and IRGFSE Analytical codes for computing verti-
cal, radial and integrated radial response functions, respectively, with skin
effect. Formulas are given in Gianzero and Anderson [121] for generating
both resistive and reactive components of the response functions. The three
codes involve the numerical evaluation of integrals.
MAPSE Analytical code for computing a 2D map of induction response
functions with skin effect as a function of ρ and z. The derivation is given
in Gianzero and Anderson [121].
CDRGF Analytical code for computing a 2D axisymmetric map of Born
response functions for 2-MHz tools. The derivation is given in papers by
Habashy [135, 240] and summarized in Section 4.1.4.
THICK The first Schlumberger 1D analytical code for computing induc-
tion response in concentric cylindrical media, written in 1957. The problem
is solved by numerically evaluating an integral involving Bessel functions,
which is given in Moran and Kunz [194]. Point dipole antennas are assumed.
THICKX A 1D spectral integral code which computes induction response
in an arbitrary number of concentric cylindrical media. Antennas are thin
rings of arbitrary radius surrounding a metal mandrel. The code was written
by Gianzero and Anderson [122] in 1973. It is sufficiently general that it can
248 CHAPTER 4. MODELING OF TOOL RESPONSE

be used for modeling tools operating from tens of hertz to tens of megahertz.
THIN The first Schlumberger 1D analytical code for computing induction
response in a thin bed with symmetrical shoulder-beds, written in 1958. The
problem is solved by numerically evaluating an integral involving exponential
functions, which is given in Moran and Kunz [194]. Point dipole antennas
are assumed.
ISMLM A 1D spectral integral code which computes induction response
in an arbitrary number of planar layered media with the sonde logging per-
pendicular to the bedding planes. Point dipole antennas are assumed. The
code was written by Anderson and Gianzero [29] in 1975.
ISMLMDIP A 1D spectral integral code for computing induction and 2-
MHz tool response in an arbitrary number of dipping beds for dip angles
ranging from 0◦ to 90◦ . Tool response is formulated as a superposition of
vertical and horizontal magnetic dipoles. The code was written by Anderson
and Habashy [32] in 1985.
HYBIND A 2D hybrid code for approximating induction response in ax-
isymmetric formations. A planar layerd model (such as ISMLMDIP or ANIS-
BEDS) is used to treat skin effect in the far field where it is greatest. The
contributions of the borehole and invaded zones are modeled with Doll’s ge-
ometrical factor theory, and the two sets of responses are combined. The
method was proposed by Kaufman [157].
TRIKHZ A 2D semi-analytic code for computing induction response in
axisymmetric formations. The vertical part of the problem is solved analyti-
cally using a finite number of discrete modes to describe the electromagnetic
waves, and the radial part of the problem is solved numerically using piece-
wise linear functions. The code was written by Chew and Anderson [70, 23]
in 1983. It models tool response in three beds, with an arbitrary number of
invaded zones in each bed.
INDINV A 2D semi-analytic code for computing induction response in
axisymmetric formations. It is an extended version of TRIKHZ, allowing
an arbitrary number of beds. The code was written by Liu and Nie [73]
in 1989. Another version of the code, CDRINV, computes the response of
2-MHz LWD tools.
FEMIND A 2D finite element code for computing induction response
in axisymmetric formations. The code was written by Chang and Ander-
4.4. GLOSSARY OF COMPUTER CODES 249

son [62, 22] in 1980. It uses block-Gaussian elimination to efficiently solve


for multiple tool positions.
ANISBEDS A 1D spectral integral code for computing induction and 2-
MHz tool response in an arbitrary number of dipping anisotropic beds. TI
(transversely isotropic) anisotropy is assumed. The code was written by
Lüling and Habashy [137, 180, 35] in 1994. A general version of the code
computes all existing electric and magnetic field components for vertically
and horizontally oriented dipoles.
XBED An analytical code for computing induction and 2-MHz tool re-
sponse in layered media with crossbedding anisotropy. The crossbedding is
described by a uniaxial conductivity tensor whose principal axes have dif-
ferent arbitrarily oriented dip and strike angles in each bed. The code was
written by Gianzero and Anderson [34] in 1997. The current version of
the code computes the response of a tool logging perpendicular to the bed
boundaries. Work is in progress to model arbitrary logging angles.
SLDMINV A 3D finite difference code for computing induction response in
arbitrary geometries. The code was developed by Druskin, Lee and Ander-
son [18] in 1995. It was derived from a more general surface electromagnetic
prospecting code by Druskin and Knizhnerman [101]. It uses the spectral
Lanczos decomposition method to solve Maxwell’s equations on a staggered
Cartesian grid.
MAXANIS A 3D finite difference code which computes induction response
in anisotropic media. The code was developed by Druskin and Davydy-
cheva [18] in 1995. It uses a super-staggered Lebedev’s grid for arbitrarily
oriented anisotropy tensors [82, 81].

4.4.2 Laterolog codes

LEP A 1D analytical code for computing laterolog response in concentric


cylindrical media. The problem is solved by numerically evaluating an in-
tegral involving Bessel functions. The first version of the code, written by
Moran and Timmons [190, 165] in 1957, modeled thin ring electrodes on an
insulating mandrel. A modification by Regat in 1966 allowed electrodes of
arbitrary vertical extent (long electrodes).
Resistor Network A 2D analog computer [131] that was used to simu-
late laterolog response in azimuthally symmetric formations between 1950
250 CHAPTER 4. MODELING OF TOOL RESPONSE

and 1980, when it was replaced by LATER. The network contained approx-
imately 300,000 resistors arranged in interchangeable panels, representing
formations with a height of approximately 200 feet and a radius of 150 feet.
LATER A 2D finite element code for computing laterolog response in az-
imuthally symmetric formations. A quasi-uniform rectangular grid in the ρz
plane is used. The code was written by Zamansky [119, 274] in collaboration
with the École des Mines, Paris, in 1978.
SKYLINE A 2D finite element code for computing laterolog response in
a 3D geometry consisting of cylindrical wedges. This configuration can be
sheared to account for dip, and the tool can be eccentered in the borehole.
The code was written in collaboration with the École des Mines, Paris, in
1982 and upgraded by Gounot [127].
CWNLAT A 2D finite element package which can solve for a variety of
scalar potentials, including Hφ and Φ in an azimuthally symmetric medium,
or Hz and Φ in the xy plane. The code was written by Lovell [177] in 1989.
Because the code models an AC current source, it is useful for studying
Groningen effect.
LATNMM A 2D semi-analytic code for computing laterolog response in
axisymmetric formations with an arbitrary number of beds. It uses the same
numerical mode matching technique as INDINV. The code was written by
Liu, Anderson and Chew [173] in 1992.
SIMULOG2D A fast 2D finite element code for computing laterolog re-
sponse in axisymmetric formations. It was written by Legendre [128] in 1997
for use with inversion software in place of the slower LATER code. The code
uses a triangular mesh with domain averaging. A complete Cholesky factor-
ization is used to solve the linear systems.
ALAT3D A 3D finite element code which solves for Φ in more or less
arbitrary geometries. The code was written by Lovell [177] in 1993. The
basis functions are linear on pentahedral elements, with the option of adding
tetrahedral nodes on interfaces between bed boundaries.
LL3D A 3D finite element code for computing laterolog response in arbi-
trary geometries. The code was written by Wang [265] in 1997. To give
stable results at dip angles from 0◦ to 90◦ , the solution region is divided into
tetrahedra according to a cylindrical coordinate system. The tetrahedra are
further subdivided if they are cut by boundaries. There is also a version of
4.4. GLOSSARY OF COMPUTER CODES 251

the code for anisotropic media.


DC3D A 3D finite difference code for computing laterolog response in
arbitrary geometries. The code was written by Lee, Druskin and Habashy
in 1998. The tool is modeled using impedance boundary conditions. Material
averaging is used to improve accuracy for high resistivity contrasts [82, 196].
252 CHAPTER 4. MODELING OF TOOL RESPONSE
Chapter 5

Using modeling methods in log


interpretation

Summary: Up to this point, the focus has been on the physics of resistivity log-
ging tools and on methods for modeling their response. This chapter addresses
the practical use of resistivity measurements, describing the interpretation of re-
sistivity logs for the purpose of evaluating the amount of hydrocarbons present in
a reservoir. Basic relationships between resistivity measurements and rock physics
are outlined. The use of correction charts is described, and simple inversions based
on iterative forward modeling are illustrated using log examples. This interpreta-
tion overview demonstrates the complexities of the logging environment and shows
the problems that modeling and inversion codes must resolve in order to generate
accurate solutions.

5.1 Relating resistivity logs to rock physics; Rt and


Archie’s equation

The primary purpose of log interpretation is to answer two basic questions:

1. Where are potential hydrocarbon producing zones located?


2. How much hydrocarbons do these zones contain?
254 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Resistivity tools aid in answering these questions by providing a deep mea-


surement of true bulk formation resistivity, Rt . In addition to resistivity, the
two other important parameters that enter into log interpretation are poros-
ity (see Section 3.1) and permeability. The porosity of a rock is the fraction
of its total volume that is pore space; porosity determines the amount of
fluid that can be present in a rock matrix. The permeability of a rock is
its capacity for transmitting fluids; permeability controls the rate at which
hydrocarbons can be produced. Porosity is measured by density, neutron
or sonic logs. Permeability is measured by spontaneous potential or gamma
ray logs, or by formation testing.
The electrical characteristics of the materials in the drilling environment
are:

- water is conductive,
- hydrocarbons are insulators,
- rocks are non-conductive for practical purposes (shales containing bound
water can complicate interpretation).

Thus all electrical conduction takes place via the water contained in the pore
space. The fundamental task in log interpretation is to compare the mea-
sured electrical resistivity of a formation with the resistivity that would exist
if all the pore space contained water. If the measured resistivity is higher
than the water-filled resistivity, then hydrocarbons are present. Greater
departures from the water-filled resistivity indicate larger amounts of hydro-
carbons. The resistivity of the water in the pore space can vary widely (see
Section 3.1), and this must be taken into account in the evaluation.
The quantitative relationship between a rock matrix and water saturation
is expressed by Archie’s equation [39], which in its simplest form is

C Rw
Sw = , (5.1)
φ Rt

where Sw is water saturation (in percent), φ is porosity (in percent), Rw


is the water resistivity, and Rt is the deep formation resistivity supplied by
induction or laterolog tools. C is a cementation factor, which is equals 1 for
carbonates and 0.9 for sands. Archie’s equation was derived from laboratory
experiments in the early 1940’s. It is the basic equation of log interpretation.
5.1. RELATING RESISTIVITY LOGS TO ROCK PHYSICS 255

Figure 5.1: Idealized log demonstrating the joint interpretation of perme-


ability, resistivity and porosity logs.

There are several steps involved in applying Archie’s equation to deter-


mine the amount of hydrocarbons present in a rock’s pore space. The proce-
dure is outlined below using an idealize log example taken from Dewan [86],
which is shown in Figure 5.1. The evaluation begins with the selection of
potential hydrocarbon bearing zones and ends by calculating the number of
barrels of oil in the reservoir.
256 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

The steps in the evaluation are:

1. Determine where potential productive zones are located

a. Locate permeable zones – Track 1


– Right base-line: impermeable shales – non-producible
– Left base-line: permeable sands – contain oil or water
– Possible hydrocarbon zones: A, B, C, D
b. Locate high resistivity zones – Track 2
– Right base-line: high resistivity – hydrocarbons or tight rocks
– Left base-line: low resistivity – water sands
– Possible hydrocarbon zones: A, C
c. Locate high porosity zones – Track 3
– Right base-line: low porosity – non-producible
– Left base-line: high porosity – producible
– Possible hydrocarbon zones: A

2. Determine how much oil or gas is present using Archie’s equation

a. Obtain Rw from a nearby water-bearing interval (D) with Sw = 1



.9 Rw
1= , gives Rw = .045 ohm−m (5.2)
.35 .3

b. Use this Rw in the zone of interest (A) to solve for Sw


+
.9 .045
Sw = , gives Sw = 32% (5.3)
.3 4

c. Calculate the amount of pore space in zone A containing hydro-


carbons

1 − Sw = 68% , hydrocarbons in porespace (5.4)

d. Calculate the volume of hydrocarbons in zone A

φ × .68 = 20.4% , bulk volume hydrocarbons (5.5)


5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 257

e. Multiply the above bulk volume of hydrocarbons by the height of


the reservoir times its areal extent to obtain the total barrels of
oil in situ
.204 × 10 (feet) × 40 (acres) × 7758 (conversion factor)
= 633, 000 barrels (5.6)
Assuming a recovery rate of 20% and a price of $25 per barrel, the potential
revenue from this reservoir would be $3.16 million.
It should be noted that although zone A in Figure 5.1 was not the zone
with the highest resistivity, it proved to be the only hydrocarbon bearing
zone by virtue of the additional information provided by the porosity and
permeability logs. Even though this is a simple example, it serves to il-
lustrate the fundamentals of log interpretation and the role that resistivity
measurements play in evaluating the amount of hydrocarbons in a reservoir.

5.2 Early 1D plus 1D “inversion” efforts

Many hydrocarbon bearing reservoirs consist of resistive beds surrounded


by conductive shoulders, such as zone A in Figure 5.1. Modeled logs in the
benchmark formation in Chapter 3 (for example, Figures 3.25, 3.26, 3.32,
3.74, 3.79, 3.88) show that induction and laterolog tools have a great deal
of difficulty reading the true formation resistivity, Rt , in this configuration.
Departures from Rt become even greater than those shown when beds are
very thin, when deep invasion is present, or when borehole effect is severe.
For induction tools, skin effect creates an additional problem by causing a
significant nonlinear decrease in the measured signal, especially in conductive
formations.
Over the years, a group of standard correction charts and algorithms was
developed to compensate for these parasitic effects. Induction skin effect
losses were corrected by a boosting algorithm. Borehole effect was corrected
using borehole correction charts. Shoulder effect was corrected either by
deconvolution (for induction) or by using shoulder correction charts (both
laterolog and induction). Rt in invaded formations was determined using
either tornado charts (for induction) or butterfly charts (for laterologs); these
charts were named after the natural phenomena that their shapes resembled.
Corrections were applied sequentially in an empirically defined order.
The objective was to simulate 2D axisymmetric inversion by combining 1D
258 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

radial corrections for borehole and invasion effects with 1D vertical correc-
tions for adjacent bed effects. Since most wells drilled prior to 1985 were
vertical and most beds of interest were more than five feet thick, serial cor-
rections provided reasonable results. However, this methodology is not valid
in principle because the fields generated by resistivity tools interact with all
of the media they penetrate in a complex nonlinear fashion. These early cor-
rection efforts were only a stopgap means of estimating Rt until advances in
computer technology made it possible to use interactive 2D and 3D modeling
and inversion in log interpretation.
Some of the most commonly used correction charts and algorithms are
summarized in the remainder of Section 5.2. Although the automatic soft-
ware processing used in conjunction with modern array tools has rendered
many of these procedures obsolete, they are described here because the on-
going analysis of existing reservoirs often includes the interpretation of early
logs.

5.2.1 Deconvolution and boosting

The processes of deconvolution and boosting were only used to correct in-
duction logs. They were replaced in the 1980’s by more modern signal pro-
cessing techniques, such as Phasor processing (described in Section 3.2.4).
The purpose of boosting was to amplify raw induction signals in order to
compensate for losses due to skin effect. The deconvolution technique that
was used on early induction logs was a simple weighted average designed
to decrease shoulder effect; it did not involve inverse convolution filtering
because of computational constraints that existed in the 1960’s and 1970’s.

Deconvolution

The commercial deconvolution of induction logs was introduced by Henri


Doll [98]. Doll’s algorithm attempted to reduce the effect of adjacent shoul-
der beds by giving greater proportional weight to the signal measured at the
sonde center than to signals measured above and below this point. Deconvo-
lution was only applied to 6FF40 and ID logs (IM was not deconvolved). The
correction was performed using a “panel” analog computer on the logging
truck. In practice, the process involved storing the raw conductivity signals
and applying a three-station windowing filter at ±78 inches above and be-
low a given logging station. The algorithm used to compute the deconvolved
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 259

signal was

σD (z) = −w1 σR (z − 78 inches) + w0 σR (z) − w1 σR (z + 78 inches), (5.7)

where σR is the raw induction R-signal and σD (z) is the resulting decon-
volved log at depth z. The distances ±78 inches were selected from in-
spection of the 6FF40 vertical geometrical factor, and the weights, w, were
determined empirically for several values of shoulder bed resistivity (SBR).
Weights for various values of SBR are given in Table 5.1. The difficulty of
predicting SBR values a priori led to using SBR = 1 as the standard setting
in later years. The SBR value is normally recorded on the log heading.

SBR w0 w1
0.25 1.00 1.00
0.5 1.06 1.03
1.0 1.10 0.05
2.0 1.16 0.08
4.0 1.20 0.10

Table 5.1: Deconvolution coefficients.

Skin effect boosting

Boosting is the amplification of the raw induction signal to compensate for


the nonlinear losses due to skin effect. Boosting was applied to IM, ID and
6FF40 logs. (5FF27 and 5FF40 were designed to have low skin effect and
their logs were not boosted.) ID and 6FF40 logs were deconvolved before
they were boosted. The boosting algorithm was an approximate fit based on
the tool response in a homogeneous medium [194]. The fitting function [215]
was
log σB = log(η · σR ) + η · β · σR , (5.8)
where where σR is the induction R-signal and σB is the boosted signal. All
conductivities are expressed in mS/m. η was chosen such that the formula
gives the correct σB at 500 mS/m, and β is a tool constant. The coefficients
η and β for 6FF40, ID and IM are given in Table 5.2.
A logarithmic formula was used instead of a simple polynomial fit be-
cause it was easier to implement on the analog panel computer. In addition,
260 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Figure 5.2: Deconvolved and boosted modeled ID logs in low and high re-
sistivity formations.

experience demonstrated that Equation (5.8) gave more stable answers over
the entire logging range of conductivities than a polynomial fit.

Sonde η β
6FF40 1.0739 0.00135
ID 1.0899 0.00135
IM 1.0494 0.00030

Table 5.2: Boosting coefficients.

Figure 5.2 [17] illustrates the use of deconvolution and boosting to im-
prove Rt predicted by ID for two computed logs with bed resistivities dif-
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 261

Figure 5.3: A modeled ID log with resistivities between those in Figure 5.2,
deconvolved at three different SBR settings.

fering by a factor of 100. A comparison of the two sets of logs shows that
boosting does a reasonably good job correcting the low resistivity log for
skin effect, while the high resistivity log requires little skin effect correction.
The high resistivity logs are more subject to shoulder effect before processing
than the low resistivity logs, especially in the series of thin beds between 65
and 82 feet and in the thick resistive bed between 82 and 100 feet. Deconvo-
lution with SBR = 1 slightly improves the low resistivity log, while it does
not adequately correct the high resistivity log.
Figure 5.3 [17] shows the effect of deconvolving the same modeled ID log
at three different SBR settings. (The logs were all boosted after deconvo-
lution.) In the resistive bed between 82 and 100 feet, increasing the SBR
262 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

setting pushes the ID apparent resistivity near the boundaries closer to Rt ,


but the correction is not adequate in the center of the bed. In the series of
thin beds between 65 and 82 feet, increasing SBR only marginally improves
the ID apparent resistivity.

Removal of deconvolution and boosting

In order to upgrade the analysis of existing reservoirs, it is often neces-


sary to reprocess old induction logs with more modern signal processing or
inversion methods . This task is not straightforward, because service com-
panies archived only deconvolved and boosted logs, and most reprocessing
requires raw data. Therefore it is desirable to have a means of working back-
wards from the processed logs to obtain the raw signals. Regenerating raw
induction signals from commercially deconvolved and boosted logs can be
accomplished using the following two-step process [17]. The operations must
be performed in the order indicated, and the data must be in conductivity
units of mS/m.
First, boosting is removed using the expression


8
ln(σDB ) = ci [ln(σB )]i , (5.9)
i=0

where σB is the boosted conductivity signal and σDB is the deboosted con-
ductivity. The coefficients ci are given in Table 5.3 for 6FF40, ID and IM.
These coefficients were derived by applying Equation (5.8) to boost the R-
signal conductivities obtained by modeling tool response in a series of ho-
mogeneous media, and then using a least squares technique to fit ln(σB ) to
ln(σR ).
For IM, the removal of boosting completely deprocesses the log. For ID
and 6FF40, deconvolution is removed using the expression

σDD = + h0 σDB (z)


+ h1 [σDB (z − 78 inches) + σDB (z + 78 inches)]
+ h2 [σDB (z − 156 inches) + σDB (z + 156 inches)]
+ h3 [σDB (z − 234 inches) + σDB (z + 234 inches)] , (5.10)

where σDB is the deboosted conductivity from Equation (5.9) and σDD is
the de-deconvolved conductivity signal. Equation (5.10) was obtained by
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 263

6FF40 ID IM
c0 −.718610E−1 −.863658E−1 −.481587E−1
c1 +.100386E+1 +.100033E+1 +.998368E+0
c2 −.936772E−2 −.162975E−3 +.367211E−2
c3 −.803369E−2 −.288303E−3 −.356167E−2
c4 −.355743E−2 −.145077E−5 +.172759E−2
c5 +.832554E−3 +.199875E−4 −.468673E−3
c6 −.104627E−3 −.233440E−5 +.709200E−4
c7 −.601166E−5 −.686051E−6 −.563695E−5
c8 +.117641E−6 +.606668E−7 +.175232E−6

Table 5.3: Coefficients for removal of boosting.

rewriting Equation (5.7) so that it defines the raw signal at a central logging
station in terms of the known weights and deconvolved log, plus the unknown
raw signal on either side of the central station. Continuously replacing the
unknown raw signals with the expressions defining them in terms of the
deconvolved log at ±78 inch intervals gives a formula for calculating the raw
signal that is accurate to 0.1 mS/m when the substitution is truncated at
seven stations. The values of hi are derived from the deconvolution weights
and are given in Table 5.4 for all SBR settings (there are no values for
SBR = 0.25 since this represents no deconvolution). Identical results can be
obtained by least squares fitting.

SBR = 0.5 SBR = 1 SBR = 2 SBR = 4


h0 0.944907541 0.912847483 0.870269384 0.844907407
h1 0.026764053 0.041578444 0.060301351 0.070891204
h2 0.000755657 0.001878287 0.004100209 0.005787037
h3 0.000021386 0.000085377 0.000282773 0.000482253

Table 5.4: Coefficients for removal of deconvolution.

5.2.2 Correction chartbooks and departure curves

Correction charts are used in conjunction with logs to remove unwanted ef-
fects caused by media that are adjacent to beds of interest. Charts are gen-
erated by first running forward models to compute tool apparent resistivity
264 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

readings in given formations. The modeled formation parameters are then


plotted as a function of specific apparent resistivity values. Using a chart
“inverts” the procedure: apparent resistivity values from logs are entered
into the chart and formation parameters are retrieved. The computed tool
response curves are sometimes called departure curves because they reflect
the departure of the apparent resistivity from the true formation resistivity.
Correction charts are published by service companies in log interpreta-
tion chartbooks. Early charts were used manually. Because this was a te-
dious and time consuming task, chartbook corrections were only performed
on potential hydrocarbon bearing zones. Since the 1970’s, computer pro-
grams that automatically perform chartbook corrections have been available
at most service company computer centers, or from third-party software ven-
dors. These programs are constructed from digitized plots of the charts. The
most common methods used to implement the software corrections are cubic
spline interpolation and least squares fitting [226].
The standard order [215] for applying chartbook corrections for resistivity
tools is to first correct for borehole effect, then for shoulder effect, and finally
for invasion effect. Examples of typical correction charts for both induction
and laterolog tools are described below.

Borehole and eccentricity effect corrections

Borehole effect and sonde eccentricity effect are corrected simultaneously.


Prior to the 1980’s, induction borehole correction was performed using bore-
hole geometrical factor curves. Figure 5.4 [230] shows a borehole correction
chart for the Dual Induction tool. Curves are plotted for several standoff
values (eccentricity is measured in reference to the center of the borehole,
while standoff is measured in reference to the borehole wall). This data was
obtained experimentally in the 1960’s, and has since been reproduced by
computer modeling [187]. The chart indicates that with the exception of IM
run at a standoff of 0.0 inches, borehole correction is only necessary for large
boreholes or very conductive muds.
The dashed lines illustrate the use of the chart for a 6FF40 sonde with
1.5 inch stand-off in a 14.6 inch diameter borehole and a mud resistivity,
Rm , of 0.35 ohm-m. The value of the borehole geometrical factor, in this
case 0.0019, is multiplied by the mud resistivity using the Rm scale to give
a hole signal of 5.5 mS/m. This value is subtracted from the apparent
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 265

10 -10

9 -5

8 0
5

2
7 5
1

6 10
.5

5 15
.3

4 .2 20

3 25
.1

2 30
.05

1 35

0 40

-1 45
4 6 8 10 12 14 16 18 20

Figure 5.4: Borehole correction chart for the Dual Induction tool using bore-
hole geometrical factors.

conductivity signal to give the borehole corrected conductivity signal, which


is then converted to apparent resistivity.
To avoid correcting logs manually, it was a common field practice to
subtract the expected borehole signal (based on mud conductivity and bit
size) from the log reading by adjusting the sonde error correction (sonde
error is described at the end of Section 3.2.4). This adjustment is normally
noted on the log heading.
Geometrical factor corrections are sufficiently accurate in average size
boreholes with moderately conductive muds because there is little skin ef-
fect in these environments. However, the borehole correction procedures
for the Phasor induction tool [158] and the AIT Array Induction Tool [129]
incorporate the full model of eccentered tool response with skin effect be-
266 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Figure 5.5: Borehole correction charts for LLS, centered.

cause these tools can be run in boreholes as large as 30 inches in diameter.


In addition shallowest AIT arrays are strongly affected by conductive mud.
These modern tools are borehole corrected by the logging truck computer
using a least squares fitting algorithm based on the modeled tool response.
The most accurate corrections are obtained when wellbore resistivity and
hole size are measured with auxiliary devices.
For laterolog tools, borehole correction is performed using charts gener-
ated from computed tool response for specific borehole sizes and mud and
formation resistivities. A borehole correction chart for LLS is shown in Fig-
ure 5.5 [230]. To use the chart, the ratio of the apparent resistivity reading
to the mud resistivity is entered on the x-axis. The borehole size is entered
proceeding upward from this point. The appropriate correction factor is
then obtained by projection on the y-axis. The apparent resistivity reading
is multiplied by the correction factor to yield the borehole corrected resis-
tivity. Borehole correction charts are available for all laterolog tools run in
both centered and eccentered modes. Modern laterolog tools use software
algorithms based on digitized versions of these charts to perform borehole
corrections.

Shoulder correction charts

Although the deconvolution of ID induction response reduces the effect of


adjacent beds to some extent, it does not remove shoulder effect completely,
especially when beds are very resistive (see Figure 5.3). IM response is also
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 267

Figure 5.6: ID shoulder correction Figure 5.7: IM shoulder correction


chart for Rs = 1 ohm-m (SBR = 1). chart for Rs = 1 ohm-m.

subject to shoulder effect, mainly in resistive thin beds (see Figure 3.25).
Shoulder correction charts are used to remove the effect of adjacent beds
from Dual Induction logs. These charts apply only to center-bed apparent
resistivity readings (Ra ).
Figure 5.6 [230] shows a typical shoulder correction chart for ID. Fig-
ure 5.7 [230] shows the corresponding chart for IM. These two charts were
generated by the author of this thesis in 1975 [215]. Shoulder effect is a
function of the contrast between bed resistivities. For ID and IM, shoul-
der effect is also a function of formation conductivity level, since the skin
effect boosting algorithm does not adequately compensate for the conduc-
tivity in layered media because it is based on tool response in homogeneous
media. Therefore ID and IM require several different shoulder correction
charts which cover various ranges of shoulder bed resistivity (Rs ). Phasor
induction tools need only one chart each for IM and ID because the Pha-
268 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

500
200
100
50
20
10
5 RLLD/Rs
RLLD/Rs

500

2 10
2
1 .5
.5
.2 .1
.1
.05
.02.01
.005 .005

Figure 5.8: Shoulder correction chart for LLD.

sor algorithm accurately corrects for skin effect over a tool’s entire range of
operation (see Section 3.2.4). Modern array induction tools do not require
shoulder correction because their vertical resolution is one to two feet.
Shoulder correction charts such as Figures 5.6 and 5.7 are generated by
using a forward modeling code to compute center-bed Ra readings for a
series of known values of bed thickness and Rt (with Rs held constant for
each chart). Selected constant values of Ra are obtained by interpolation of
the modeled data, and the Ra curves are plotted on the shoulder correction
chart as a function of Rt and bed thickness.
To use the chart, the bed thickness derived from inflection points on a
log is entered at the bottom along the x-axis. The center-bed Ra reading
from the log is entered proceeding upward from this point, referring to the
scale at the right of the chart. Rt is then obtained by projection on the
y-axis.
The magnitude of the departures of the Ra curves from Rt is an indication
of the amount of shoulder effect present over the range of bed thicknesses.
The oscillation of the Ra curves at high Rt levels results from the amplifi-
cation of small excursions in the vertical response functions at high-contrast
interfaces (see Figures 3.17 and 3.22). ID oscillations are further accentuated
by deconvolution in beds of small to moderate size.
Dual Laterolog logs sometimes require shoulder correction, particularly
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 269

in high contrast thin beds. Figure 5.8 [230] shows a shoulder correction chart
for LLD. This chart was generated in the 1970’s using analog resistor network
data, and has since been reproduced by computer modeling. Laterolog tools
need only one chart because their response is only a function of resistivity
contrast and not a function of resistivity level. Laterolog shoulder correction
charts are used in the same manner as induction charts, except the ratio
RLLD /Rs is entered instead of Ra in order to express the laterolog reading
in the normalized units of the chart. The appropriate correction factor is
read on the y-axis.

Invasion effect—tornado and butterfly charts

Conventional Dual Induction-SFL (or LL8) tools provide resistivity logs with
three different depths of investigation. Tornado charts use these three sep-
arate resistivity readings to improve the estimation of Rt in thick invaded
beds. Tornado charts also provide a quantitative evaluation of the invasion
resistivity, Rxo , and the invasion diameter, di .
Figure 5.9 [230] shows an example of a tornado chart. The chart was
created by the author of this thesis in 1972 [215]. Published charts do
not take shoulder effect into account; they assume that invaded beds of
interest are sufficiently thick to have negligible shoulder effect, or that logs
have been corrected at least qualitatively using shoulder correction charts.
Service companies normally publish two or more different tornado charts for
various ranges of Rxo /Rm . More than one chart is needed because conductive
invasion significantly decreases the depth of investigation of the induction
tools. This decrease occurs because the skin effect boosting algorithm is
based on tool response in homogeneous media, which fails to adequately
compensate for the combined conductivity of the formation plus the invaded
zone.
Tornado charts are generated by using forward modeling codes to com-
pute ID, IM and SFL responses for selected values of Rm , Rxo , Rt and di .
Ratios of the modeled tool responses (RIM /RID and RSF L /RID ), along with
Rt /RID , are plotted on the tornado chart.
To use the chart, ratios of the tool responses are calculated from the
log apparent resistivity readings and entered on the x-axis and y-axis of
the chart. Once this coordinate is plotted, Rt /RID is read from the chart
(after visual interpolation), and Rt can then be calculated. Values for di
270 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Figure 5.9: Tornado chart for the Dual Induction-SFL tool with Rt /Rm =
100.

and Rxo /Rt can also be obtained visually, and Rxo can be calculated since
Rt is known. Computer programs are available which read log data and use
software algorithms to perform these operations (see Section 5.3).
Similar charts, called butterfly charts, are used to interpret laterolog tool
response in thick invaded beds. Butterfly charts cover both Rxo < Rt and
Rxo < Rt cases, which causes curves to fan out in two directions and thus
differentiates their shape from tornado charts. Figure 5.10 [230] shows a
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 271

Figure 5.10: Butterfly chart for the Dual Laterolog-Rxo tool.

butterfly chart for the Dual Laterolog-Rxo tool. The chart was created by
the author of this thesis in 1972.
Butterfly charts are generated from modeled laterolog response and used
in the same manner as tornado charts. Logs should first be corrected for
borehole effect, and for shoulder effect, if necessary. Rxo is obtained directly
from a microresistivity pad tool such as the microlaterolog or MicroSFL.
Ratios of the tool responses are entered on the x-axis and y-axis of the
chart. The value of di is read from the chart, and Rt is calculated either
272 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

from Rt /Rxo or Rt /RLLD .

Dip effect

Before horizontal drilling became common practice in the mid-1980’s, there


were no charts or algorithms for correcting resistivity logs in deviated wells
or dipping beds such as the corrections described above for vertical wells.
Estimates obtained from modeling the response of induction and laterolog
tools crossing a dipping interface showed that dip effect only became signif-
icant at angles greater than 30◦ . Since most wells drilled before 1985 were
vertical, and naturally occurring geologic dip is usually less than 30◦ , dip
was considered to be a second order effect and was largely ignored.
The introduction of horizontal drilling made it impossible to neglect dip
effect. In horizontal and highly deviated wells, apparent dip is much greater
than naturally occurring geologic dip. Complex new interpretation problems
appear at high dip angles. Induction tools that were designed to have opti-
mum response in vertical wells can be affected by beds as far away as fifteen
feet from the borehole in highly deviated and horizontal wells [35]. 2-MHz
tools were introduced in the late 1980’s, and their use in horizontal wells
to steer drilling toward a target bed requires modeling the precise effect of
dip on tool response [20]. Anisotropy which is invisible to resistivity tools in
vertical wells becomes a significant effect in horizontal wells (see Equation
(1.6)).
Horizontal wells are more costly to drill than vertical wells. The accurate
evaluation of productive zones is an important factor in keeping expenses
down, and it soon became apparent that quantifying the effect of dip was an
economic necessity. The University of Houston was one of the first institu-
tions to put significant effort into dip correction. They developed a code for
modeling induction response in dipping beds [143] and published shoulder
correction charts [144] for various dip angles. Several of these charts are
shown in Figure 5.11.
Charts such as these are only valid when the shoulder-bed resistivity is
the same above and below a bed of interest. In addition, each chart can be
used for only a limited range of dip angles bracketing the angle for which it
was computed. Because of these limitations, Schlumberger favored iterative
forward modeling and inversion over correction charts for quantifying dip
effect in highly deviated and horizontal wells.
5.2. EARLY 1D PLUS 1D “INVERSION” EFFORTS 273

Apparent Bed Thickness (ft) Apparent Bed Thickness (ft)

Ra
o (ohm-m)
0o Ra 45
(ohm-m)
Rt (ohm-m)

Rt (ohm-m)

ILd ILd
Rs = 1 ohm-m Rs = 1 ohm-m

True Bed Thickness (ft) True Bed Thickness (ft)

Apparent Bed Thickness (ft) Apparent Bed Thickness (ft)

Ra
o
30o Ra 60 (ohm-m)
(ohm-m)
Rt (ohm-m)
Rt (ohm-m)

ILd ILd
Rs = 1 ohm-m Rs = 1 ohm-m

True Bed Thickness (ft) True Bed Thickness (ft)

Figure 5.11: Shoulder correction charts for the ID induction tool with Rs = 1
ohm-m, and dip angles of 0◦ , 30◦ , 45◦ and 60◦ .
274 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Shortly after this time, a code for computing induction and 2-MHz tool
response in dipping beds [32] was made available at Schlumberger regional
computer centers [14, 19]. This code was used to improve Rt estimation in
highly deviated wells by means of iterative forward modeling [19, 117] and to
steer drilling in horizontal wells [169]. The improved Rt estimates obtained
by modeling (for example, see Figure 1.6) led to much more accurate reserve
predictions, with the improvements subsequently confirmed by production
histories. Using this same code, a dip correction algorithm based on filtering
was developed for both Phasor and array induction tools [45]. The algorithm
is accurate up to 60◦ dip. More recently, maximum entropy inversion [49]
was applied to array induction logs to estimate Rt , Rxo and di in invaded thin
beds at high relative dip angles. The accuracy of the inversion was validated
using 3D modeling [18]. Thanks to today’s array tools and computationally
efficient 2D and 3D modeling and inversion software, chartbooks can finally
“rest in peace.”

5.3 A 2D axisymmetric iterative forward modeling case


study

When fast 2D axisymmetric forward modeling codes first became available


in the mid-1980’s, they immediately demonstrated how inaccurate chart-
book corrections were for estimating Rt in invaded thin beds. Fast for-
ward modeling codes also allowed the development of efficient inversion
software, and the 2D inversion of resistivity logs was attempted shortly af-
ter this time. Among the inversion methods used were: constrained least
squares (Levenberg-Marquardt ridge regression) [170, 268, 267], maximum
entropy [115], parametric [136] and distorted Born [72, 172]. These inversion
efforts often produced nonunique or obviously incorrect solutions because of
the inadequate information content of the existing measurements. Recently,
the contrast source inversion method [261, 2, 262, 4, 3] has shown promise in
the inversion of cross-well problems. However, this method has not yet been
applied to single-well resistivity measurements. During the 1990’s, inversion
was temporarily abandoned in favor of iterative forward modeling. Modeling
permitted a higher level of user intervention, making it easier to constrain
solutions within a petrophysically meaningful solution space. It also allowed
local geological knowledge and information from non-resistivity logs to be
incorporated in formation models. The inversion codes of that time did not
5.3. A 2D ITERATIVE FORWARD MODELING CASE STUDY 275

have these capabilities.


The mid-1980’s also saw the addition of the dimension of time to log
interpretation. This occurred when 2-MHZ logging while drilling (LWD) re-
sistivity tools with multiple depths of investigation were introduced [210, 74].
Because LWD tools are located immediately above the drill bit, they make
measurements typically within one to five minutes [11] of bit penetration.
Wireline measurements are generally made two to ten days after drilling.
Thus the evolution of invasion over time can be monitored by the joint in-
terpretation of LWD and wireline logs.
In addition to providing a more complete reservoir analysis, joint LWD–
wireline interpretation also decreases ambiguity in the formation model.
Lithology and porosity do not change with time. Therefore bed bound-
aries, Rt and Rxo can be constrained to be the same for both LWD and
wireline modeling. The only parameter that varies with time is the invasion
diameter, di . (If annulus invasion exists, there is also a variable annulus
diameter and an annulus resistivity which may be constant or variable.)
Iterative forward modeling was routinely applied to time-lapse invasion
interpretation. Its use can best be illustrated by means of a case study. Fig-
ure 5.12 shows the LWD and wireline field logs to be interpreted. This exam-
ple is taken from an analysis of a well drilled in onshore southern Louisiana
previously published by this author [11]. The final modeled reproduction of
the field log is shown in Figure 5.13. The solutions obtained for Rxo , Rt and
di at both LWD and wireline times are also shown in Figure 5.13.
The 2-MHz Compensated Dual Resistivity (CDR) field log is in the center
track of Figure 5.12. Even though these logs were recorded within minutes
of drilling, there is significant separation between the shallow phase shift
resistivity (RPS) and the deeper attenuation resistivity (RAD) curves in
three regions. The Phasor Induction-SFL logs in the right track exhibit
larger separations between the deep and medium induction curves (IDPH
and IMPH) and the shallow SFL curve in these same regions. Because this
is a vertical well and the beds are over ten feet thick, it is safe to assume that
curve separation is caused primarily by invasion and not by anisotropy or
shoulder effect (all logs were borehole corrected). The attenuation resistivity
reads the same value as the deep induction log in most cases, but occasionally
reads a lower value.
The left track of Figure 5.12 displays the SP log and wireline hole di-
ameter, along with two calculated invasion diameters. The wireline di was
276 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Figure 5.12: LWD and wireline field log from a well in southern Louisiana.

Figure 5.13: Modeled reproduction of the field log in Figure 5.12.


5.3. A 2D ITERATIVE FORWARD MODELING CASE STUDY 277

obtained from a Phasor tornado chart algorithm. The LWD di was estimated
using tabulated values of the ratio RPS/RAD versus di for various Rxo and
Rt values [12]. The largest calculated wireline invasion diameters correlate
well with the permeable sand beds indicated by the SP log.
The basic questions facing the log interpreter are:

1. What is Rt ?
2. Why does the separation between curves vary within the same bed for
logs run at the same time, and why is the separation different at LWD
and wireline time?
3. What can be learned about the formation properties from the curve
separation?

Iterative forward modeling was used to reproduce the LWD and wireline
logs in an attempt to answer these questions. Induction and 2-MHz tool
response was modeled with a fast 2D semi-analytic code [73]. SFL response
was modeled with a 2D finite element laterolog code [274, 119]. Differences
in vertical resolution and depth of investigation among the various tools are
not a problem because the modeling codes account for these effects exactly.
The initial model was constructed as follows:

- Inflection points of the CDR phase shift resistivity log were used to
select bed boundaries, since this is the log with the highest vertical
resolution.
- Rt and Rxo were estimated from the wireline logs.
- A single-front invasion profile was assumed at both LWD and wireline
time.
- LWD di was estimated from the calculated LWD invasion diameters.
- Wireline di was estimated from the calculated tornado chart invasion
diameters.

For all the iterative modeling runs, Rt , Rxo and the bed boundaries were
constrained to have the same value at LWD and wireline time; only di was
allowed to differ.
The final modeling results shown in Figure 5.13 were obtained after four
iterations. Agreement with the field logs is excellent, confirming that a uni-
form lithology with variations in fluid penetration is a reasonable assumption
278 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

for time-lapse modeling. It should generally take only three or four itera-
tions for a moderately skilled log interpreter to obtain a good match between
modeled logs and field logs. If more iterations are necessary, there is usually
an error in one or more fundamental assumptions about the formation model
(such as neglecting dip effect, or not accounting for annulus invasion).
The modeling results answer question (1) above concerning the value
of Rt . The answer to question (2) regarding curve separation is related to
the vertical variations of the invasion diameters, and needs a more detailed
explanation. In the three invaded sand beds, the LWD di decreases from the
bottom to the top of the sand, while the wireline di increases. Sidewall core
analysis of these beds shows that the sand is fining upward. Since early-time
invasion volume decreases with decreasing grain size [8], the LWD di will be
shallower at the top of the beds where the sands are compacted and less
permeable.
Late-time invasion is affected more by vertical gravity segregation or
buoyancy [96, 105]. This well was drilled with fresh mud. Since fresh mud
filtrate is less dense than the formation salt water, it rises to the top of the
sand beds over time. This results in an invasion diameter at wireline time
that increases from the bottom to the top of a bed.
Regarding question (3), vertical permeability plays an important role in
determining the invasion diameter at the time of wireline logging [105, 7].
Wireline di is only moderate in zones of high vertical permeability because
the rate at which filtrate rises exceeds the rate at which it enters a zone.
The opposite occurs in zones of low vertical permeability, where the filtrate
cannot escape rapidly. When a zone of low vertical permeability occurs above
a zone of high vertical permeability the two effects are magnified, such as
at the top of the bed between 3250 and 3265 feet. In addition to gaining
fluid from the wellbore, this zone is fed by filtrate migrating upward from
the lower part of the bed. At the top of a high permeability zone bounded
by a permeabilty barrier, the filtrate spreads out into a thin, deep invasion
front.
Although the 2D modeling used in this case study is more accurate than
chartbook corrections, it cannot overcome the fundamental problem that
there are many more unknowns than measurements. The ultimate test of
accuracy for any given iterative modeling (or inversion) solution is to check
for consistency between the mathematical results and all available petro-
physical information. When this is done, modeling becomes a useful tool for
5.4. A LEAST SQUARES INVERSION EXAMPLE IN THIN BEDS 279

understanding petrophysics and reservoir dynamics, in addition to providing


Rt .

5.4 A least squares inversion example in isotropic thin


beds

The essential task of log interpretation is the solution of the inverse problem,
i.e., the determination of formation parameters from logging data. The first
part of Chapter 5 reviewed empirical inversion efforts: Section 5.2 described
early chartbook corrections, while Section 5.3 described the more recent
use of iterative forward modeling. Iterative forward modeling provides an
accurate solution because it simulates the exact tool response in a given
formation. However, a high level of user interaction is required, which is a
major disadvantage. A more desirable solution would be a stable and accu-
rate inversion algorithm which allows the initial incorporation of geological
information and requires no interaction after the software is launched.
In the early 1980’s, the least squares method was one of the first software
algorithms to be successfully implemented [170] for the automated inversion
of resistivity logs. Least squares inversion had been used on seismic data
since the late 1960’s because of its mathematical robustness in the presence
of insufficient or inaccurate information [171]. The least squares method
solves for the formation parameters by minimizing the sum of the squares
of errors between a forward model and the logging data. Most least squares
algorithms are iterative in nature, starting with an initial guess for a set of
formation parameters and then generating a sequence of sets of formation
parameters which decrease the sum of the squares of the errors. Eventually
the solution converges to a final set of parameters for which the sum of the
squares of errors is a global minimum.
In order to illustrate the inversion of resistivity logging data, the least
squares method is applied to invert an induction log in 1D isotropic layered
media. This is admittedly a simple example. However, it serves to introduce
the notation of inversion and raises the practical problems associated with
the inversion of resistivity logs. The computer code for performing the inver-
sion was implemented using the strategy described by Lines and Treitel [171].
The mathematical algorithm is described first, followed by an outline of the
steps involved in adapting the algorithm to invert the induction log example.
280 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

The set of N logging data points is defined by the vector r. The corre-
sponding set of N model responses is defined by the vector f . The model is
a function of M parameters which are elements of the vector p.
Let pj 0 be the jth element of the initial estimate of p and let f 0 be the
initial model response. If the model response is a piecewise linear function
of the parameters, a perturbation of the model response about p0 can be
represented by the first order Taylor expansion, which is
&
0

M
∂ff &&
f =f + & (pj − pj 0 ) , (5.11)
∂ pj & p=p0
j=1

or in matrix notation
f =f 0 +Zδ, (5.12)
where Z is the N by M Jacobian matrix of partial derivatives with elements
∂fi
Zij = , (5.13)
∂ pj
and δ = p − p0 is the parameter change vector.
The choice of perturbations in p will be made so as to minimize the
sum of squares of the errors between the model response and the data. Let
e represent the error vector expressing the difference between the model
response f and the logging data r , that is
r −f = e. (5.14)
Combining Equation (5.12) and Equation (5.14) by equating f yields
r −f 0 = Zδ +e. (5.15)
The vector r −ff 0 , which is the difference between the initial model response
and the observed data, is called the discrepancy vector g , so that
g =r −f 0. (5.16)
In addition,
e =g −Zδ. (5.17)

In the simplest least squares approach, the cumulative error S = eT e


is minimized with respect to the parameter change vector δ . From Equa-
tion (5.17)
S = eT e = (gg − Z δ )T (gg − Z δ ) . (5.18)
5.4. A LEAST SQUARES INVERSION EXAMPLE IN THIN BEDS 281

Minimization of S with respect to δ requires that


S
∂S
= 0. (5.19)
∂δδ
Substituting Equation (5.18) into Equation (5.19) gives
∂ T T
(δδ Z Z δ − g T Z δ − δ T Z T g + g T g ) = 0 . (5.20)
∂δδ
Carrying out differentiation with respect to δ gives the so-called “normal
equation”
ZT Z δ = ZT g , (5.21)
whose solution for the parameter change vector δ is

Z T Z )−1 Z T g .
δ = (Z (5.22)

The parameter vector is updated, such that

p = p0 + δ , (5.23)

and the procedure is repeated until δ reaches a specified small value.


Equation (5.22) is known as the Gauss-Newton solution, which also has
important applications in inverse theory and digital filtering. The term “nor-
mal equation” arises from the property that the least squares error vector e
is perpendicular to the column vectors of the Jacobian matrix Z .
If the matrix Z Z T is nearly singular, δ tends to grow without bound. To
remedy this problem, a constraining condition is introduced so that δ is cho-
sen to minimize a cost function. This approach is known as the Levenberg-
Marquardt method of damped least squares.
One of the main problems that has historically prevented the widespread
commercial use of inversion in log interpretation is the difficulty of validating
the accuracy of results that are generated from field logs. Since formation
resistivities are not known a priori, inversion results are often considered to
be of questionable accuracy when they disagree with the interpretation of an
experienced log analyst. As a result, inversion methods are usually validated
by testing them on computed logs which are generated from know resistivity
values and bed boundary locations. In order to better approximate a field
log, noise is introduced on the computed log. The log is then inverted and the
accuracy of the inversion is checked by comparing results with the formation
parameters that generated the simulated log.
282 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Figure 5.14: Raw ID simulated log Figure 5.15: Results of least squares
to be inverted. inversion.

Figure 5.14 shows the simulated log used to test the least squares algo-
rithm: an ID log computed in the benchmark formation without invasion.
The log was calculated with a layered medium forward modeling code [32].
Figure 5.15 shows the results obtained from the inversion of this log. The
bed boundary locations used in this inversion were assumed to be the same
as those of the computed log (sensitivity to errors in bed boundary location
will be shown in Figure 5.17). Only center-bed readings were inverted in
order to to generate square matrices which are easier to work with. Using
data as far as possible from bed boundaries also avoids potential problems
caused by inaccurate bed boundary selection. The log apparent resistivity
readings were used as the first iterate.
The steps involved in applying the least squares algorithm to this log are
described below in sufficient detail for interested readers to construct their
5.4. A LEAST SQUARES INVERSION EXAMPLE IN THIN BEDS 283

own inversion codes. The inversion is carried out in conductivity units, while
the logs are plotted in resistivity units to agree with convention.

1. Select the conductivity increment to be used to compute the partial


derivatives for the Jacobian: 0.1 provided satisfactory results.
2. Select the log points, r, to be used in the inversion: center-bed readings
were used.
3. Determine the bed boundary locations: bed boundaries were assumed
to be the same as those of the computed log.
4. Determine the initial guesses for the model parameters, p (σt ): center-
bed apparent conductivity readings were used.
5. Calculate g according to Equation (5.17): a layered medium forward
modeling code was used to generate f 0 , with the log values used for r .
6. Compute the Jacobian, Z , according to Equation (5.13): the layered
medium code was used to compute ∂fi , with ∂pj = 0.1 (using center-
bed readings makes the Jacobian strongly tri-diagonal).
7. Evaluate Equation (5.22) to obtain δ : matrix operation subroutines
were used; there were no problems with the matrix inversion.
8. Add δ to p: this generates the next iterate.
9. Go to step 5 above if δ is less than some small specified value, stop
when δ converges: 5 iterations were generally sufficient for obtaining
convergence.

Figure 5.16 shows results obtained by inverting the same log with a +1
mS/m error in the center-bed readings. This is a typical value of sonde error
drift encountered in the field. The largest errors in Rt occur in the three
resistive beds between 80 and 122 feet, where the +1 mS/m error is the
largest percentage (5%) of the formation conductivity. The results in these
beds are approximately 5% too low. The same log was also inverted with
a −1 mS/m error, giving Rt values (not shown) that are approximately 5%
too high in the resistive beds. The errors in Rt in the conductive beds and in
the moderately resistive shoulder beds are less than 1%, and not discernible
in Figure 5.16.
Figure 5.17 shows results obtained by inverting the log in Figure 5.14
with a systematic 6 inch error in the bed boundary location. This is typical
of an error caused by squeeze or anti-squeeze effects (see the description
284 CHAPTER 5. USING MODELING IN LOG INTERPRETATION

Figure 5.16: Least squares inversion Figure 5.17: Least squares inversion
results with a +1 mS/m error in the results with a 6-inch error in bed
log center-bed readings. boundary locations.

associated with Figure 3.84) when using a bed boundary detection algorithm
based on log inflection points. The agreement between the formation Rt and
Rt obtained from inversion is very good, except in the thinnest resistive
bed between 80 and 83 feet, where shoulder effect is greatest. Inversion
results are generally more accurate when data near bed boundaries is either
omitted or given less weight in the inversion. This avoids forcing agreement
with a 1D or 2D model near bed boundaries where logs are most sensitive
to perturbations caused by 3D effects, such as dip, unsymmetrical invasion
or faulting. In general, data near bed boundaries should receive careful
consideration not only for least squares inversion, but for other inversion
methods as well.
Chapter 6

Parametric inversion in anisotropic


layered earth

Summary: This final chapter describes the application of parametric inversion


to obtain the resistivities Rh and Rv in each medium of an anisotropic layered
earth. The inversion is tested on two different resistivity measurements: the exist-
ing 2-MHz Compensated Dual Resistivity tool and a theoretical triaxial induction
tool. The layered anisotropic medium modeling code used in the inversion is briefly
summarized, and the parametric inversion algorithm is described in detail. Inver-
sion results show that the 2-MHz tool only has enough sensitivity to invert for Rh
and Rv when independent information is imposed to constrain bed boundary loca-
tions and dip and strike angles. The triaxial induction tool has sufficient sensitivity
to accurately invert for not only Rh and Rv , but for bed boundary locations and
dip and strike angles as well. Additional work is planned to extend the inversion to
fully 3D geometries.

6.1 Introduction

The growth of horizontal drilling during the 1980’s revealed that anisotropy
had a significant effect on resistivity tool response. When logs from verti-
cal and highly deviated wells in the same formation were compared, it was
discovered that anisotropy caused apparent resistivity readings to increase
286 CHAPTER 6. PARAMETRIC INVERSION

z z

Receiver Rz MzR
y y

L x L
x
T
Transmitter Tz Mz
y y

x x

Figure 6.1: Coil arrangement for a two-coil induction sonde (left) and the
equivalent dipole model (right).

as a function of increasing deviation angle in shales and in laminated zones.


These differences in resistivity level made it difficult to identify marker beds,
which created major problems when steering a horizontal well toward a tar-
get pay zone.
The discovery of the magnitude of anisotropy effect also prompted a
reassessment of the interpretion of laminated zones in vertical wells. Hydro-
carbon pay zones in thinly laminated sand–shale sequences are often missed
by conventional resistivity tools. These tools measure horizontal resistivity
because their currents flow mainly parallel to the bedding plane. Since the
horizontal resistivity is dominated by the low resistivity shale laminae and
not by the high resistivity hydrocarbon-bearing sand laminae (see Equation
(1.2)), the presence of hydrocarbons is masked. A measurement of the verti-
cal resistivity which is dominated by the high resistivity hydrocarbon-bearing
sands (see Equation (1.3)) would thus be a better hydrocarbon indicator.
A multi-component triaxial induction tool can provide much more in-
formation for interpreting anisotropy in both vertical and highly deviated
wells. Such a tool incorporates three mutually orthogonal transmitter and
receiver coils, which enables one in principle to determine the horizontal and
vertical resistivity and the transverse anisotropic resistivity through inver-
sion of the measured voltages. Baker-Atlas has recently introduced a tool
which implements this concept [164, 163, 38].
6.1. INTRODUCTION 287

z z

Receivers Rz Ry MzR MyR


y y
MxR
Rx L L
x x
Ty T T
Transmitters Tz Mz My
y y
T
Mx
Tx
x x

Figure 6.2: Coil arrangement for a triaxial induction sonde (left) and the
equivalent dipole model (right).

The coil configuration for a conventional two-coil induction tool is shown


in Figure 6.1 (left). Sufficiently small coils can be replaced by point magnetic
dipoles (right). All conventional induction tool coils are vertical magnetic
dipoles. The coil configuration for a multicomponent (triaxial) induction
tool is shown in Figure 6.2.
For the modeling and inversion described in this chapter, these coil sys-
tems are located in a layered anisotropic medium as shown in Figure 6.3.
Each anisotropic layer has a transverse isotropy (TI) parallel to the bed
boundaries. This is a reasonable assumption based on normal depositional
processes (see Section 1.4). The bed boundary direction is fixed as parallel
to the x-y plane, with each bed boundary having a unique value in the z-
direction. To simulate a variety of realistic logging conditions, the tool axis
is allowed to vary between perpendicular and parallel to the bedding plane.
Coils are modeled as point magnetic dipoles. It has been shown both
theoretically [240] and experimentally [20] that a point dipole model is ac-
curate at distances greater than several coil radii. The actual size of field
tool coils ranges from a radius of one to three inches. Since transmitter-to-
receiver spacings are usually much greater than 10 inches, the dipole model
is sufficiently accurate.
In designing new tools, it is desirable for coils to approximate pure point
dipole antennas as closely as possible, because dipole response can be mod-
eled rapidy and is straightforward to invert and interpret. In order to ensure
288 CHAPTER 6. PARAMETRIC INVERSION

z
z’
σv
R σh
θ
σv
φ σh
y
σv
σh

σv
T σh
x
σv
σh

Figure 6.3: Layered anisotropic medium to be modeled and inverted. θ is


the dip angle, and φ is the strike angle. (See Figure 2.14 for a more detailed
representation of the Cartesian reference frame.)

that field tool antennas are pure dipoles, the actual coil dimensions are mod-
eled during the design process. In addition to the coil radius, the vertical
extent of coils and the mandrel materials and shape are taken into consider-
ation. After this validation process, a point dipole model without a mandrel
is then assumed for the purposes of interpretation and inversion.
The borehole is also not included in the configuration of Figure 6.3 be-
cause borehole effect for induction-type tools is only significant in a small
number of cases in conductive muds when Rt /Rm is greater than 200 [222,
74]. These simplifications allow the forward problem to be modeled rapidly
using a spectral integral formulation.
The objective of the inversion described in this chapter is to retrieve the
TI anisotropic formation parameters from measurements made by receiver
coils when moving a tool through the formation shown in Figure 6.3. The re-
mainder of this chapter gives an overview of the forward model and describes
the inversion procedure.
6.2. FORWARD MODELING 289

6.2 Forward modeling

The relevant basic equations from Chapter 2 are repeated here for the conve-
nience of the reader. In Cartesian coordinates, the TI anisotropic formation
parameters in each layer of Figure 6.3 are given by
 
σh 0 0
σ= 0 σh 0 , (6.1)
0 0 σv
 
h 0 0
 = 0  0 h 0 , (6.2)
0 0 v
 
µh 0 0
µ= 0 µh 0 . (6.3)
0 0 µv

The conductivity and dielectric permittivity are combined into the complex-
valued dielectric constant
 ∗ 
 h 0 0
∗ =  + iσ
σ /ω =  0 ∗h 0 . (6.4)
0 0 ∗v

Maxwell’s equations for a magnetic dipole source are given by

∇ × H + iω∗ · E = 0, (6.5)
∇ × E − iωµ
µ · H = −K K e. (6.6)

The forward modeling code is based on a Schlumberger program called


ANISBEDS [137]. This program computes the electromagnetic fields of a
point magnetic dipole radiator in a layered, dipping, TI anisotropic medium.
Each layer has an independent horizontal and vertical electric conductivity,
dielectric permittivity and magnetic permeability. The program computes
the complex, vector-valued magnetic field H and the electric field E at any
position with respect to an arbitrarily oriented magnetic dipole transmitter.
Thus the program can be used to model various induction-type measure-
ments in this formation geometry.
Without going into detail, the basic steps in the program are as follows.
First, the dyadic tensor for the TI anisotropic Green function is derived.
290 CHAPTER 6. PARAMETRIC INVERSION

Next, a spectral integral representation is derived in each TI anisotropic


layer. After applying boundary conditions at the interfaces and introducing
local reflection and transmission coefficients at each boundary, and subse-
quently introducing global reflection and transmission coefficients that are
recursively determined across the layer sequence, one obtains the Green func-
tion in a layered, TI anisotropic medium.
The transmitter is an arbitrarily oriented point magnetic dipole located
at r 0 . It has a dipole moment density M (rr). Therefore, in Equation (6.6)

K e = M δ(rr − r 0 ). (6.7)

Here the dipole moment is given in units of Vm, while the dipole moment of
a current loop is given in Am2 . To reconcile this disrepancy, the current loop
magnetic dipole moment is multiplied by the circular frequency iω. The two
conventions are related by
   
1 Vs
M loop
M point Vm = iωµM Am2 . (6.8)
s Am
The Green functions convert the magnetic dipole density in the correspond-
ing fields,

H (rr) = iω∗h drr G(rr, r ) · M (rr ) δ(rr − r), (6.9)

E (rr) = drr Γ(rr, r  ) · M (rr ) δ(rr − r ). (6.10)

The resulting formulas for all possible situations of interest become very
complicated. To illustrate how the magnetic field at the receiver position
depends on the TI anisotropic parameters, two results are shown for a mag-
netic dipole source located at the origin in a homogeneous medium and in
layered media.

Homogeneous medium

In a homogeneous medium, the magnetic field for the vertical magnetic


dipole M z is
∞ 
Mz kρ2 (h)
ikz |z| µ
H (rr) = dkρ (h)
e kρ J0 (kρ ρ) iz ∓ ikz(h) J1 (kρ ρ) iρ .
4πωµz kz µz
0
(6.11)
6.2. FORWARD MODELING 291

The ∓ sign in front of the eρ term depends on the position of the observation
point with respect to the source. The + sign applies if the receiver is below
the source (z < 0).
The magnetic field for the horizontal magnetic dipole in a homogeneous
medium is more complex because the full formation anisotropy becomes
active in this case. At the same time the azimuthal symmetry of the TI
ansiotropic medium simplifies the simultaneous field analysis of the M x and
M y components of the dipole moment. The magnetic field for the horizontal
magnetic dipole is
∞  (h)
ω eikz |z| µ
H (rr) = − dkρ kρ ±i M x cos φ + M y sin φ) J1 (kρ ρ) iz
kρ (M
4π k02 µz
0
(h) 
kz (h)
|z| 1
− eikz M x cos φ + M y sin φ) J0 (kρ ρ) −
(M J1 (kρ ρ) iρ
k02 kρ ρ
(h)
kz (h)
|z| 1
− eikz M y cos φ − M x sin φ)
(M J1 (kρ ρ) iφ
k02 kρ ρ
(e)
eikz |z| 1
− (e)
M x cos φ + M y sin φ)
(M J1 (kρ ρ) iρ
kz kρ ρ
(e)  
eikz |z| 1
− (e)
M y cos φ − M x sin φ) J0 (kρ ρ) −
(M J1 (kρ ρ) iφ . (6.12)
kz kρ ρ
The ± sign in front of the ez term describes the vertical direction of the
magnetic field depending on the position of the observation point with re-
spect to the source. The + sign applies if the receiver is above the source
(z > 0).
In Equation (6.11) and Equation (6.12)
 1
h 2 2
kz(e) = k02 − k , (6.13)
v ρ
 1
µh 2 2
kz(h) = k02 − k . (6.14)
µv ρ
The angle φ comes from the relation between Cartesian and cylindrical
coordinates as
ix = cos φ iρ − sin φ iφ ,
iy = sin φ iρ + cos φ iφ . (6.15)
292 CHAPTER 6. PARAMETRIC INVERSION

In a homogeneous medium, a closed-form solution can also be found by


the method of Moran and Gianzero [193]. In the program ANISMED, the
spectral integrals of Equation (6.11) and Equation (6.12) are compared to
this closed-form solution. The comparison provides an accuracy estimate for
the numerical integration.

Layered medium with source and receiver in the same bed


In a layered medium with the source and receiver in the same bed, a
vertical dipole M = M z iz δ 3 (rr) at the origin gives the magnetic field above
the source (zR ≥ zT ) as
∞ (h) (h)
ωm µm kρ2 1 + ΓDm e2ikzm dm
H (rr) = − Mz dkρ (h) (h) (h) (h)
4πkm µzm kzm 1 − ΓU m ΓDm e2ikzm Lm
0

1 µm  (h) (h) 
kρ J0 (kρ ρ) eikzm zR + ΓU m eikzm (2hm −zR ) iz
(h)
km µzm

 (h) (h) 
ΓU m eikzm (2hm −zR )
(h)
− ikzm
(h)
J1 (kρ ρ) eikzm zR
− iρ . (6.16)

A horizontal dipole M = (M M xix + M y iy )δ 3 (rr) at the origin gives the


magnetic field above the source (zR ≥ zT ) as
∞ (h) (h)
ωm 1 + ΓDm e2ikzm dm
H (rr) = − dkρ kρ (h) (h) (h)
4πkm 1 − ΓU m ΓDm e2ikzm Lm
0

1 µm  (h) (h) 
kρ J1 (kρ ρ) eikzm zR + ΓU m eikzm (2hm −zR )
(h)
km µzm
(−i M x cos φ − i M y sin φ) iz
  
1 (h) (h)
eikzm zR − ΓU m eikzm (2hm −zR )
(h)
(h)
+ i kzm J0 (kρ ρ) − J1 (kρ ρ)
kρ ρ
(−i M x cos φ − i M y sin φ) iρ
1  (h) (h) 
J1 (kρ ρ) eikzm zR − ΓU m eikzm (2hm −zR )
(h) (h)
+ i kzm
kρ ρ

M x sin φ − M y cos φ) iφ
(M

∞ (e) (e)
ωm 1 + ΓDm e2ikzm dm
− dkρ kρ (e) (e) (e)
4πkm 1 − ΓU m ΓDm e2ikzm Lm
0
6.3. 2-MHZ INVERSION IN LAYERED MEDIA 293


i  (e) (e) 
J1 (kρ ρ) eikzm zR + ΓU m eikzm (2hm −zR )
(e)
kρ ρ
(−i M x cos φ − i M y sin φ) iρ
  
1 (e) (e)
eikzm zR + ΓU m eikzm (2hm −zR )
(e)
+ J0 (kρ ρ) − J1 (kρ ρ)
kρ ρ

M x sin φ + M y sin φ) iφ . (6.17)
(−M

Similar expressions for all possible cases can be found. These expressions
are given in a report on the ANISBEDS code by Habashy and Lüling [137]. In
Equation (6.16) and Equation (6.17), Γ(e) and Γ(h) represent global reflection
factors of the TE (transverse electric) and TM (transverse magnetic) type
that are recursively determined across the layer sequence above and below
the layer under consideration. These reflection factors depend on all the TI
anisotropic parameters of the layer sequence.
Equation (6.11) through Equation (6.17) indicate the structure of the
relations between the measurements, i.e., the three components of the mag-
netic field H at the measure points and the anisotropy parameters to be
inverted from them. The sensitivity of both 2-MHz [180] and induction [35]
tools to anisotropy in deviated wells has been documented using the ANIS-
BEDS code.

6.3 Inversion in layered media using existing 2-MHz


measurements

The geometry considered is the TI anisotropic layered medium formation of


Figure 6.3. In this section, the objective is to invert for the horizontal and
vertical resistivity in each layer using measurements made by an existing
tool, the Compensated Dual Resistivity (CDR) tool.

6.3.1 2-MHz tool response in anisotropic media

The CDR tool broadcasts a 2-MHz electromagnetic wave alternately from


two transmitters and measures the amplitudes and phases of the complex
voltages at two receivers. The antenna configuration is shown in Figure 6.4.
The receiver voltage is related to Hz through a constant factor involving the
294 CHAPTER 6. PARAMETRIC INVERSION

Borehole

Transmitter 1

Measure Receiver 1

Point Receiver 2

Transmitter 2

Figure 6.4: Antenna configuration for the 2-MHz CDR tool.

coil turns [180]. Between the receivers, the complex voltages have a phase
shift and attenuation which vary depending on the formation resistivity. The
phase shift and attenuation generated by Transmitter 1 between Receiver 1
and Receiver 2, and by Transmitter 2 between Receiver 2 and Receiver 1,
are averaged together to symmetrize the response. The averaged phase shift
and attenuation are then converted to two separate resistivities: RPS (phase
shift, shallow) and RAD (attenuation, deep). The conversion procedure
uses a look-up table based on polynomial approximations of computed tool
response in homogeneous media of known isotropic resistivity, Rt . Thus the
conversion of the raw voltages to resistivity involves assumptions which can
complicate the interpretation of anisotropy.
When anisotropy is present in deviated wells or dipping formations, the
RPS measurement reads higher than the RAD measurement (see Section
1.5, and Figure 1.9 in particular). The amount of separation between the
two curves varies as a function of the deviation or dip angle, Rh and Rv .
Figure 6.5 shows the sensitivity of RPS and RAD to variations in the hori-
zontal resistivity level (Rh ) and the anisotropy contrast (Rv /Rh ) when the
deviation angle between the tool and formation is 75◦ . Figure 6.6 shows the
6.3. 2-MHZ INVERSION IN LAYERED MEDIA 295

Figure 6.5: CDR sensitivity to anisotropy at 75◦ dip.

sensitivity at 90◦ . The larger separation between the curves at 90◦ indicates
a greater sensitivity. (At 0◦ , the curves collapse to the diagonal Rv /Rh = 1
line.)
When bed boundaries are present, the interpretation of anisotropy be-
comes more complex since the resistivities in adjacent beds can have consid-
erable effect on the measurements in beds of interest. In deviated wells, the
effect of adjacent beds extends over much longer distances than in vertical
wells (see Section 2.1.5). Also, in highly deviated or horizontal wells, there
is the extra complication of polarization horns which occur at high contrast
bed boundaries. These horns can easily be mistaken for an additional bed
(see for example the large polarization horn at 0.0 depth in Figure 6.7).
In cases such as these, it is helpful to incorporate information from other
measurements to constrain resistivity levels and bed boundary locations.

6.3.2 The inversion algorithm

The inversion is an iterative approach based on the Gauss-Newton method.


The algorithm described here is adapted from Habashy, et al. [140, 136].
296 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.6: CDR sensitivity to anisotropy at 90◦ dip.

It employs a quadratic model of a cost function, which is minimized to


find the optimum solution. The cost function is defined as the square of
the sum of the relative residual errors given by the difference between the
logging data and the estimated response, normalized to the former. The
step length is adjusted so that the mismatch between the measured and
estimated responses is sufficiently decreased after each iteration. The step
length is computed by a line search algorithm.

Notation

The set of unknown model parameters is denoted by the vector x, such


that    
x1 Rh 1
 x2   
   Rv 1 
 x   
   Rh 2 
  
3

Model parameters = x =  xn = Rv 2 
   . (6.18)
 ..   .. 
 .   . 
   
 xN −1  
 Rh N/2


xN Rv N/2
6.3. 2-MHZ INVERSION IN LAYERED MEDIA 297

For the problem under consideration, adjacent odd-even subscripts of x cor-


respond to Rh and Rv pairs in the same bed. N is the number of unknowns,
with N/2 being the number of beds assumed in the formation model. The
vector of model parameters x is represented as the difference between the
vector of actual model parameters y and a background reference model y R ,
that is,
x = y − yR. (6.19)
The reference model includes all a priori information about the model pa-
rameters, such as information derived from independent measurements.
The set of measured data points (log values) is denoted by the vector m,
such that,
   
m1 RP S 1
   
 m2   RAD1 
   
 m3   RP S 2 
   
Data points = m =  mm = RAD2 
   . (6.20)
 ..   .. 
 .   . 
   
 mM −1  
 RP S M/2


mM RADM/2

Adjacent odd-even subscripts of m represent RP S (phase shift) and RAD


(attenuation) apparent resistivity measurement pairs at the same depth
point. M is the number of data points, with M/2 being the number of
RP S-RAD measurement pairs.
S (x
x) is the vector of M simulated tool response values generated as a
function of the model parameters x. e(xx) is defined as the vector of resid-
uals whose m-th element is the residual error (also referred to as the data
mismatch) of the m-th measurement. The residual error is defined as the
difference between the measured and predicted normalized responses
   
e1 (x x) x) − m1
S1 (x
 x)   x) − m2 
 e2 (x   S2 (x 
   
x) = 
e(x em (x x) = x) − mm
Sm (x  = S (x
x) − m. (6.21)
   
 ..   .. 
 .   . 
x)
eM (x x) − mM
SM (x
298 CHAPTER 6. PARAMETRIC INVERSION

The cost function

x),
The inversion is posed as the minimization of the cost function C (x
which has the form
1 
x) =
C (x ee(x x2 .
x)2 + λx (6.22)
2
The scalar factor λ (0 < λ < ∞) is a regularization parameter for deter-
mining the relative importance of the two terms of the cost function. The
choice of λ, described in a subsequent section, will produce an estimate of
the model x that has a finite minimum norm and which globally fits the data.
The second term of the cost function is included to regularize the optimiza-
tion problem. It safeguards against cases when measurements are redundant
or lacking sensitivity to certain model parameters causing a nonunique solu-
tion. It also suppresses any possible magnification of errors in the parameter
estimation due to noise which is unavoidable present in the measurements.
The inverted model parameters x are constrained to be within their phys-
ical bounds using a nonlinear transformation which is described in a later
section. Such a nonlinear transformation maps a constrained minimization
problem to an unconstrained one.

Normalization of the vector of residuals

Two forms of the cost function of Equation (6.22) are employed to put
the various measurements on equal footings. The two forms differ in the way
the vector of residuals e(x
x) is defined. In the first form, e(x
x) is defined as

x)
Sm (x
x) =
em (x − 1, (6.23)
mm
and therefore & &2
M &
 &
x)
& Sm (x &
ee(x 2
x) = & − 1& . (6.24)
& mm &
m=1
In the second form
x) − mm
Sm (x
x) =
em (x , (6.25)
x
x / M
and therefore M
m=1 |Sm (x
x ) − mm |2
x) =  
ee(x 2  . (6.26)
m=1 |mm |
M 2 / M2
6.3. 2-MHZ INVERSION IN LAYERED MEDIA 299

The Newton minimization approach

To solve the above nonlinear optimization problem, we employ a Newton


minimization approach [125] which is based on a local quadratic model of
the cost function. The quadratic model is formed by taking the first three
terms of the Taylor series expansion of the cost function around the current
xk ) as follows
k-th iteration (x
1 T
xk + pk ) ≈ C(x
C(x xk ) · pk +
xk ) + g T (x p · G(x
xk ) · pk , (6.27)
2 k
where the superscript T indicates transposition, and pk = xk+1 − xk is the
step size in xk toward the minimum of the cost function C(x x). The vector
x) = ∇C(x
g (x x) is the gradient vector of the cost function C(x
x) and is given
by the expression
 
g (x x) = gn ≡ ∂xn C,
x) = ∇C(x n = 1, 2, 3, . . . , N
= J T (x
x) · e(x
x) + λ x, (6.28)

x) is the M × N
where xn is the n-th component of the model vector x. J (x
Jacobian matrix given by
 
x) = Jm, n ≡ ∂xn em ,
J (x m = 1, 2, 3, . . . , M ; n = 1, 2, 3, . . . , N ,
(6.29)
or  
∂x1 S1 ∂xn S1 . . . ∂xN S1
 
 ∂x1 Sm ∂xn Sm . . . ∂xN Sm 
x) = 
J (x  .. .. .. ,
 (6.30)
 . . ... . 
∂x1 SM ∂xn SM . . . ∂xN SM
where e is the normalized form of S . Since the most time-consuming part
of the inversion is the calculation of the elements of the Jacobian, it is im-
portant to perform these computations efficiently. We compute the elements
numerically with good results by taking the difference between tool response
computed with ANISBEDS at the points xn and 1.01xn .
x) = ∇∇C(x
In Equation (6.27), G(x x) is the Hessian of the cost function
C(x) which is a real symmetric N × N matrix given by
x
 
G(x x) = Gi, j ≡ ∂x2n xm C,
x) = ∇∇C(x i, j = 1, 2, 3, . . . , N
= λ I + J T (x
x) · J (x
x) + Q(x
x). (6.31)
300 CHAPTER 6. PARAMETRIC INVERSION

In the above equation


M
x) =
Q(x x) F Tm (x
em (x x), (6.32)
m=1

x) the m-th element of the vector of residuals em (x


with em (x x), and
 
x) = ∇∇em (x
F m (x x) = ∂x2i xj em , i, j = 1, 2, 3, . . . , N . (6.33)

The minimum of the right hand side of Equation (6.27) is achieved if pk is


a minimum of the quadratic function
1 T
xk ) · p +
φ(pp) = g T (x p · G(x
xk ) · (pp). (6.34)
2
The function φ(pp) has a stationary point (a minimum, a maximum or a
saddle point) at p0 only if

∇φ(pp0 ) = G · p0 + g = 0. (6.35)

Therefore, the stationary point is the solution to the set of linear equations

G · p0 = −gg . (6.36)

The Gauss-Newton minimization approach

In the Gauss-Newton search method, one discards the second order deriva-
tives because they are computationally expensive to generate. In this case,
the Hessian reduces to

x) = λ I + J T (x
G(x x) · J (x
x), (6.37)

which is a positive semi-definite matrix. The Hessian G can be constructed


to be a positive definite matrix by the proper choice of λ, which will be
described in a later section.
The search direction pk , which is given by the vector that solves Equa-
tion (6.36) with the Hessian approximated by Equation (6.37), is called the
Gauss-Newton search direction. The method in which this vector is used as
a search direction is called the Gauss-Newton search. The Gauss-Newton
minimization approach has a rate of convergence which is slightly less than
6.3. 2-MHZ INVERSION IN LAYERED MEDIA 301

quadratic but significantly better than linear. It provides quadratic conver-


gence in the neighborhood of the minimum.

Line searches

The search vector pk of Equation (6.36) is guaranteed to be a descent


direction for the approximated quadratic form of the cost function of Equa-
tion (6.22). However, the step (i.e., the new iterate xk + pk ) may not suffi-
ciently reduce the cost function of Equation (6.22) and may not even decrease
its value, indicating that C(xx) is poorly modeled by a quadratic form in the
vicinity of xk . One approach to alleviate this problem [85] is to adopt a
line search algorithm where one searches for an appropriate real positive
step length νk along the search direction pk which yields an acceptable next
iterate, xk+1 = xk + νk pk , that sufficiently decreases the cost function.
In selecting this step length, we adopt an algorithm where a step length
νk > 0 is found which reduces the cost function. The procedure first employs
the full Gauss-Newton search step. If νk = 1 fails to reduce the cost function,
we backtrack by reducing νk along the direction of the Gauss-Newton step
until an acceptable next iterate xk+1 = xk + νk pk is found.
(l)
If, at the (k + 1)-th iteration, νk is the current step length and it does
not reduce the cost function, we compute the next backtracking step length
(l+1)
νk by searching for the minimum of the function

f (ν) ≡ C(x
xk + ν pk ), (6.38)

which can be approximated by a quadratic expression as

f (ν) ≈ a + b ν + c ν 2 . (6.39)

The real constants a, b and c are determined from the current information
x) as follows:
on the cost function C(x

xk ),
f (ν = 0) = C(x (6.40)
&
xk + ν pk )&&
dν f (ν = 0) = δ Ck+1 = ∂ν C(x xk ) · pk ,
= g T (x (6.41)
ν=0
and
(l)
xk + νk(l) pk ).
f (ν = νk ) = C(x (6.42)
This gives
xk ),
a = C(x (6.43)
302 CHAPTER 6. PARAMETRIC INVERSION

b = δ Ck+1 , (6.44)
and  
1
c=  2 xk + νk(l)pk ) − C(x
C(x xk ) − νk(l) δ Ck+1 . (6.45)
(l)
νk
(l+1)
Thus νk , which is the minimum of f (ν) for l = 0, 1, 2, . . . is given by
 
(l) 2
(l+1) b νk δ Ck+1
νk =− =− (l) (l)
, (6.46)
2c 2 xk +
C(x νk p k ) − C(x
xk ) − νk δ Ck+1

xk + νk(l)pk ) > C(x


from which it can be seen that if C(x xk ), then

(l+1) 1 (l) 1
0 < νk < νk < m+1 , m = 0, 1, 2, . . . . (6.47)
2 2
(0)
Thus we start with νk = 1 and proceed with the backtracking procedure
of Equation (6.46) until the cost is reduced. In general, it is not desirable to
(l+1)
decrease νk too much since this may excessively slow down the iterative
process, requiring many iterations to achieve little progress toward the min-
(l+1) (l) (l+1) (l)
imum. To prevent this slow down, we set νk = 0.1 νk if νk < 0.1 νk
and then procees with the iteration. In addition, νk in not allowed to de-
crease below 0.1 (i.e., νmin = 0.1 to guard against an excessively small value
of ν.

The choice of the regularization parameter

Several criteria exist for selecting the regularization parameter λ. One


such criterion [138] is to apply a steepest descent method in the initial steps of
the iteration process. This corresponds to choosing large values for λ which
puts more weight of the second term of the cost function of Equation (6.22)
since the first term is only crudely approximated by the quadratic model
of Equation (6.27). As the iteration progresses, the reconstructed model
approached its true value. This results in Equation (6.27) becoming more
accurate, and therefore more weight (corresponding to small values of λ)
should be placed on minimizing the cost function. One of the working criteria
for choosing λ is to bound it by the inequality

maximum (small τm  s)  λ  minimum (large τm  s), (6.48)


6.3. 2-MHZ INVERSION IN LAYERED MEDIA 303

where τm are the eigenvalues of the positive definite real symmetric matrix

H = J T (x
x) · J (x
x). (6.49)

The second part of the inequality in Equation (6.48) guarantees that the
spectral content of the inversion operator remains unaltered, while the first
part of the inequality regularizes the inversion problem by suppressing the
null-space of the inversion operator.

Criteria for terminating the iteration process

The iteration process is stopped when one of the following conditions


occurs:

1. The root mean square of the relative error reaches a prescribed value
η determined from estimates of noise in the data, i.e.,
 1/2
1
ee2 ≤ η, (6.50)
M
where η is a predetermined a priori value that is provided by the user.
In the hypothetical case of noise free data, η = 0.
2. The difference between two successive iterates, (k + 1) and k, of the
model parameters is within a prescribed tolerance factor, tol, of the
current iterate, that is,

|xj, k+1 − xj, k | ≤ tol × |xj, k+1 |, j = 1, 2, 3, . . . , N. (6.51)

3. The number of iterations exceeds a prescribed maximum.


4. The difference between the cost function at two successive iterates,
(k + 1) and k, of the model parameters is within a prescribed tolerance
factor, tole, of the cost function at the current iterate, that is,

xk+1 ) − C(x
|C(x xk )| ≤ tole × C(x
xk+1 ). (6.52)

Nonlinear transformations for constrained minimization

There are a large number of nonlinear transformations which can map


a constrained minimization problem to an unconstrained one. Among the
304 CHAPTER 6. PARAMETRIC INVERSION

many that were tried, the following transformation was found to work the
best.
If xmax is an upper bound on the model parameter x, and xmin is a lower
bound, then in order to ensure that xmin < x < xmax at all iterations, we
introduce the transformation
xmax − xmin 2
x = xmin + c , −∞ < c∞. (6.53)
c2 + 1
It is apparent that

x → xmin , as c → 0, (6.54)
x → xmax , as c → ±∞. (6.55)

It can be shown that


xmax − x  1/2
∂c Sm = dc x ∂x Sm = 2 (xmax − x)(x − xmin ) ∂x Sm , (6.56)
xmax − xmin
where Sm is the m-th measurement. The two successiver iterates, xk+1 and
xk of x are related by
xmax − xmin 2 xmax − xmin
xk+1 = xmin + ck+1 = xmin + (ck + qk )2 (6.57)
c2k+1 + 1 (ck + qk )2 + 1

where ) *1/2
xk − xmin
ck = , (6.58)
xmax − xk
and qk = ck+1 −ck is the Gauss-Newton search step in c toward the minimum
of the cost function. Defining
xmax − x  1/2
p=2 (xmax − x)(x − xmin ) q = dc x q, (6.59)
xmax − xmin
we obtain the following relationship between two successive iterates xk+1 and
xk of x (assuming as adjustable step length νk along the search direction
xmax − xmin
xk+1 = xmin + α2 , (6.60)
αk2 + (xk − xmin )(xmax − xk )3 k

where
1
αk = (xk − xmin )(xmax − xk ) + (xmax − xmin ) νk pk . (6.61)
2
6.3. 2-MHZ INVERSION IN LAYERED MEDIA 305

Figure 6.7: CDR computed logs in Figure 6.8: Results of inversion for
an isotropic bed above an anisotropic Rh and Rv from the CDR computed
bed at 80◦ dip. logs in Figure 6.7.

Note that
xk+1 → xmax , if xk → xmax or xmin , (6.62)
The variable p defined by Equation (6.59) is the solution of Equation (6.36).
Finally, it should be noted that the transformation of Equation (6.53) in-
troduces false minima at x = xmax and x = xmin , since ∂c Sm vanishes at
both x = xmax and x = xmin . Note also that (from Equation (6.62)) this
transformation skews the emphasis toward xmax rather than toward xmin .

6.3.3 2-MHz inversion results

The inversion algorithm was tested using simulated logs generated by ANIS-
BEDS. The simulated logs for a single interface case are shown in Figure 6.7,
and the corresponding inversion results are shown in Figure 6.8. The objec-
tive is to invert for the horizontal and vertical resistivities within each bed
from the two apparent resistivity logs. It is assumed that the dip (or devia-
306 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.9: Effect of error in bed Figure 6.10: Effect of error in bed
boundary detection for 2-foot data boundary detection for 2-foot data
sampling with points at the bound- sampling with points at the bound-
ary included in the inversion. ary omitted from the inversion.

tion) angle can be obtained from a dipmeter or imaging log. Bed boundary
locations are estimated from inflection points on the logs for small dip angles,
or from peak values of polarization horns for large dip angles. In this case,
the bed boundary is assumed to be within ±1 inch of the boundary of the
simulated log. The initial values for the model parameters in each bed are
set equal to the log center-bed readings; RAD is used as the initial estimate
for Rh and the normally higher RPS is used for Rv , For both Rh and Rv ,
the lower limit xmin is constrained to be 0.2 ohm-m, and the upper limit
xmax is constrained to be 2000 ohm-m, which are the physical limits of the
field tool readings. Rv is not constrained to be greater than Rh The agree-
ment between the modeled and inverted resistivities shown in Figure 6.8 is
excellent.
The effect of errors in the bed boundary location was then systematically
studied for this case. Figure 6.9 shows the inverted Rv and Rh obtained
with data points at the bed boundary included in the inversion. Figure 6.10
similar results, only in this case the data points at the boundary have been
omitted from the inversion. In both cases, the data was sampled at 2-foot in-
tervals in true vertical depth. The curves terminate where the inversion was
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 307

unable to find a solution in 50 iterations. Although results are quite accurate


when data points at the boundary are included (Figure 6.9), the accuracy
deteriorates rapidly when errors in the boundary location are greater than 3
inches. Reasonable but less accurate results are obtained for 10 inch errors
in bed boundary location when data points at the bed boundary are omitted
from the inversion (Figure 6.10).

6.4 Inversion in layered media using triaxial measure-


ments

In this section, the objective is to invert for the horizontal and vertical
resistivity in each layer of a TI anisotropic formation using measurements
made by a tool with triaxial antennas. The geometry to be considered is
the layered TI anisotropic formation shown in Figure 6.3 (the same type
of geometry that was used for the CDR inversion. Dipole antennas are
assumed, with the dipole moments at right angles to one another as shown
in Figure 6.2. To focus on the inversion results and avoid tool design issues, a
tool with a single triaxial transmitter and a single triaxial receiver is modeled
and inverted. An optimized tool would make use of either mutually balanced
induction focusing or 2-MHz voltage differencing. The responses of both
of these types of multi-coil tools can easily be constructed by combining
measurements made with several different antenna spacings.
In order to reveal potential problems associated with the inversion of
triaxial induction data, it is useful to first examine the relative sensitivities
of coils with x, y and z dipole orientations to basic environmental effects.
These effects are: skin effect, anisotropy in homogeneous media, radial depth
of investigation and bed boundary effect.

6.4.1 Triaxial tool response in some limiting cases

Conventional induction tools are composed of coaxial coils as shown in Fig-


ure 6.1. In borehole logging, coaxially oriented coils are commonly called
vertical magnetic dipoles (VMD). Induction coils are also equivalent to the
Tz -Rz components of the antennas shown in Figure 6.2. The response of
induction tools has been studied in great detail since the late 1940’s and the
results of these studies were already summarized in this thesis.
308 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.11: Response of two-coil VMD (left) and HMD (right) sondes in
homogeneous isotropic media.

Tools composed of coplanar coils are commonly called horizontal mag-


netic dipoles (HMD) or transverse magnetic dipoles (TMD). The Tx -Rx and
the Ty -Ry antennas shown in Figure 6.2 are examples of coplanar coils. (The
different directionality of the Tx -Rx and Ty -Ry antennas is not relevant in
the axisymmetric case.) HMD tools have been periodically investigated as a
source of additional information about the logging environment [193]. How-
ever, HMD tools have had limited success in borehole logging, primarily
because their logs exhibit large excursions near bed boundaries in vertical
wells (see for example Figure 6.23), which makes interpretation extremely
difficult. The recent intensified interest in anisotropy, along with advances
in real-time inversion, have made the extra effort needed to interpret and
process HMD measurements less problematical. Since the response of HMD
antennas has not been addressed elsewhere in this thesis, it will be briefly
examined here before describing the inversion results.

Homogeneous isotropic media (skin effect)

Figure 6.11 compares the response of two-coil VMD (left) and HMD
(right) sondes in homogeneous isotropic media. Both sondes have a spacing
of 40 inches and operate at a frequency of 20 kHz. Note that the HMD R-
signal curve becomes non-linear at lower true formation conductivities than
the corresponding VMD curve. In addition, the HMD R-signal curve be-
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 309

Figure 6.12: Two-coil VMD sensitivity to anisotropy at 45◦ dip for a 20 kHz
tool.

comes double-valued at formation conductivities above 5000 mS/m because


of attenuation. This indicates that HMD measurements are more subject
to skin effect than VMD measurements made using the same frequency and
spacing. (HMD and VMD response in both isotropic and anisotropic homo-
geneous media was calculated using the closed-form solution of Moran and
Gianzero [193].)

Homogeneous anisotropic media

VMD response in homogeneous anisotropic media is examined first to


serve as a comparison for HMD response. Figure 6.12 shows the sensitivity
of a 40-inch two-coil conventional induction (VMD) sonde to anisotropy when
the deviation angle between the tool and the formation is 45◦ . The frequency
is 20 kHz. The induction R-signal and X-signal are plotted for a range of
horizontal conductivities (σh ) and anisotropy contrasts (σh /σv ). The small
amount of separation between the curves indicates that a VMD sonde has
almost no sensitivity to anisotropy at 45◦ dip. (At 0◦ dip, all curves collapse
to the σh /σv = 1 curve.)
Figure 6.13 shows the sensitivity of the same VMD sonde to anisotropy
when the deviation angle between the tool and the formation is 90◦ . It is
310 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.13: Two-coil VMD sensitivity to anisotropy at 90◦ dip for a 20 kHz
tool.

apparent that the sensitivity increases significantly as the dip angle increases.
Figure 6.14 shows the sensitivity of a 40-inch two-coil VMD sonde to
anisotropy at 90◦ dip, only here the frequency has been increased to 100
kHz. A comparison of Figure 6.13 and Figure 6.14 shows that the sensi-
tivity to anisotropy at 100 kHz decreases compared to that at 20 kHz. The
decrease in sensitivity is indicated by the smaller area spanned by the curves
in Figure 6.14. In addition, some of the curves in Figure 6.14 are double-
valued and negative, which could cause problems when inverting this data (a
voltage difference measurement can cancel out these problems to a certain
extent).
The next three figures examine anisotropy effect for HMD sondes. Fig-
ure 6.15 shows the sensitivity of a 40-inch two-coil HMD sonde to anisotropy
when the deviation angle between the tool and the formation is 0◦ . The
frequency is 20 kHz. The large amount of separation between the curves
indicates that a HMD sonde has significant sensitivity to anisotropy at 0◦
dip.
Figure 6.16 shows the sensitivity of a 40-inch two-coil HMD sonde to
anisotropy at 0◦ dip, only here the frequency has been increased to 100 kHz.
A comparison of Figure 6.15 and Figure 6.16 shows that the sensitivity to
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 311

Figure 6.14: Two-coil VMD sensitivity to anisotropy at 90◦ dip for a 100
kHz tool.

Figure 6.15: Two-coil HMD sensitivity to anisotropy at 0◦ dip for a 20 kHz


tool.
312 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.16: Two-coil HMD sensitivity to anisotropy at 0◦ dip for a 100 kHz
tool.

anisotropy at 100 kHz decreases compared to that at 20 kHz, as it also


does for the VMD. For both the HMD and the VMD, the primary cause of
the decrease in sensitivity at the higher frequency is signal attenuation and
phase shift caused by skin effect. Skin effect can of course be reduced by
either lowering the frequency or decreasing the coil spacings. However, any
reduction in skin effect obtained by decreasing coil spacings is is bought at
the cost of a shallower depth of investigation.
Figure 6.17 shows the sensitivity of a 40-inch two-coil HMD sonde to
anisotropy when the deviation angle between the tool and the formation is
75◦ . The frequency is 20 kHz. At 75◦ dip, the sensitivity has decreased
significantly compared to that at 0◦ (Figure 6.15). In Figure 6.17 the curves
have all shifted toward the σh /σv = 1 curve, and when the dip angle increases
to 90◦ , they will all collapse to that curve. Thus VMD and HMD sensitivity
to anisotropy is complimentary; for small dip angles where the VMD is least
sensitive, the HMD is most sensitive, and for large dip angles where the
VMD is most sensitive, the HMD is least sensitive.
Figure 6.15 through Figure 6.17 addressed HMD response to anisotropy
with the horizontal transmitter and receiver dipoles oriented in the y direc-
tion, as illustrated in Figure 6.18 (a) for 0◦ dip. At 90◦ dip when the HMD
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 313

Figure 6.17: Two-coil HMD sensitivity to anisotropy at 75◦ dip for a 20 kHz
tool.

loses sensitivity to σv , the transmitter and receiver are oriented with the
dipole moments parallel to σv as illustrated in Figure 6.18 (b). HMD sensi-
tivity to σv at 90◦ dip can be recovered by rotating the horizontal dipoles so
that their moments are no longer parallel to σv , as shown in Figure 6.18 (c).

(a) (b) (c)


σv σv σv

σh σh σh

MyR
T T
T My MyR My MyR
My

Figure 6.18: HMD antennas with dipoles perpendicular (a), parallel (b) and
(c) at 45◦ to the direction of σv .
314 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.19: Two-coil HMD sensitivity to anisotropy at 90◦ dip for a 20 kHz
tool. The antennas are rotated 45◦ .

Figure 6.19 shows the response of a HMD sonde to anisotropy when the
antennas are rotated 45◦ . The sensitivity to anisotropy is greatly improved.
Thus both x-x and y-y dipole components (which can be combined to create
a rotated HMD) are needed to find a unique solution in cases where the dip
angle rotates outside the z-y plane (arbitrary strike).

Figure 6.20: Geometry for modeling depth of investigation.


6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 315

Figure 6.21: Comparison of the depth of investigation of two-coil HMD and


VMD sondes at 20 kHz.

Radial depth of investigation

Depth of investigation can be examined by modeling tool response to


invasion while systematically increasing the invasion radius. Low conductiv-
ities were used in this formation to avoid complicating the examination of
depth of investigation with skin effect. A shallow tool will read Rt (or σt )
only when the invasion radius is small. However, a deep tool will read near
Rt even when the invasion radius is relatively large. The geometry modeled
to compare VMD and HMD depth of investigation is shown in Figure 6.20.
Figure 6.21 compares the depth of investigation of 40-inch VMD and
HMD two-coil sondes. The frequency of both tools is 20 kHz. A quick
look at Figure 6.21 seems to indicate that the HMD is significantly deeper
than the VMD since it reads consistently closer to σt for large invasion
radii. However, note that HMD response exceeds σt for invasion radii less
that 20 inches. The offset created by this excursion above σt accentuates
the differences in depth of investigation between the HMD and VMD. For a
HMD, cylindrial boundaries act as waveguides which distort their response to
invasion and can also cause severe borehole effect when the contrast between
the borehole and formation conductivity is large.
Figure 6.22 compares the depth of investigation of two 40-inch HMD two-
coil sondes with one operating at 2 kHz and the other at 200 kHz. There is
signifant attenuation in the HMD signal at 200 kHz, which causes 200 kHz
316 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.22: Comparison of the depth of investigation of two-coil HMD


sondes with one operating at 2 kHz and the other at 200 kHz.

measurement to be much shallower than the 2 kHz measurement. Borehole


effect would be greater at 200 kHz, since the excursion above σt at small
radii is larger.

Bed boundary effect

Figure 6.23, Figure 6.24, Figure 6.25 and Figure 6.26 illustrate HMD
and VMD response in a thin bed at 0◦ , 30◦ , 60◦ and 89◦ dip respectively.
The spacing between the transmitter and receiver coils is 40 inches and the
frequency is 20 kHz. Three apparent resistivity curves are plotted in each
figure. Ra−zz is the apparent resistivtity for a z-directed transmitter and
receiver (VMD). Ra−yy is the apparent resistivtity for a y-directed transmit-
ter and receiver (HMD). Ra−xx is the apparent resistivtity for a x-directed
transmitter and receiver (a HMD with the dipole moments perpendicular to
the y-z plane, as illustrated in Figure 6.2). Only three couplings are shown
for the sake of clarity. The R-signals are plotted because they scale reason-
ably well to resistivity. (The R-signal is derived from the imaginary part of
H and the X-signal from the real part of H .)
The formation modeled is an 8-foot anisotropic bed surrounded by isotropic
shoulder beds of unequal resistivity (the logs in this formation will be in-
verted in the following section). The resistivity values in each bed are indi-
cated by the straight lines labeled Rh and Rv . The logs were not corrected
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 317

Figure 6.23: Response of two-coil x-x, Figure 6.24: Response of two-coil x-x,
y-y and z-z dipole sondes in an aniso- y-y and z-z dipole sondes in an aniso-
tropic formation at 0◦ dip. tropic formation at 30◦ dip.

Figure 6.25: Response of two-coil x-x, Figure 6.26: Response of two-coil x-x,
y-y and z-z dipole sondes in an aniso- y-y and z-z dipole sondes in an aniso-
tropic formation at 60◦ dip. tropic formation at 89◦ dip.
318 CHAPTER 6. PARAMETRIC INVERSION

for skin effect. The larger amount of skin effect on the HMD measurements
is apparent from the greater departure of the Ra−xx and Ra−yy curves from
Rt in the conductive shoulder beds (higher Ra translates to lower σa due to
attenuation).
In Figures 6.23 through Figure 6.26, as the dip angle increases, Ra−zz
becomes more sensitive to Rv , and the curve grows closer to Rv in the center
bed. Also as the dip angle increases, Ra−xx and Ra−yy become less sensitive
to Rv , and these curves depart further from Rv in the center bed. (At 0◦ dip,
Ra−xx and Ra−yy are equal because of azimuthal symmetry.) The relative
departures of the apparent resistivity curves from Rv and Rh are difficult to
attribute to anisotropy because there is also considerable shoulder effect in
the 8-foot bed.
Note also the large excursions (horns) on the Ra−xx and Ra−yy curves
at 0◦ dip, which become smaller but do not disappear as the dip angle
increases. These horns make HMD logs difficult to interpret in both vertical
and deviated wells. Ra−zz exhibits the familiar “polarization” horns at 89◦
dip which were apparent at high dip on CDR logs.

6.4.2 Triaxial inversion results

A tool with a single spacing of 40 inches (approximately one meter) between


triaxial antennas is modeled and inverted. The set of measured data points,
m, is composed of the nine complex components of the magnetic field H
as measured by the triaxial tool (i.e., an x-directed transmitter with x, y
and z-directed receivers, a y-directed transmitter with x, y and z-directed
receivers, and a z-directed transmitter with x, y and z-directed receivers).
This yields 18 data points at each logging station, as compared to 2 points
at each station for the CDR as given in Equation (6.20). Unless otherwise
noted, the interval between log samples is one foot in true vertical depth.
Certain couplings of H are zero in degenerate cases. To accelerate inver-
sion run time, these couplings can be omitted in the inversion.
For TI anisotropy with no dip and no strike (i.e., θ = 0◦ and φ = 0◦ ), H
takes the form  
Hxx 0 0
H = 0 Hyy 0 , (6.63)
0 0 Hzz
with Hxx = Hyy because of azimuthal symmetry.
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 319

For TI anisotropy with dip but no strike (i.e., θ


= 0◦ and φ = 0◦ ), H
takes the form  
Hxx 0 Hxz
H = 0 Hyy 0 . (6.64)
Hzx 0 Hzz
In the special case of a homogeneous anisotropic medium, Hxz = Hzx .
For TI anisotropy with both dip and strike (i.e., θ
= 0◦ and φ
= 0◦ ), H
is a full matrix and takes the form
 
Hxx Hxy Hxz
H =  Hyx Hyy Hyz  . (6.65)
Hzx Hzy Hzz

In the special case of a homogeneous anisotropic medium, Hxz = Hzx , Hyx =


Hxy and Hzy = Hyz . For all other cases, Hyx = Hxy because of the assumed
TI anisotropy.
The the triaxial inversion was implemented using two different sets of
the unknown model parameters, x.

- Case 1: The set of x values is composed of the horizontal (Rh ) and


vertical (Rv ) resistivities in each bed, as given by Equation (6.18). It
is assumed that the bed boundary locations can be estimated with
sufficient accuracy, and that the dip and strike angles can be obtained
by an independent measurement such as a dipmeter or imaging tool.
- Case 2: The set of x values is composed of the dip (θ) and strike (φ)
angles and the locations of all bed boundaries (zm ), in addition to Rh
and Rv in each bed. Thus, in this case, all formation parameters are
assumed to be unknown.

Simulated logs consisting of the full H matrix were modeled with ANIS-
BEDS at 1-foot intervals in true vertical depth and then inverted. In all
cases, the center-bed values of Ra−zz were used as the first guess for Rh , and
the corresponding center-bed values of Ra−yy were used as the first guess
for Rv . The modeled data were truncated at 4 significant figures to better
approximate the accuracy of field data. Several different bed thicknesses and
resistivity levels were modeled and inverted. However, only results for the
8-foot bed cases shown in Figures 6.23 through 6.26 will be examined here
for the sake of brevity, since they represent typical solutions.
320 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.27: Triaxial inversion results Figure 6.28: Triaxial inversion results
for the formation in Figure 6.25 with a for the formation in Figure 6.25 with
+5◦ error in the dip angle. a +3 inch error in the bed boundaries.

Figure 6.29: Triaxial inversion results Figure 6.30: Triaxial inversion results
for the formation in Figure 6.25 using for the formation in Figure 6.25 using
only the center-bed values of H . only the diagonal values of H .
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 321

Figures 6.27 through 6.30 shows a set of inversion results for the 8-foot
bed at 60◦ dip (and 0◦ strike) modeled in Figure 6.27. Each of these figures
illustrates errors in the inversion for Rh and Rv caused by making a different
erroneous assumption about the data. If it is assumed that the dip and
strike angles and bed boundary locations are known exactly, the parametric
inversion can obtain Rh and Rv in each bed to 4 significant figures in 4
iterations, using the full H matrix sampled a 1-foot intervals (these results
are not shown here because the figure is very similar to Figure 6.29 and not
very interesting).
Figure 6.27 shows the inversion results when a +5◦ error has been made
in the assumed dip angle. The errors in Rh and Rv are significant, but the
results are still reasonable. Errors of −5◦ in the assumed dip angle generate
comparable errors in Rh and Rv .
Figure 6.28 shows the inversion results when a systematic +3 inch error
has been made in both of the assumed bed boundary locations. The errors in
Rh and Rv are similar to those made by assuming the wrong dip angle. For
this 40-inch tool, errors greater than 3 inches in the boundary locations cause
a rapid deterioration in the accuracy of Rh and Rv . The error in Rh and Rv is
greater when one boundary is moved too far in the positive direction and the
adjacent boundary is moved too far in the negative direction, in comparison
to a systematic movement of both boundaries too far in the same direction
(i.e., the error is less when the true bed thickness is preserved, even if the
bed is moved). Also, the error in Rh and Rv is generally greater when a
bed is estimated to be too small rather than too large. However, even with
these difficulties, the inversion of triaxial data with bed boundary errors is
much more stable than CDR inversion with comparable boundary errors,
converging to a reasonable but incorrect solution in less than 10 iterations.
Returning to the assumption that the dip and strike angles and bed
boundary locations are known, Figure 6.29 shows the inversion results when
only the center bed values for H are used in the inversion. The error in Rh
and Rv is less than 2%, and the small difference between the actual formation
parameters and the inversion results is not apparent in the figure.
Figure 6.30 shows the inversion results when only the diagonal values
of H (i.e., Hxx , Hyy and Hzz as given in Equation (6.64)) are used in the
inversion. The interval between log samples is again one foot. Since the
terms that have been omitted are small in comparison to the diagonal terms
in this case, the inversion results are quite good. For the results shown in
322 CHAPTER 6. PARAMETRIC INVERSION

Figures 6.27 through 6.30, each inversion takes less than one minute to run
on a Sun SPARC Ultra 30 workstation.
Because inaccuracies in estimating bed boundary locations and the dip
angle are the major sources of errors in the inversion, and because triaxial
measurements provide a large amount of data with good sensitivity to the
dipping layered anisotropic environment, it is reasonable to next attempt a
simultaneous inversion for the bed boundary locations and dip and strike
angles, along with Rh and Rv . The vector of unknown model parameters
x was modified to include the dip and strike angles and all bed boundary
locations, in addition to Rh and Rv in each bed. Thus, in this case x becomes
   
x1 Rh 1
 x2   Rv1 
   
   
 x3   Rh 2 
   
 xn   Rv2 
   
 ..   .. 
 .   . 
   
 x2NB−1   R 
   h 2NB/2 
   
x= x2NB  =  Rv 2NB/2  , (6.66)
   
 x2NB+1   θ 
   
 x2NB+2   φ 
   
   
 x2NB+3   z1 
   
 x2NB+4   z2 
   
 ..   .. 
 .   . 
xN zNB−1

where NB is the number of beds in the formation model, θ is the dip angle, φ
is the strike angle and zm are the bed boundary locations. N is the number
of unknowns, with N = 2NB + 2 + (NB − 1). For both the dip and strike
angles, the lower limit xmin is constrained to be 0◦ and the upper limit xmax
is constrained to be 90◦ . The bed boundaries are constrained to be within
±2 feet of the initial guess. The constraints on Rh and Rv of xmin = 0.2
ohm-m and xmax = 2000 ohm-m still apply.
Results of the inversion for all formation parameters will be shown for
two 8-foot bed cases: a small dip angle (30◦ ) and a large dip angle (89◦ ).
Figure 6.31 shows the modeled apparent resistivity curves for the x-x, y-y
and z-z couplings at 30◦ dip and 60◦ strike. The response of the z-z coupling
is the same as it is at 30◦ dip and 0◦ strike (shown in Figure 6.24). The
responses of the x-x and y-y couplings at 60◦ strike differ from those at 0◦
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 323

Figure 6.31: Response of two-coil x-x, Figure 6.32: Triaxial inversion re-
y-y and z-z dipole sondes in an aniso- sults for the formation in Figure 6.31
tropic formation at 30◦ dip, 60◦ strike. using the full H matrix.

Figure 6.33: Triaxial inversion results Figure 6.34: Triaxial inversion re-
for the formation in Figure 6.31 using sults for the formation in Figure 6.31
only the center-bed values of H . using only the diagonal values of H .
324 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.35: Response of two-coil x-x, Figure 6.36: Triaxial inversion re-
y-y and z-z dipole sondes in an aniso- sults for the formation in Figure 6.35
tropic formation at 89◦ dip, 30◦ strike. using the full H matrix.

Figure 6.37: Triaxial inversion results Figure 6.38: Triaxial inversion re-
for the formation in Figure 6.35 using sults for the formation in Figure 6.35
only the center-bed values of H . using only the diagonal values of H .
6.4. TRIAXIAL INVERSION IN LAYERED MEDIA 325

strike by approximately 5% in the center of the anisotropic bed, and by as


much as 80% near the bed boundaries. Strike as a cause of departures on logs
near bed boundaries has also been documented in a study of crossbedding
(different dip and strike angles in each bed) in vertical wells [34]. At 0◦
strike, Hxy , Hyx , Hyz and Hzy are zero. At 60◦ strike, these terms are one
to three orders of magnitude less than the diagonal values of H .
Figure 6.32 shows the inversion results using the full H matrix. The
values obtained from the inversion for Rh , Rv , dip, strike and bed boundary
locations are within 1% of the modeled formation parameters. The initial
guesses for the dip and strike angles were 30◦ from the actual solution, and
the initial guesses for the bed boundary locations were 0.5 feet from the
solution. Using these guesses, the inversion converges in 4 iterations. If
initial guesses are further from the solution, the inversion converges to the
same solution in 5 to 6 iterations.
Figure 6.33 shows the inversion results when only the center-bed values of
H are used in the inversion. The accuracy has deteriorated in comparison to
the results for 1-foot sampling (Figure 6.32) because the center-bed readings
do not contain enough information to pin down the bed boundary locations.
Figure 6.34 shows the inversion results with 1-foot sampling when only
the diagonal values of H are used in the inversion. The diagonal terms have
sufficient sensitivity to strike to provide an accurate solution. However, in
this case the initial guess for the strike had to be within 20◦ of the actual
solution. If the initial guess for the strike was farther from the solution,
the inversion sometimes gave a strike value of 90◦ (xmax ). Thus discarding
the off-diagonal terms eliminates information that is needed to accurately
resolve strike.
Figure 6.35 shows the modeled apparent resistivity curves for the x-x, y-y
and z-z couplings for the large dip case of 89◦ dip with 30◦ strike. Again, the
response of the z-z coupling is the same as the corresponding response at 0◦
strike (Figure 6.26). The responses of the x-x and y-y couplings at 30◦ strike
differ from those at 0◦ strike by a smaller amount than the previous case:
a 3% differences in the center of the anisotropic bed, and approximately a
50% difference near the bed boundaries. At 30◦ strike, Hxy , Hyx , Hyz and
Hzy are one to three orders of magnitude less than the diagonal values of H .
Figure 6.36 shows the inversion results at 89◦ dip using the full H matrix.
As in the previous 30◦ dip case, the values obtained from the inversion for
Rh , Rv , dip, strike and bed boundary locations are within 1% of the modeled
326 CHAPTER 6. PARAMETRIC INVERSION

formation parameters. The inversion converges in 5 iterations.


Figure 6.37 shows the inversion results at 89◦ dip when only the center-
bed values of H are used in the inversion. The accuracy has deteriorated
about the same amount as in the previous 30◦ dip case.
Figure 6.38 shows the inversion results at 89◦ dip with 1-foot sampling
when only the diagonal values of H are used in the inversion. The accuracy
is slightly worse than for the previous 30◦ dip case.

6.5 Summary and future plans

The CDR inversion results described in Section 6.3.3 demonstrated that a


vertical magnetic dipole does not provide enough sensitivity to invert for the
anisotropic resistivities Rh and Rv in layered media at all dip angles. Even
at large dip angles where the CDR sensitivity to anisotropy is greatest,
the CDR inversion becomes unstable when bed boundary locations cannot
be determined with certainty. It was shown in Section 6.4.2 that triaxial
measurements contain enough information to invert not only for Rh and
Rv , but for the dip and strike angle and bed boundary locations as well.
However, since horizontal magnetic dipoles have significant sensitivity to
borehole and invasion effects, there is no guarantee that the good inversion
results obtained in layered media will be duplicated when radial boundaries
are included in the model.
During the past few years, there has been much work done to improve
the speeds of 3D codes (such as MAXANIS) that are used to model triaxial
tool response in anisotropic media. Run-times of several minutes per logging
station are not uncommon, making it possible to use these codes for inversion
in a research environment. However, further acceleration is necessary before
3D codes can be used in the routine interpretation of triaxial tool response.
Although 3D inversion in anisotropic media is outside the scope of this thesis,
it will be studied during the coming year by this author. Future 3D inversion
must also include testing in cases where the dip and strike angles are different
in each medium (i.e., crossbedding).
We have seen that horizontal magnetic dipoles have much different re-
sponse characteristics than vertical magnetic dipoles. Therefore, it is useful
to briefly compare their responses in a layered medium with a borehole to
get a better perspective for potential problems associated with the inversion
6.5. SUMMARY AND FUTURE PLANS 327

Figure 6.39: Geometry for modeling HMD borehole effect in a vertical well.

of triaxial response when radial boundaries are present.


Figure 6.39 shows the geometry modeled to compare HMD and VMD
borehole effect: a vertical well with a resistive borehole passing through a
single interface separating an isotropic bed and an anisotropic bed. The
modeled response of a two-coil HMD tool is shown in Figure 6.40 (R-signal)
and Figure 6.41 (X-signal). Also shown for comparison are modeled logs
computed without the borehole. (The logs were generated with the program
MAXANIS, and the no-borehole results were additionally validated using
ANISBEDS.) A moderately small coil spacing of 20 inches is used to accen-
tuate borehole effect. The frequency is 20 kHz. The logs are plotted on a
linear scale because the R-signal becomes negative near the interface. There
is significant borehole effect on the HMD R-signal in the lower conductive
bed as shown by the large difference between the response with and without
the borehole. The presence of the borehole has also decreased the size of the
excursions near the bed boundary. The deeper HMD X-signal shows very
little borehole effect.
Figure 6.42 and Figure 6.43 show the corresponding R-signal and X-signal
logs for a two-coil VMD sonde with the same coil spacing and frequency
as the HMD. There is much less borehole effect on the VMD R-signal in
the conductive bed in comparison to the HMD. Note in addition that the
328 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.40: Two-coil HMD R-signal Figure 6.41: Two-coil HMD X-signal
with and without borehole effect for with and without borehole effect for
a single interface separating isotropic a single interface separating isotropic
and anisotropic beds. and anisotropic beds.

VMD curve is fairly smooth as it crosses the bed boundary. This makes
it possible to decouple borehole effect from other formation effects and to
perform an accurate borehole correction prior to inversion, thus eliminating
the borehole parameters from the inversion algorithm. The irregular shape
of the HMD R-signal as the tool crosses the boundary makes it unclear
whether an independent borehole correction could be performed accurately.
(The current Baker-Atlas triaxial tool inversion for vertical wells first obtains
a solution for layered media and then separately corrects for borehole and
invasion effects [164].)
There is also the question of whether invasion effect on the HMD is so
severe that it limits the tool’s sensitivity to the parameters of primary inter-
est: Rh and Rv in the uninvaded formation. Figure 6.44 shows the geometry
modeled to examine HMD sensitivity to anisotropy beyond an invaded zone.
The invasion radius in systematically increased to see if the tool can read
the Rh and Rv values in the uninvaded formation when the radius becomes
6.5. SUMMARY AND FUTURE PLANS 329

Figure 6.42: Two-coil VMD R-signal Figure 6.43: Two-coil VMD X-signal
with and without borehole effect for with and without borehole effect for
a single interface separating isotropic a single interface separating isotropic
and anisotropic beds. and anisotropic beds.

large. Figure 6.45 shows the apparent horizontal (Rha ) and vertical (Rva )
resistivities for a two-coil HMD sonde with a 40-inch coil spacing. The fre-
quency is 20 kHz. Rha and Rva were obtained using a Newton-Raphson
algorithm based on tool response in homogeneous anisotropic media. Rha
and Rva are near Rh and Rv for invasion radii less than 20 inches. Sensitiv-
ity to anisotropy gradually decreases and disappears as the invasion radius
approaches 40 inches.
The tool can of course be made to read deeper by increasing the distance
between the transmitter and receiver coils. Figure 6.46 shows the modeled
response for a 80-inch coil spacing. The frequency is also 20 kHz. There
is good sensitivity to anisotropy up to a radius of 40 inches, and sensitivity
gradually disappears as the invasion radius approaches 80 inches. These
results are in agreement with the isotropic cases shown in Figure 6.21, and
with induction modeling described in Section 3.2.1 which shows that a tool’s
average depth of investigation corresponds to approximately half the coil
330 CHAPTER 6. PARAMETRIC INVERSION

Figure 6.44: Geometry for modeling sensitivity to anisotropy beyond an


invaded zone.

spacing. A Baker-Atlas model study [163] reached similar conclusions; they


showed that the sensitivity pattern of Hxx to Rv is similar to that of Hzz to
Rh (i.e., the induction response function shown in Figure 3.6).
Ever since the first electrical log was run in 1927, the major goal has been
to design a tool that can accurately give the resistivity of each layer pen-
etrated by the wellbore with a minimum of user intervention. It was soon
discovered that this is not a simple task; currents circulate preferentially
within the borehole rather than inside beds, borehole mud creeps into the
formation through invasion, beds of economic interest become smaller as re-
covery methods improve, and the resistivity is no longer a single number but
a tensor because of anisotropy. The ideal resistivity inversion would provide
a map of the formation of the quality of a medical CAT-scan image. This
cannot be accomplished by illuminating the formation with a uni-directional
dipole located inside the borehole. Since logging tools are restricted to re-
side inside the borehole, the only way to illuminate the formation at multiple
angles is with triaxial or similar multi-directional antennas. This will pro-
vide the additional information needed to make major improvements in the
accuracy of borehole resistivity inversion.
6.5. SUMMARY AND FUTURE PLANS 331

Figure 6.45: Sensitivity of a 40-inch two-coil HMD sonde to anisotropy be-


yond an invaded zone.

Figure 6.46: Sensitivity of an 80-inch two-coil HMD sonde to anisotropy


beyond an invaded zone.
332 CHAPTER 6. PARAMETRIC INVERSION
Bibliography

[1] A. Abubakar. Three-Dimensional Nonlinear Inversion of Electrical


Conductivity. PhD thesis, Delft University, 2000.

[2] A. Abubakar and P. M. van den Berg. Non-linear three-dimensional in-


version of cross-well electrical measurements. Geophysical Prospecting,
48(1):109–134, 2000.

[3] A. Abubakar and P. M. van den Berg. Three-dimensional extended


contrast source inversion. In Proceedings of the 2000 International
Symposium on Antennas and Propagation. Institute of Electronics, In-
formation and Communication Engineers, 2000. Volume E00-A.

[4] A. Abubakar and P. M. van den Berg. Three-dimensional inverse scat-


tering applied to cross-well induction sensors. IEEE Transactions on
Geoscience and Remote Sensing, 38(4):1669–1681, 2000.

[5] L. A. Allaud and M. H. Martin. Schlumberger: The History of a


Technique. John Wiley and Sons, New York, 1977.

[6] L. A. Allaud and J. Ringot. The high-resolution dipmeter tool. The


Log Analyst, 10(3):3–10, 1969.

[7] D. Allen, F. Auzerais, E. Dussan, P. Goode, T. S. Ramakrish-


nan, L. Schwartz, D. Wilkinson, E. Fordham, P. Hammond, and
R. Williams. Invasion revisited. Schlumberger Oilfield Review, 3(3):10–
23, 1991.

[8] D. Allen, D. Bergt, D. Best, B. Clark, I. Falconer, J. M. Hache,


C. Kienitz, M. Lesage, J. Rasmus, C. Roulet, and P. Wraight. Logging
while drilling. Schlumberger Oilfield Review, 1(1):4–17, 1989.
334 BIBLIOGRAPHY

[9] D. Allen, R. Dennis, J. Edwards, S. Franklin, J. Livingston, A. Kirk-


wood, J. White, L. Lehtonen, B. Lyon, and G. Simms. Modeling logs
for horizontal well planning and evaluation. Schlumberger Oilfield Re-
view, 7(4):47–63, 1995.

[10] D. F. Allen. Laminated sand analysis. In SPWLA 25th Annual Logging


Symposium Transactions. Society of Professional Well Log Analysts,
1984. Paper XX.

[11] D. F. Allen, B. I. Anderson, T. D. Barber, Q. H. Liu, and M. G. Lüling.


Supporting interpretation of complex, axisymmetric invasion by mod-
eling wireline induction and 2-MHz LWD resistivity tools. In SPWLA
34th Annual Logging Symposium Transactions. Society of Professional
Well Log Analysts, 1993. Paper U.

[12] D. F. Allen and M. G. Lüling. Integration of wireline resistivity data


with dual depth of investigation 2-MHz resistivity data. In SPWLA
30th Annual Logging Symposium Transactions. Society of Professional
Well Log Analysts, 1989. Paper C.

[13] L. M. Alpin. The method of the electric logging in the borehole with
casing, 1939. U.S. Patent 56,026.

[14] B. Anderson. The analysis of some unsolved induction interpretation


problems using computer modeling. The Log Analyst, 27(5):60–73,
1986.

[15] B. Anderson and T. Barber. Strange induction logs–a catalog of envi-


ronmental effects. In SPWLA 28th Annual Logging Symposium Trans-
actions. Society of Professional Well Log Analysts, 1987. Paper G.

[16] B. Anderson and T. Barber. Using computer modeling to provide


missing information for interpreting resistivity logs. In SPWLA 29th
Annual Logging Symposium Transactions. Society of Professional Well
Log Analysts, 1988. Paper H.

[17] B. Anderson and T. Barber. Deconvolution and boosting parameters


for obsolete Schlumberger induction tools. The Log Analyst, 40(2):133–
137, 1999.

[18] B. Anderson, T. Barber, V. Druskin, P. Lee, E. Dussan, L. Knizh-


nerman, and S. Davydycheva. The response of multiarray induction
BIBLIOGRAPHY 335

tools in highly dipping formations with invasion and in arbitrary 3D


geometries. The Log Analyst, 40(5):327–344, 1999.

[19] B. Anderson, T. Barber, J. Singer, and T. Broussard. ELMOD–putting


electromagnetic modeling to work to improve resistivity log interpre-
tation. In SPWLA 30th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1989. Paper M.

[20] B. Anderson, S. Bonner, M. Lüling, and R. Rosthal. Response of 2-


MHz LWD restivity and wireline induction tools in dipping beds and
laminated formations. The Log Analyst, 33(5):461–475, 1992.

[21] B. Anderson, I. Bryant, M. Lüling, B. Spies, and K. Helbig. Oilfield


anisotropy: Its origins and electrical characteristics. Schlumberger Oil-
field Review, 6(4):48–56, 1994.

[22] B. Anderson and S. Chang. Synthetic induction logs by the finite


element method. The Log Analyst, 23(6):17–26, 1982.

[23] B. Anderson and W. C. Chew. A new high speed technique for cal-
culating synthetic induction and DPT logs. In SPWLA 25th Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 1984. Paper HH.

[24] B. Anderson and W. C. Chew. A general semi-analytic method for


the rapid computer simulation of the response of borehole electrical
logging tools. In SEG 55th Annual Meeting Transactions. Society of
Exploration Geophysicists, 1985. Paper BHG 4.3.

[25] B. Anderson and W. C. Chew. SFL interpretation using high speed


synthetic computer generated logs. In SPWLA 26th Annual Logging
Symposium Transactions. Society of Professional Well Log Analysts,
1985. Paper K.

[26] B. Anderson and W. C. Chew. Transient response of some borehole


mandrel tools. Geophysics, 54(2):216–224, 1989.

[27] B. Anderson, V. Druskin, T. Habashy, P. Lee, M. Lüling, T. Barber,


G. Grove, J. Lovell, R. Rosthal, J. Tabanou, D. Kennedy, and L. Shen.
New dimensions in modeling resistivity. Schlumberger Oilfield Review,
9(1):40–56, 1997.
336 BIBLIOGRAPHY

[28] B. Anderson, V. Druskin, P. Lee, M. G. Lüling, E. Schoen, J. Tabanou,


P. Wu, S. Davydycheva, and L. Knizhnerman. Modeling 3D effects on
2-MHz LWD resistivity logs. In SPWLA 38th Annual Logging Sym-
posium Transactions. Society of Professional Well Log Analysts, 1997.
Paper N.

[29] B. Anderson and S. Gianzero. Induction sonde response in stratified


media. The Log Analyst, 24(1):25–31, 1983.

[30] B. Anderson, Q. H. Liu, R. Taherian, J. Singer, W. C. Chew, R. Freed-


man, and T. Habashy. Interpreting the response of the Electromagnetic
Propagation Tool in heterogeneous environments. The Log Analyst,
35(2):65–83, 1994.

[31] B. Anderson, G. Minerbo, M. Oristaglio, T. Barber, B. Freedman,


and F. Shray. Modeling electromagnetic tool response. Schlumberger
Oilfield Review, 4(3):22–32, 1992.

[32] B. Anderson, K. A. Safinya, and T. Habashy. Effects of dipping beds


on the response of induction tools. In Proceedings 1986 SPE Annual
Technical Conference and Exhibition. Society of Petroleum Engineers,
1986. Paper SPE 15488.

[33] B. I. Anderson and T. D. Barber. Induction Logging. Schlumberger,


Houston, TX, 1997. Document SMP-7056.

[34] B. I. Anderson, T. D. Barber, and S. C. Gianzero. The effect of


crossbedding anisotropy on induction tool response. Petrophysics,
42(2):137–149, 2001.

[35] B. I. Anderson, T. D. Barber, and M. G. Lüling. The response of


induction tools to dipping, anisotropic formations. In SPWLA 36th
Annual Logging Symposium Transactions. Society of Professional Well
Log Analysts, 1995. Paper D.

[36] B. I. Anderson, M. G. Lüling, R. Taherian, and T. Habashy. The


response of dielectric logging tools to dipping thin beds. In Proceed-
ings 1994 SPE Annual Technical Conference and Exhibition. Society
of Petroleum Engineers, 1994. Paper SPE 28438.

[37] B. I. Anderson, K. A. Safinya, T. Habashy, A. Davidson, and


R. Gilmore. The response of the Electromagnetic Propagation Tool
BIBLIOGRAPHY 337

to bed boundaries. In Proceedings 1987 SPE Annual Technical Con-


ference and Exhibition. Society of Petroleum Engineers, 1987. Paper
SPE 16768.
[38] Anonymous. 3D Explorer induction logging service. Baker Hughes
sales brochure, 1999.
[39] G. E. Archie. The electrical resistivity log as an aid in determining
some reservoir characteristics. In Transactions of the AIME, Petroleum
Division, volume 146, pages 54–62. American Institute of Mining and
Metallurgical Engineers, 1942.
[40] I. Asimov. Understanding Physics. Dorset Press, New York, 1988.
[41] E. A. Badea. 3D finite element analysis of induction logging. In SEG
69th Annual Meeting Transactions. Society of Exploration Geophysi-
cists, 1999. Paper EM3.2.
[42] T. Barber. Introduction to the Phasor dual induction tool. Journal of
Petroleum Technology, 37(10):1699–1706, 1985.
[43] T. Barber. Real time environmental corrections for the Phasor induc-
tion tool. In SPWLA 26th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1985. Paper EE.
[44] T. Barber. Induction log vertical resolution enhancement–physics and
limitations. In SPWLA 29th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1988. Paper O.
[45] T. Barber and A. Q. Howard. Correcting the induction log for dip
effect. In Proceedings 1989 SPE Annual Technical Conference and
Exhibition. Society of Petroleum Engineers, 1989. Paper SPE 19607.
[46] T. Barber, A. Orban, G. Hazen, T. Long, R. Schlein, S. Alderman,
J. Tabanou, and J. Seydoux. A multiarray induction tool optimized
for efficient wellsite operation. In Proceedings 1995 SPE Annual Tech-
nical Conference and Exhibition. Society of Petroleum Engineers, 1995.
Paper SPE 30583.
[47] T. Barber and R. Rosthal. Using a multiarray induction tool to achieve
logs with minimum environmental effects. In Proceedings 1991 SPE
Annual Technical Conference and Exhibition. Society of Petroleum En-
gineers, 1991. Paper SPE 22725.
338 BIBLIOGRAPHY

[48] T. D. Barber. Phasor processing of induction logs including shoulder


and skin effect correction, 1984. U.S. Patent 4,513,376.

[49] T. D. Barber, T. Broussard, G. Minerbo, Z. Sijercic, and D. Mur-


gatroyd. Interpretation of multiarray induction logs in invaded for-
mations at high relative dip angles. The Log Analyst, 40(3):202–217,
1999.

[50] T. D. Barber, R. Chandler, and J. Hunka. Induction logging with


metallic sonde support, 1987. U.S. Patent 4,651,101.

[51] T. D. Barber, W. Vandermeer, and W. Flanagan. Method for deter-


mining induction sonde error, 1989. U.S. Patent 4,800,496.

[52] K. J. Bartenhagen, J. C. Bradford, and D. Logan. Cased hole formation


resistivity: Changing the way we find oil and gas. In Proceedings
2001 SPE Permian Basin Oil and Gas Recovery Conference. Society
of Petroleum Engineers, 2001. Paper SPE 70042.

[53] D. R. Beard, Q. Zhou, and E. L. Bigelow. A new, fully digital, full-


spectrum induction device for determining accurate resistivity with
enhanced diagnostics and data integrity verification. In SPWLA 37th
Annual Logging Symposium Transactions. Society of Professional Well
Log Analysts, 1996. Paper B.

[54] D. R. Beard, Q. Zhou, and E. L. Bigelow. Practical application of


a new multichannel and fully digital spectrum induction system. In
Proceedings 1996 SPE Annual Technical Conference and Exhibition.
Society of Petroleum Engineers, 1996. Paper SPE 36504.

[55] P. Béguin, D. Benimeli, A. Boyd, I. Dubourg, A. Ferreira, A. Mc-


Dougall, G. Rouault, and P. van der Wal. Recent progress on forma-
tion resistivity measurement through casing. In SPWLA 41st Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 2000. Paper CC.

[56] R. Beste, T. Hagiwara, G. King, R. Strickland, and G. A. Merchant.


A new high resolution array induction tool. In SPWLA 41st Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 2000. Paper C.
BIBLIOGRAPHY 339

[57] S. Bonner, A. Bagersh, B. Clark, M. Dajee, M. Dennison, M. Fredette,


O. Grogan, J. S. Hall, J. Jundt, E. Kwok, J. R. Lovell, R. A. Rosthal,
and D. A. Allen. A new generation of electrode resistivity measure-
ments for formation evaluation while drilling. In SPWLA 35th Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 1994. Paper OO.

[58] S. D. Bonner, J. R. Tabanou, P. T. Wu, J. P. Seydoux, K. A. Moriarty,


B. K. Seal, E. Y. Kwok, and M. W. Kuchenbecker. New 2-MHz mul-
tiarray borehole-compensated resistivity tool developed for MWD in
slim holes. In Proceedings 1996 SPE Annual Technical Conference and
Exhibition. Society of Petroleum Engineers, 1996. Paper SPE 30547.

[59] A. Boyd, H. Darling, J. Tabanou, B. Davis, B. Lyon, C. Flaum,


J. Klein, R. M. Sneider, A. Sibbit, and J. Singer. The lowdown on
low resistivity pay. Schlumberger Oilfield Review, 7(3):4–18, 1995.

[60] J. M. Bricaud and A. Poupon. Continuous dipmeter survey: The pote-


clinometer and the microfocused devices. In Transactions 5th World
Petroleum Congress, pages 225–239, 1959.

[61] T. J. Calvert, R. N. Rau, and L. E. Wells. Electromagnetic propa-


gation: A new dimension in logging. In Proceedings 1977 SPE Cali-
fornia Regional Meeting. Society of Petroleum Engineers, 1977. Paper
SPE 6542.

[62] S. Chang and B. Anderson. Simulation of induction logging by the


finite element method. Geophysics, 49(11):1943–1958, 1984.

[63] M. V. K. Chari and P. P. Silvester. Finite Elements in Electrical and


Magnetic Field Problems. John Wiley and Sons, New York, 1980.

[64] R. Chemali, S. Gianzero, and S. M. Su. The effect of shale anisotropy


of focused resistivity devices. In SPWLA 28th Annual Logging Sym-
posium Transactions. Society of Professional Well Log Analysts, 1987.
Paper H.

[65] R. Chemali, S. Gianzero, and S. M. Su. The Dual Laterolog in com-


mon complex situations. In SPWLA 29th Annual Logging Symposium
Transactions. Society of Professional Well Log Analysts, 1988. Pa-
per N.
340 BIBLIOGRAPHY

[66] P. Cheung, D. Pittman, A. Hayman, R. Laronga, P. Vessereau,


A. Ounadjela, O. Desport, S. Hansen, R. Kear, M. Lamb, T. Borbas,
and B. Wendt. Field test results of a new oil-base mud formation im-
ager tool. In SPWLA 42nd Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 2001. Paper XX.

[67] W. C. Chew. Modeling of the high frequency dielectric logging tool:


Theory, applications and results. IEEE Transactions on Geoscience
and Remote Sensing, 26(4):382–398, 1988.

[68] W. C. Chew. Waves and Fields in Inhomogeneous Media. VanNostrand


Reinhold, New York, 1990.

[69] W. C. Chew and B. Anderson. Propagation of electromagnetic waves


through geological beds in a geophysical probing environment. Radio
Science, 20(3):611–621, 1985.

[70] W. C. Chew, S. Barone, B. Anderson, and C. Hennessy. Diffraction


of axisymmetric waves in a borehole by bed boundary discontinuities.
Geophysics, 49(10):1586–1595, 1984.

[71] W. C. Chew and S. L. Chuang. Profile inversion of a planar medium


with a line source or a point source. In Proceedings IGARSS ’84.
IGARSS, 1984. Strasbourg, France, Aug. 27-30.

[72] W. C. Chew and Q. H. Liu. Inversion of induction tool measurements


using the distorted Born iterative method and conjugate gradient fast
Fourier-Hankel transform. IEEE Transactions on Geoscience and Re-
mote Sensing, 32(4):878–884, 1994.

[73] W. C. Chew, Z. Nie, Q. H. Liu, and B. Anderson. An efficient so-


lution for the response of electrical well logging tools in a complex
environment. IEEE Transactions on Geoscience and Remote Sensing,
29(2):308–313, 1991.

[74] B. Clark, D. F. Allen, D. Best, S. D. Bonner, J. Jundt, M. G. Lüling,


and M. O. Ross. Electromagnetic propagation logging while drilling:
Theory and experiment. In Proceedings 1988 SPE Annual Technical
Conference and Exhibition. Society of Petroleum Engineers, 1988. Pa-
per SPE 18117.
BIBLIOGRAPHY 341

[75] B. Clark, M. G. Lüling, J. Jundt, M. Ross, and D. Best. A dual


depth resistivity measurement for formation evaluation while drilling.
In SPWLA 29th Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1988. Paper A.

[76] L. Combee. Transient Diffusive Electromagnetic Fields in Layered An-


isotropic Media. PhD thesis, Delft University, 1991.

[77] D. Coope, L. C. Shen, and F. S. Huang. The theory of 2 MHz resis-


tivity tool and its application to measurement while drilling. The Log
Analyst, 25(3):35–46, 1984.

[78] P. T. Cox and W. F. Warren. Development and testing of the Texaco


dielectric tool. In SPWLA 24th Annual Logging Symposium Transac-
tions. Society of Professional Well Log Analysts, 1983. Paper H.

[79] D. H. Davies, O. Faivre, M. T. Gounot, B. Seeman, J. C. Trouiller,


D. Benimeli, A. E. Ferreira, D. J. Pittman, J. W. Smits, M. Randri-
anavony, B. I. Anderson, and J. Lovell. Azimuthal resistivity imaging:
A new generation laterolog. In Proceedings 1992 SPE Annual Techni-
cal Conference and Exhibition. Society of Petroleum Engineers, 1992.
Paper SPE 24676.

[80] P. J. Davis and P. Rabinowitz, editors. Methods of Numerical Integra-


tion. Academic Press, New York, 1975.

[81] S. Davydycheva and V. Druskin. Staggered grid for Maxwell’s equa-


tions in arbitrary 3D inhomogeneous media. In Proceedings of the In-
ternational Symposium on Three-Dimensional Electromagnetics, pages
181–189, 1995.

[82] S. Davydycheva, V. Druskin, and T. Habashy. Solution of Maxwell’s


equations in an arbitrary 3D inhomogeneous anisotropic media with
sharp discontinuities using a finite difference scheme on a regular
Cartesian grid with material averaging. Schlumberger-Doll Research
internal report, September 1996. Report EMG-002-96-15.

[83] P. deChambrier. The microlog continuous dipmeter. Geophysics,


18(4):929–951, 1953.
342 BIBLIOGRAPHY

[84] A. T. deHoop. Electromagnetic radiation from sources in piecewise ho-


mogeneous structures with non-concentric circularly cylindrical bound-
aries (theory). Schlumberger-Doll Research internal report, March
1983. Project MEL-003.

[85] J. E. Dennis and R. B. Schnabel. Numerical Methods for Uncon-


strained Optimization and Nonlinear Equations. Prentice Hall, En-
glewood Cliffs, NJ, 1983.

[86] J. T. Dewan. Essentials of Modern Open-hole Log Interpretation.


PennWell Books, Tulsa, OK, 1983.

[87] A. J. deWitte and D. A. Lowitz. Theory of the induction log. In SP-


WLA 2nd Annual Logging Symposium Transactions. Society of Pro-
fessional Well Log Analysts, 1961. Paper 9.

[88] H. G. Doll. Introduction to induction logging and application to log-


ging of wells drilled with oil base mud. Journal of Petroleum Technol-
ogy, 1(6):148–162, 1949.

[89] H. G. Doll. The micro-laterolog. Journal of Petroleum Technology,


5(1):17–32, 1950.

[90] H. G. Doll. The microlog. In Transactions of the AIME, volume


189, pages 155–164. American Institute of Mining and Metallurgical
Engineers, 1950.

[91] H. G. Doll. The laterolog: A new resistivity logging method with


electrodes using an automatic focusing system. Journal of Petroleum
Technology, 3(11):305–316, 1951.

[92] H. G. Doll. Differential coil system for induction logging, 1952. U.S.
Patent 2,582,315.

[93] H. G. Doll. Electromagnetic well logging system, 1952. U.S. Patent


2,582,314.

[94] H. G. Doll. Electrical logging apparatus, 1955. U.S. Patent 2,712,628.

[95] H. G. Doll. Electrical resistivity well logging method and apparatus,


1955. U.S. Patent 2,712,627.
BIBLIOGRAPHY 343

[96] H. G. Doll. Filtrate invasion in highly permeable sands. The Petroleum


Engineer, January 1955.

[97] H. G. Doll. Methods and apparatus for electrical logging of wells, 1955.
U.S. Patent 2,712,630.

[98] H. G. Doll. Method and apparatus for providing improved vertical


resolution in induction well logging including electrical storage and
delay means, 1965. U.S. Patent 3,166,709.

[99] H. G. Doll, J. L. Dumanoir, and M. Martin. Suggestions for better


electric log combinations and improved interpretation. Geophysics,
25(4):854–882, 1960.

[100] V. Druskin and L. Knizhnerman. A spectral semi-discrete method


for the numerical solution of 3D nonstationary problems in electri-
cal prospecting. Izvestia Academia Science USSR, Physics of Solid
Earth, 24(8):641–648, 1988. (English edition published by American
Geophysical Union).

[101] V. Druskin and L. Knizhnerman. Spectral approach to solving three-


dimensional Maxwell’s diffusion equations in the time and frequency
domains. Radio Science, 29(4):937–953, 1994.

[102] W. C. Duesterhoeft. Propagation effects in induction logging. Geo-


physics, 26(2):192–204, 1961.

[103] W. C. Duesterhoeft and H. W. Smith. Propagation effects on radial


response in induction logging. Geophysics, 27(4):463–469, 1962.

[104] A. Dumont, M. Kubacsi, and J. L. Chardac. The oil-based mud dip-


meter tool. In SPWLA 28th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1987. Paper LL.

[105] E. Dussan, B. Anderson, and F. Auzerais. Estimating vertical perme-


ability from resistivity logs. In SPWLA 35th Annual Logging Sympo-
sium Transactions. Society of Professional Well Log Analysts, 1994.
Paper UU.

[106] C. J. Dyos. Inversion of the induction log by the method of maximum


entropy. In SPWLA 28th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1987. Paper T.
344 BIBLIOGRAPHY

[107] H. N. Edmundson and H. Urrows. Raymond Sauvage: Recollections


of Schlumberger wireline’s first four years. The Technical Review,
35(1):4–15, 1987.

[108] P. Eisenmann, M. T. Gounot, B. Juchereau, J. C. Trouiller, and S. J.


Whittaker. Improved Rxo measurements through semi-active focusing.
In Proceedings 1994 SPE Annual Technical Conference and Exhibition.
Society of Petroleum Engineers, 1994. Paper SPE 28437.

[109] M. Ekstrom, C. A. Dahan, M. Y. Chen, P. M. Lloyd, and D. J. Rossi.


Formation imaging with microelectrical scanning arrays. The Log An-
alyst, 28(3):294–306, 1987.

[110] P. A. S. Elkington, J. R. Samworth, and M. C. Enstone. Vertical reso-


lution enhancement by combination and transformation of associated
responses. In SPWLA 31st Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1990. Paper HH.

[111] D. Ellis. Well Logging for Earth Scientists. Elsevier Science Publishing,
New York, 1987.

[112] A. A. Fitch, editor. Developments in Geophysical Exploration


Methods–3, chapter 6, pages 195–238. Applied Science Publishers,
London, 1982. “Electrical anisotropy: Its effect on well logs,” by J. H.
Moran and S. Gianzero.

[113] P. D. Fredericks, F. D. Hearn, and M. M. Wisler. Formation eval-


uation while drilling with a Dual Propagation Resistivity tool. In
Proceedings 1989 SPE Annual Technical Conference and Exhibition.
Society of Petroleum Engineers, 1989. Paper SPE 19622.

[114] R. Freedman and G. P. Grove. Interpretation of EPT logs in the


presence of mudcakes. SPE Formation Evaluation, 5(4):449–457, 1990.

[115] R. Freedman and G. N. Minerbo. Maximum entropy inversion of in-


duction log data. SPE Formation Evaluation, 6(2):259–268, 1991.

[116] R. Freedman and J. P. Vogiatzis. Theory of microwave dielectric


constant logging using the electromagnetic wave propagation method.
Geophysics, 44(5):969–986, 1979.
BIBLIOGRAPHY 345

[117] A. Fylling and J. Spurlin. Induction simulation, the log analysts’ per-
spective. In 11th European Formation Evaluation Symposium Trans-
actions. Norwegian Chapter of SPWLA, 1988. Paper T.

[118] S. Gianzero. Effect of sonde eccentricity on responses of conventional


induction logging tools. IEEE Transactions on Geoscience Electronics,
GE-16(4):332–339, 1978.

[119] S. Gianzero. The mathematics of resistivity and induction logging.


The Technical Review, 29(1):4–32, 1981.

[120] S. Gianzero and B. Anderson. An integral transform solution to the


fundamental problem in resistivity logging. Geophysics, 47(6):946–956,
1982.

[121] S. Gianzero and B. Anderson. A new look at skin effect. The Log
Analyst, 23(1):20–34, 1982.

[122] S. Gianzero and B. Anderson. Mathematical theory for the fields due
to a finite A.C. source in an infinitely thick bed with an arbitrary
number of co-axial layers. The Log Analyst, 25(3):25–32, 1984.

[123] S. Gianzero and Y. Lin. The effect of standoff on the response of in-
duction logging tools. In SPWLA 26th Annual Logging Symposium
Transactions. Society of Professional Well Log Analysts, 1985. Pa-
per FF.

[124] S. Gianzero, G. A. Merchant, M. Haugland, and R. Strickland. New de-


velopments in 2 MHz electromagnetic wave resistivity measurements.
In SPWLA 35th Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1994. Paper MM.

[125] P. E. Gill, W. Murray, and M. H. Wright. Practical Optimization.


Academic Press Limited, London, 1981.

[126] R. Gilmore, B. Clark, and D. Best. Enhanced saturation determina-


tion using the EPT-G Endfire antenna array. In SPWLA 28th Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 1987. Paper K.

[127] M. T. Gounot. Modilization des outils de résistivité. Schlumberger


Paris Engineering internal report, July 1984. Memo 2:12.
346 BIBLIOGRAPHY

[128] R. Griffiths, J. W. Smits, O. Faivre, I. Dubourg, E. Legendre, and


J. Doduy. Better saturation from new array laterolog. In SPWLA
40th Annual Logging Symposium Transactions. Society of Professional
Well Log Analysts, 1999. Paper DDD.
[129] G. P. Grove and G. N. Minerbo. An adaptive borehole correction
scheme for array induction tools. In SPWLA 32nd Annual Logging
Symposium Transactions. Society of Professional Well Log Analysts,
1991. Paper P.
[130] A. Gruner-Schlumberger. The Schlumberger Adventure. Arco Publish-
ing, New York, 1982.
[131] H. Guyod. Electric analogue of resistivity logging. Geophysics,
20(3):615–629, 1955.
[132] H. Guyod. Factors affecting the responses of laterolog-type logging
systems (LL3 and LL7). Journal of Petroleum Technology, 16(2):211–
219, 1964.
[133] H. Guyod and L. E. Shane. Geophysical Well Logging. Hubert Guyod,
Publisher, Houston, TX, 1969.
[134] A. Gyllensten and A. Boyd. Cased hole formation resistivity tool trial.
In Proceedings 2001 SPE Middle East Oil Show Conference. Society of
Petroleum Engineers, 2001. Paper SPE 68081.
[135] T. Habashy and B. Anderson. Reconciling differences in depth of inves-
tigation between 2-MHz phase shift and attenuation resistivity mea-
surements. In SPWLA 32nd Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1991. Paper E.
[136] T. Habashy, Y. Hsu, J. Xia, and J. C. Trouiller. Parametric inversion
of the Dual Laterolog. Schlumberger-Doll Research internal report,
November 1992. Report EMG-001-92-50.
[137] T. Habashy and M. G. Lüling. A point magnetic dipole radiator in a
layered, TI-anisotropic medium. Schlumberger-Doll Research internal
report, April 1994. Report EMG-001-94-14.
[138] T. M. Habashy, W. C. Chew, and E. Y. Chow. Simultaneous recon-
struction of permittivity and conductivity profiles in a radially inho-
mogeneous slab. Radio Science, 21(4):635–645, 1986.
BIBLIOGRAPHY 347

[139] T. M. Habashy, R. W. Groom, and B. R. Spies. Beyond the Born


and Rytov approximations: A nonlinear approach to electromagnetic
scattering problems. Journal of Geophysical Reaearch, 98(B2):1759–
1775, 1993.

[140] T. M. Habashy and T. S. Ramakrishnan. An inversion methodology


for MI2 : A weighted least squares constrained minimization approach.
Schlumberger-Doll Research internal report. Draft.

[141] R. Hakvoort, A. Fabris, M. Frenkel, J. Koelman, and A. Loermans.


Field measurements and inversion results of the High-Definition Lat-
eral Log. In SPWLA 39th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1998. Paper C.

[142] R. G. Hakvoort, A. Fabris, M. A. Frenkel, J. M. V. A. Koelman,


and A. M. Loermans. Field measurements and inversion results of
the High-Definition Lateral Log. In SPWLA 39th Annual Logging
Symposium Transactions. Society of Professional Well Log Analysts,
1998. Paper C.

[143] R. H. Hardman and L. C. Shen. Theory of induction sonde in dipping


beds. Geophysics, 51(3):800–809, 1986.

[144] R. H. Hardman and L. C. Shen. Charts for correcting effects of for-


mation dip and hole deviation on induction logs. The Log Analyst,
28(4):349–356, 1987.

[145] B. Harrison, editor. Russian-style Formation Evaluation, chapter 11,


pages 129–156. London Petrophysical Society, London, 1995. “Uno-
cused resistivity,” by M. Vincent and F. G. Williams.

[146] B. Harrison, editor. Russian-style Formation Evalustion, chapter 13,


pages 169–174. London Petrophysical Society, London, 1995. “Focused
resistivity (induction),” by M. Kennedy.

[147] J. R. Hearst and P. Nelson. Well Logging for Physical Properties.


McGraw Hill Book Co., New York, 1985.

[148] L. Henry and B. Seeman. Methods and apparatus for investigating


earth formations wherein a fixed relationship is maintained between
emitted current and measured potential difference, 1970. U.S. Patent
3,539,910.
348 BIBLIOGRAPHY

[149] A. Q. Howard. A new invasion model for resistivity log interpretation.


The Log Analyst, 33(2):96–110, 1992.

[150] J. S. Huchital, R. Hutin, Y. Thoraval, and B. Clark. The Deep


Propagation Tool (a new electromagnetic logging tool). In Proceed-
ings 1981 SPE Annual Technical Conference and Exhibition. Society
of Petroleum Engineers, 1981. Paper SPE 10988.

[151] J. Hunka, T. D. Barber, R. A. Rosthal, G. N. Minerbo, E. A. Head,


A. Q. Howard, G. A. Hazen, and R. N. Chandler. A new resistivity
measurement system for deep formation imaging and high-resolution
formation evaluation. In Proceedings 1990 SPE Annual Technical Con-
ference and Exhibition. Society of Petroleum Engineers, 1990. Paper
SPE 20559.

[152] G. B. Iskovich, A. Mezzatesta, K. M. Strack, and L. Tabarovsky. High-


Definition Lateral Log resistivity device: Basic physics and resolution.
In SPWLA 39th Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1998. Paper V.

[153] J. D. Jackson. Classical Electrodynamics. John Wiley and Sons, New


York, 1975.

[154] Y. M. Jan and R. L. Campbell. Borehole correction of MWD gamma


ray and short normal resistivity logs. In SPWLA 25th Annual Logging
Symposium Transactions. Society of Professional Well Log Analysts,
1984. Paper PP.

[155] A. A. Kaufman. Conductivity determination in a formation having a


cased well, 1989. U.S. Patent 4,796,186.

[156] A. A. Kaufman. The electric field in a borehole with a casing. Geo-


physics, 55(1):29–38, 1990.

[157] A. Kaufmann. Theory of Induction Logging. Siberian Dept. of Nauka


Press, Novosibersk, 1965.

[158] C. Kienitz, C. Flaum, J. R. Olesen, and T. Barber. Accurate logging


in large boreholes. In SPWLA 27th Annual Logging Symposium Trans-
actions. Society of Professional Well Log Analysts, 1986. Paper III.
BIBLIOGRAPHY 349

[159] J. Klein, P. Martin, and D. F. Allen. The petrophysics of electrically


anisotropic reservoirs. In SPWLA 36th Annual Logging Symposium
Transactions. Society of Professional Well Log Analysts, 1995. Pa-
per HH.
[160] R. L. Kleinberg, W. C. Chew, E. Y. Chow, B. Clark, and D. D. Grif-
fin. Microinduction sensor for the oil-based mud dipmeter. In Proceed-
ings 1987 SPE Annual Technical Conference and Exhibition. Society
of Petroleum Engineers, 1987. Paper SPE 16761.
[161] J. A. Kong. Electromagnetic Wave Theory. John Wiley and Sons, New
York, 1990.
[162] O. Kosenkov. Russian induction sondes, 1995. E-mail communication.
[163] B. F. Kriegshäuser, O. N. Fanini, S. Forgang, G. Iskovich, M. Ra-
binovich, L. Tabarovsky, L. Yu, M. Epov, P. Gupta, and J. van der
Horst. A new multicomponent induction logging tool to resolve aniso-
tropic formations. In SPWLA 41st Annual Logging Symposium Trans-
actions. Society of Professional Well Log Analysts, 2000. Paper D.
[164] B. F. Kriegshäuser, O. N. Fanini, S. Forgang, R. A. Mollison, L. Yu,
P. K. Gupta, J. M. V. Koelman, and J. van Popta. Increased oil
in place in low resistivity reservoirs from multicomponent induction
log data. In SPWLA 41st Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 2000. Paper A.
[165] K. S. Kunz and J. H. Moran. Some effects of formation anisotropy
on resistivity measurements in boreholes. Geophysics, 23(4):770–794,
1958.
[166] P. Lacour-Gayet. The Groningen effect: Causes and partial remedy.
The Technical Review, 29(1):37–47, 1981.
[167] P. Lacour-Gayet. Method and apparatus for detecting an anomaly in
a resistivity measurement of an earth formation, 1982. U.S. Patent
4,335,353.
[168] J. LaVigne, T. Barber, and T. Bratton. Strange invasion profiles–
what multiarray induction logs can tell us about how oil-base mud
affects the invasion process and wellbore stability. In SPWLA 38th
Annual Logging Symposium Transactions. Society of Professional Well
Log Analysts, 1997. Paper B.
350 BIBLIOGRAPHY

[169] J. Leake and F. Shray. Logging while drilling keeps horizontal well on
small target. Oil and Gas Journal, 89(38):53–58, 1991.

[170] Y. Y. Lin, S. Gianzero, and R. Strickland. Inversion of induction log-


ging data using the least squares technique. In SPWLA 25th Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 1984. Paper AA.

[171] L. R. Lines and S. Treitel. Tutorial: A review of least squares inversion


and its application to geophysical problems. Geophysical Prospecting,
32(2):159–186, 1984.

[172] Q. H. Liu. Non-linear inversion of electrode-type resistivity mea-


surements. IEEE Transactions on Geoscience and Remote Sensing,
32(3):499–507, 1994.

[173] Q. H. Liu, B. Anderson, and W. C. Chew. Modeling low-frequency


electrode-type resistivity tools in invaded thin beds. IEEE Transac-
tions on Geoscience and Remote Sensing, 32(3):494–498, 1994.

[174] Q. H. Liu and W. C. Chew. Diffraction of non-axisymmetric waves in


cylindrically layered media by horizontal discontinuities. Radio Sci-
ence, 27(5):569–581, 1992.

[175] S. Locke. Wedge problems for induction sondes. Schlumberger-Doll


Research internal report, March 1969. Project 21-51-19.

[176] S. Locke. Direct simulation of the resistor network. Schlumberger-Doll


Research internal report, April 1980. Project MEL-004.

[177] J. R. Lovell. Finite Element Methods in Resistivity Logging. PhD


thesis, Delft University, 1993.

[178] J. R. Lovell and W. C. Chew. Effect of tool eccentricity on some


electrical well logging tools. IEEE Transactions on Geoscience and
Remote Sensing, 28(1):127–136, 1990.

[179] J. R. Lovell, R. A. Rosthal, C. L. Arceneaux, R. A. Young, and


L. Buffington. Structural interpretation of resistivity-at-the-bit images.
In SPWLA 36th Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1995. Paper TT.
BIBLIOGRAPHY 351

[180] M. G. Lüling, R. A. Rosthal, and F. Shray. Processing and modeling


2-MHz resistivity tools in dipping, laminated anisotropic formations.
In SPWLA 35th Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1994. Paper QQ.
[181] R. Maillet and H. G. Doll. Sur un théorème relatif aux mi-
lieux électriquement anisotropes et ses applications a la prospection
électrique en courant continu. Ergänzungshefte für angewandte Geo-
physik, 3:109–124, 1932.
[182] M. Margulies. Modeling of resistivity sondes: A numerical approach.
Schlumberger-Doll Research internal report, March 1980. Project
MEL-004.
[183] M. Martin. Note sur la Résistivité. Schlumberger, Paris, France, 1948.
[184] H. M. Maurer and J. Hunziker. Early results of through casing resis-
tivity field tests. In SPWLA 41st Annual Logging Symposium Trans-
actions. Society of Professional Well Log Analysts, 2000. Paper DD.
[185] J. H. McClellan, T. W. Park, and L. Rabiner. A computer program for
designing optimum FIR linear phase digital filters,. IEEE Transactions
on Audio and Electroacoustics, AU-21(6):505–526, 1973.
[186] W. H. Meyer, L. W. Thompson, M. M. Wisler, and J. Q. Wu. A
new slimhole multiple propagation resistivity tool. In SPWLA 35th
Annual Logging Symposium Transactions. Society of Professional Well
Log Analysts, 1994. Paper NN.
[187] J. W. Miles and G. N. Minerbo. AIT borehole correction implemen-
tation. Schlumberger Houston Engineering internal report, January
1988. Report AIT-21-40.
[188] G. N. Minerbo and J. Miles. Borehole correction system for an array
induction well logging apparatus, 1991. U.S. Patent 5,041,975.
[189] D. C. Moore, editor. Productive Low Resistivity Well Logs of the Off-
shore Gulf of Mexico. New Orleans and Houston Geological Societies,
New Orleans, LA, 1993.
[190] J. Moran and J. Timmons. The mathematical theory of ring electrode
laterolog response in thick beds. Schlumberger-Doll Research internal
report, September 1957. Project RR-602.
352 BIBLIOGRAPHY

[191] J. H. Moran. Induction method and apparatus for investigating earth


formations utilizing two quadrature phase components of a detected
signal, 1964. U.S. Patent 3,147,429.

[192] J. H. Moran. Induction logging–geometrical factors with skin effect.


The Log Analyst, 23(6):4–10, 1982.

[193] J. H. Moran and S. Gianzero. Effects of formation anisotropy on re-


sistivity logging measurements. Geophysics, 44(7):1266–1286, 1979.

[194] J. H. Moran and K. S. Kunz. Basic theory of induction logging and


application to study of two-coil sondes. Geophysics, 27(6):829–858,
1962.

[195] P. M. Morse and H. Feshbach. Methods of Theoretical Physics. Mc-


Graw Hill Book Co., New York, 1953.

[196] S. Moskow, V. Druskin, T. Habashy, P. Lee, and S. Davydycheva. A


finite difference scheme for elliptic equations with rough coefficients
using a Cartesian grid nonconforming to interfaces. SIAM Journal on
Numerical Analysis, 36(2):442–464, 1999.

[197] G. Mowat, J. Lovell, M. Miyari, and K. Tesuka. Modeling an ultralong


electrical device: Application in a fractured geothermal reservoir. In
SEG 60th Annual Meeting Transactions. Society of Exploration Geo-
physicists, 1990. Paper EM3.4.

[198] V. N. Nikitina. The general solution of an axially symmetrical problem


in induction logging theory. Izvestia Academia Nauka SSSR, Series
Geophysics, (4):607–616, 1960.

[199] J. F. Nye. Physical properties of crystals; their representation by ten-


sors and matrices. Oxford at the Clarendon Press, Oxford, 1976.

[200] J. Oberkircher, G. Steinberger, and B. Robbins. Applications for a


multiple depth of investigation MWD resistivity device. In SPWLA
34th Annual Logging Symposium Transactions. Society of Professional
Well Log Analysts, 1993. Paper OO.

[201] D. Omeragic. Sensitivity study for a system of transverse induction


coils. Schlumberger Sugar Land Product Center internal report, Febru-
ary 1998.
BIBLIOGRAPHY 353

[202] M. L. Oristaglio and G. W. Hohmann. Diffusion of electromagnetic


fields in a two-dimensional earth: A finite difference approach. Geo-
physics, 49(7):870–894, 1984.

[203] A. R. Plyler, editor. Engineer Training Manual–Basic Resistivity.


Schlumberger, Houston, TX, 1979.

[204] M. Pudensi and L. Ferreira. Method to calculate the reflection and


transmission of guided waves. Journal of the Optical Society of Amer-
ica, 72(1):126–130, 1982.

[205] A. Pupon. Induction logging, 1957. U.S. Patent 2,790,138.

[206] L. Rabiner and B. Gold. Theory and Application of Digital Signal


Processing, chapter 3. Prentice Hall, Englewood Cliffs, NJ, 1975. “The
theory and approximation of finite-duration impulse response digital
filters”.

[207] T. S. Ramakrishnan and D. J. Wilkinson. Water cut and fractional


flow logs from array induction measurements. In Proceedings 1996 SPE
Annual Technical Conference and Exhibition. Society of Petroleum En-
gineers, 1996. Paper SPE 36503.

[208] R. C. Ransom. Practical Formation Evaluation. John Wiley and Sons,


New York, 1995.

[209] P. F. Rodney, S. G. Mack, M. S. Bittar, and R. P. Bartel. An MWD


multiple depth of investigation electromagnetic wave resistivity sensor.
In SPWLA 32nd Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1991. Paper D.

[210] P. F. Rodney, M. M. Wisler, L. W. Thompson, and R. A. Meador.


The Electromagnetic Wave Resistivity MWD tool. In Proceedings
1983 SPE Annual Technical Conference and Exhibition. Society of
Petroleum Engineers, 1983. Paper SPE 12167.

[211] R. A. Rosthal. Method and apparatus for determining horizontal con-


ductivity and vertical conductivity of earth formations, 1994. U.S.
Patent 5,329,448.

[212] R. A. Rosthal, R. A. Young, J. R. Lovell, L. Buffington, and C. L.


Arceneaux. Formation evaluation and geological interpretation from
354 BIBLIOGRAPHY

the resistivity-at-the-bit tool. In Proceedings 1995 SPE Annual Tech-


nical Conference and Exhibition. Society of Petroleum Engineers, 1995.
Paper SPE 30550.

[213] D. M. Rubin. Cross-bedding, Bedforms, and Paleocurrents. Society of


Economic Paleontologists and Mineralogists, Tulsa, OK, 1987.

[214] R. J. Runge and D. G. Hill. The role of anisotropy in ULSEL. In SP-


WLA 12th Annual Logging Symposium Transactions. Society of Pro-
fessional Well Log Analysts, 1971. Paper J.

[215] D. Rust and B. Anderson. An intimate look at induction logging. The


Technical Review, 23(2):30–54, 1975.

[216] K. A. Safinya, T. Habashy, C. Randall, B. Clark, and A. Perez-Falcon.


Experimental and theoretical study of the Electromagnetic Propaga-
tion Tool in (radially) layered and homogeneous media. In Proceed-
ings 1985 SPE Annual Technical Conference and Exhibition. Society
of Petroleum Engineers, 1985. Paper SPE 14188.

[217] K. A. Safinya, P. LeLan, M. Villegas, and P. S. Cheung. Improved


formation imaging with extended microelectrical arrays. In Proceed-
ings 1991 SPE Annual Technical Conference and Exhibition. Society
of Petroleum Engineers, 1991. Paper SPE 22726.

[218] J. R. Samworth, M. C. Spenser, H. K. Patel, and N. R. Atack. The


array induction tool advances slim hole logging technology. In SP-
WLA 16th European Formation Evaluation Symposium Transactions.
Society of Professional Well Log Analysts, 1994. Paper Y.

[219] R. Schaefer, T. D. Barber, and C. Dutcher. Phasor processing of


induction logs including shoulder and skin effect correction, 1984. U.S.
Patent 4,471,436.

[220] Schlumberger, Houston, TX. Log Interpretetion Princi-


ples/Applications, 1987. Document SMP-7017.

[221] Schlumberger, Houston, TX. Phasor Induction Tool, 1989. Document


SMP-9060.

[222] Schlumberger, Houston, TX. Schlumberger Log Interpretetion Charts,


1991. Document SMP-7006.
BIBLIOGRAPHY 355

[223] C. Schlumberger. Étude sur la prospection électrique du sous-sol. Gau-


thier Villars, Paris, France, 1920.

[224] Schlumberger Limited, New York, NY. Schlumberger Borehole Cor-


rection Charts, 1971.

[225] Schlumberger Limited, New York. Dipmeter Interpretation, 1986. Doc-


ument SMP-7002.

[226] Schlumberger Well Services–Engineering, Houston, TX. Field Log In-


terpretation Computing Center Greenbook, 1991.

[227] Schlumberger Well Surveying Corporation, Houston, TX. Interpreta-


tion Handbook for Resistivity Logs, 1950. Document Number 4.

[228] Schlumberger Well Surveying Corporation, Houston, TX. Schlum-


berger Log Interpretetion Charts, 1962.

[229] Schlumberger Well Surveying Corporation, Houston, TX. Schlum-


berger Log Interpretetion Charts, 1972.

[230] Schlumberger Well Surveying Corporation, Houston, TX. Schlum-


berger Log Interpretetion Charts, 1979.

[231] N. Schuster. Method and apparatus for investigating earth formations


by emitting survey and auxiliary currents from the same electrode,
1973. U.S. Patent 3,760,260.

[232] N. Schuster. Deep investigating induction logging with mirror image


coil arrays, 1984. U.S. Patent 4,472,684.

[233] N. A. Schuster, J. D. Badon, and E. R. Robbins. Application of the


ISF/sonic combination tool to Gulf Coast formations. In Transactions
21st Annual Fall Meeting, pages 177–186. Gulf Coast Association of
Geological Societies, 1971.

[234] O. Serra. Formation MicroScanner Image Interpretation. Schlum-


berger, Houston, TX, 1991. Document SMP-7028.

[235] A. Sholberg. Apparatus for determining the resistivity of a subsurface


earth formation at different lateral distances from a borehole wall,
1973. U.S. Patent 3,772,589.
356 BIBLIOGRAPHY

[236] J. W. Smits, D. Benimeli, I. Dubourg, O. Faivre, D. Hoyle, V. Touril-


lon, J. C. Trouiller, and B. I. Anderson. High resolution from a new
laterolog with azimuthal imaging. In Proceedings 1995 SPE Annual
Technical Conference and Exhibition. Society of Petroleum Engineers,
1995. Paper SPE 30584.

[237] J. W. Smits, I. Dubourg, M. G. Lüling, G. N. Minerbo, J. M. V. A.


Koelman, L. B. J. Hoffman, A. T. Lomas, R. K. van der Oosten, M. J.
Schiet, and R. N. Dennis. Improved resistivity interpretation using
a new array laterolog tool and associated inversion processing. In
Proceedings 1998 SPE Annual Technical Conference and Exhibition.
Society of Petroleum Engineers, 1998. Paper SPE 49328.

[238] Society of Professional Well Log Analysts, Houston, TX. The Art of
Ancient Log Analysis, 1979.

[239] A. Sommerfeld. Partial Differential Equations in Physics. Academic


Press, New York, 1949.

[240] B. R. Spies and T. M. Habashy. Sensitivity analysis of crosswell elec-


tromagnetics. Geophysics, 60(3):834–845, 1995.

[241] D. H. Staelin, A. W. Morgenthaler, and J. A. Kong. Electromagnetic


Waves. Prentice Hall, Englewood Cliffs, NJ, 1994.

[242] G. Strang and G. J. Fix. An Analysis of the Finite Element Method.


Prentice Hall, Englewood Cliffs, NJ, 1973.

[243] J. A. Stratton. Electromagnetic Theory. McGraw Hill Book Co., New


York, 1941.

[244] J. Suau. Spherical focusing method and apparatus for determining


the thickness of a zone in an earth formation traversed by a borehole,
1978. U.S. Patent 4,087,740.

[245] J. Suau, P. Grimaldi, A. Pupon, and P. Souhaite. The Dual Laterolog–


Rxo tool. In Proceedings 1972 SPE Annual Technical Conference and
Exhibition. Society of Petroleum Engineers, 1972. Paper SPE 4018.

[246] J. R. Tabanou, B. Anderson, S. Bruce, T. Borneman, K. Hodenfield,


and P. Wu. Which resistivity should be used to evaluate thinly bed-
ded reservoirs in high-angle wells? In SPWLA 40th Annual Logging
BIBLIOGRAPHY 357

Symposium Transactions. Society of Professional Well Log Analysts,


1999. Paper E.

[247] J. R. Tabanou and B. I. Anderson. Remote sensing of subsurface strata


with low frequency electrode devices. Unpublished report, 1990.

[248] L. Tabarovsky and M. Rabinovich. High-speed 2D inversion of in-


duction logging data. In SPWLA 37th Annual Logging Symposium
Transactions. Society of Professional Well Log Analysts, 1996. Pa-
per P.

[249] D. Tanguy. Induction well logging, 1962. U.S. Patent 3,067,383.

[250] D. A. Tanguy and W. A. Zoeller. Applications of measurements while


drilling. In Proceedings 1981 SPE Annual Technical Conference and
Exhibition. Society of Petroleum Engineers, 1981. Paper SPE 10324.

[251] S. G. Thadani and H. E. Hall. Propagated geometric factors in induc-


tion logging. In SPWLA 22nd Annual Logging Symposium Transac-
tions. Society of Professional Well Log Analysts, 1981. Paper WW.

[252] J. C. Tingey, R. J. Nelson, and K. E. Newsham. Comprehensive anal-


ysis of Russian petrophysical measurements. In SPWLA 36th Annual
Logging Symposium Transactions. Society of Professional Well Log An-
alysts, 1995. Paper S.

[253] J. Tittman. Geophysical Well Logging. Academic Press, New York,


1986.

[254] J. Tittman. Formation anisotropy: Reckoning with its effects. Schlum-


berger Oilfield Review, 2(1):16–23, 1990.

[255] C. Torres-Verdin and T. Habashy. Rapid numerical simulation of ax-


isymmetric single well induction data with the extended Born approx-
imation. Schlumberger-Doll Research internal report, November 1995.
Report EMG-002-95-22.

[256] G. H. Towle, W. W. Whitman, and J. H. Kim. Electric log modeling


with a finite difference method. The Log Analyst, 29(3):184–195, 1988.

[257] L. Tsang, A. K. Chan, and S. Gianzero. Solution of the fundamen-


tal problem in resistivity logging with a hybrid method. Geophysics,
49(10):1596–1604, 1984.
358 BIBLIOGRAPHY

[258] W. B. Vail. Method and apparatus for measurement of resistivity of


geological formations from within cased boreholes, 1989. U.S. Patent
4,820,989.

[259] P. M. Vallinga, J. R. Harris, and M. A. Yuratich. A multi-electrode


tool, allowing more flexibility in resistivity logging. In 14th European
Formation Evaluation Symposium Transactions. London Chapter of
SPWLA, 1991. Paper E.

[260] P. M. Vallinga and M. A. Yuratich. Accurate assessment of hydrocar-


bon saturations in complex reservoirs from multi-electrode tool resis-
tivity measurements. In SPWLA/CWLS 34th Annual Logging Sym-
posium Transactions. Society of Professional Well Log Analysts, 1993.
Paper E.

[261] P. M. van den Berg and R. E. Kleinman. A contrast source inversion


method. Inverse Problems, 13(6):1607–1620, 1997.

[262] P. M. van den Berg, A. L. van Broekhoven, and A. Abubakar. Extended


contrast source inversion. Inverse Problems, 15(5):1325–1344, 1999.

[263] M. van der Horst, V. Druskin, and L. Knizhnerman. Modeling the


response of induction logging tools in 3D geometries with the spectral
Lanczos decomposition method. In Proceedings of the International
Symposium on Three-Dimensional Electromagnetics, pages 491–498,
1995.

[264] T. Walsgrove, B. Anderson, and T. Barber. Case study of an aniso-


tropic crossbedded reservoir–improving predictions of reservoir proper-
ties through modeling the resistivity. In SPWLA 40th Annual Logging
Symposium Transactions. Society of Professional Well Log Analysts,
1999. Paper V.

[265] H. M. Wang, L. C. Shen, and G. J. Zhang. Dual Laterolog response in


3D environments. In SPWLA 39th Annual Logging Symposium Trans-
actions. Society of Professional Well Log Analysts, 1998. Paper X.

[266] T. Wang and G. W. Hohmann. A finite difference, time-domain so-


lution for three-dimensional electromagnetic modeling. Geophysics,
58(6):797–809, 1993.
BIBLIOGRAPHY 359

[267] W. W. Whitman, J. H. Schoen, G. H. Towle, and J. H. Kim. An auto-


matic inversion of normal resistivity logs. The Log Analyst, 31(1):10–
19, 1990.

[268] W. W. Whitman, G. H. Towle, and J. H. Kim. Inversion of normal


and lateral well logs with borehole compensation. The Log Analyst,
30(1):1–9, 1989.

[269] N. A. Wiltgen. The essentials of basic Russian well logs and analysis
techniques. In SPWLA 35th Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1994. Paper II.

[270] R. Woodhouse. The laterolog Groningen phantom can cost you money.
In SPWLA 19th Annual Logging Symposium Transactions. Society of
Professional Well Log Analysts, 1978. Paper R.

[271] F. Yang and S. H. Ward. Inversion of borehole normal resistivity logs.


Geophysics, 49(9):1541–1548, 1984.

[272] K. S. Yee. Numerical solution of initial boundary value problems in-


volving Maxwell’s equations in isotropic media. IEEE Transactions on
Antennas and Propagation, AP-14(3):302–307, 1966.

[273] M. A. Yuratich and W. J. Meger. The application of finite difference


methods to normal resistivity logs. In SPWLA 25th Annual Logging
Symposium Transactions. Society of Professional Well Log Analysts,
1984. Paper V.

[274] P. Zamansky. Simulation des laterologs par la méthode des éléments


finis. Schlumberger Paris Engineering internal report, March 1980.
Project 21-44-00.

[275] G. J. Zhang. High-order geometrical factors of induction logging. Acta


Geophysica Sinica, 25(4), 1982.

[276] G. J. Zhang and L. C. Shen. Response of a normal resistivity tool in


a borehole crossing a bed boundary. Geophysics, 49(2):142–149, 1984.
360 BIBLIOGRAPHY
Summary
This thesis specifically addresses new research in the parametric inver-
sion of triaxial induction logging data to obtain the true formation resis-
tivity values in anisotropic layered media. It also includes an overview of
the response characteristics of some of the most commonly used resistivity
tools. The overview is provided for two reasons: (1) a knowledge of tool
physics is helpful for obtaining accurate inversion results, and (2) there are
no other references that bridge the gap between computational physics and
log interpretation. In the overview, the response characteristics of all major
resistivity logging tools are systematically compared for the first time by
analyzing their computed logs in the same benchmark formation. The most
computationally efficient modeling methods for resistivity tools are summa-
rized and references are provided so that the reader can use this thesis as
a guide for writing his own modeling codes. The thesis takes a historical
perspective, showing how specific problems have provided the motivation
for improvements in tool design, mathematical modeling and interpretation
methodology.
The electrical resistivity is one of the most important rock parameters
used to quantify the amount of hydrocarbons present in a reservoir. In elec-
trical logging, current sources are lowered into a borehole, and the apparent
resistivity of a rock formation is obtained from scaled voltages recorded at
regular depth intervals. In simplest terms, high resistivity indicates the pres-
ence of oil or gas in rock pores, since hydrocarbons are insulators. On the
other hand, low resistivity indicates the presence of water, the other fluid
that may be present.
Environmental effects complicate the determination of the true resistiv-
ity of a specific volume of a formation. Logging tools measure the over-all
bulk average resistivity of a relatively large volume of the formation, and
perturbations caused by regions adjacent to beds of interest can have con-
siderable effect. Some of the most significant effects are caused by: (1) the
borehole filled with drilling mud, (2) zones encircling the borehole invaded
by the drilling mud, and (3) adjacent beds of differing resistivity.
Because it is necessary to evaluate these environmental effects, and be-
cause resistivity tool response can be highly nonlinear, mathematical model-
ing has always been intimately associated with resistivity log interpretation.
Starting in the 1930’s, an elementary form of “inversion” was performed
by entering log data in correction charts; this was done manually at first
362

and later by using computer algorithms. Correction charts were generated


using forward modeling codes that computed tool response in given 1D con-
figurations by solving Maxwell’s equations. 2D problems were inverted by
applying the 1D charts sequentially.
Significant improvements in computing capabilities occurred in the 1980’s
which allowed rigorous 2D (axisymmetric) modeling and inversion software
to be developed. However, inversion methods enjoyed limited success outside
of research environments because they were slow and plagued by nonunique-
ness. In the mid-1980’s, iterative forward modeling was proposed by the
author of this thesis as a stopgap solution to these inversion problems. Iter-
ative forward modeling is still used more often than inversion because, with
modeling, it is easier to constrain formation parameters by introducing local
knowledge and non-resistivity information.
Tools composed of multiple arrays of antennas were introduced in the
1990’s. Array tools helped to address the nonuniqueness problem in verti-
cal wells, but by that time the growing use of horizontal drilling required
even slower 3D modeling and inversion. The axisymmetric antennas of re-
sistivity tools which were designed for vertical wells did not provide enough
information to perform accurate 3D inversion.
The growth of horizontal drilling also drew attention to the surprisingly
large effect of resistivity anisotropy (variation of resistivity with direction).
In vertical wells, anisotropy was regarded as a secondary effect primarily
because conventional resistivity tools are only sensitive to the horizontal
resistivity component. In horizontal wells, logs sometimes read 3 to 4 times
higher than logs in vertical exploration wells in the same bed. Modeling
showed that these differences were caused by caused by anisotropy. Simple
inversions algorithms based on tool response in homogeneous anisotropic
media were implemented to solve for the horizontal and vertical resistivity
values from log readings. Although these algorithms worked well in the
center of thick beds, they gave inaccurate solutions near bed boundaries and
in thin beds.
Anisotropy is not uncommon in sedimentary strata. Many solid parti-
cles have flat or elongated shapes that are usually oriented parallel to the
plane of deposition. This results in a pore structure that allows electric cur-
rent to flow more easily parallel to the bedding plane than perpendicular
to it. Anisotropy also depends on scale. Formations consisting of a series
of isotropic beds of different lithology, such as laminated sands and shales,
363

also behave anisotropically if a logging tool is significantly longer than the


bed thicknesses. In vertical wells, the presence of hydrocarbons is masked
in these laminated zones because resistivity tools are only sensitive to the
horizontal resistivity which is dominated by the less resistive shales. A mea-
surement of the vertical resistivity which is dominated by the more resistive
hydrocarbon-bearing sands would be a better indicator of the presence of oil
or gas.
Experience has shown that 2-MHz induction tools used in logging while
drilling are moderately sensitive to anisotropy in horizontal wells. Because
of this sensitivity, the 2-MHz tool was chosen to test the parametric inver-
sion algorithm in layered anisotropic media. In the parametric inversion,
the objective is to solve for the horizontal and vertical resistivities within
each bed using the resistivity log data. The inversion algorithm is an itera-
tive approach based on the Gauss-Newton method that employs a quadratic
model of a cost function. The method is based on constrained minimization
where upper and lower bounds are imposed on the inverted parameters. The
forward model used in the inversion is a frequency domain (AC) code which
computes the response of arbitrarily oriented magnetic dipoles in anisotropic
layered media.
Simulated 2-MHz logs were computed using the forward model and then
inverted, both with and without noise. It was assumed that bed boundary
locations could be obtained from inflection points or an independent higher
resolution log, and the dip and strike angles could be obtained from imaging
or dipmeter logs. The results of the inversion were accurate in many cases,
but there were two major problem areas: (1) the vertical (z-directed) mag-
netic dipole antennas of the 2-MHz tool were insensitive to anisotropy when
the deviation angle was less than 40 degrees, and (2) an error of more than
two inches in the bed boundary locations caused large errors in the inversion
results.
To address these problems, the inversion was next applied to logs gener-
ated by a triaxial induction tool composed of mutually orthogonal x, y, and
z-directed dipoles. The x and y-directed dipoles provide good sensitivity
to the vertical resistivity even in vertical wells. A frequency in the tens of
kilohertz range was chosen in order to obtain a deep radial penetration of
the signal. For the triaxial inversion, the bed boundary locations and dip
and strike angles were not fixed, but were included in the inversion solution
(all formation parameters were assumed to be unknown). Inversion results
showed that the triaxial tool has sufficient sensitivity to accurately invert for
364

the horizontal and vertical resistivity values in all beds, the bed boundary
locations and the dip and strike angles. Future work is planned to extend the
inversion to fully 3D anisotropic geometries, including borehole and invasion
effects.
Samenvatting

In dit proefschrift wordt speciale aandacht geschonken aan nieuw on-


derzoek op het gebied van de parametrische inversie van data verkregen via
drie-assige inductielogging met het doel het werkelijke soortelijke weerstands-
profiel in anisotrope, gelaagde media te reconstrueren. Daarnaast is een
overzicht opgenomen van responsie-eigenschappen van enkele van de meest
gebruikte gelijkstroomboorgatinstrumenten. Dit overzicht is om twee re-
denen samengesteld: (1) kennis van de fysica van boorgatinstrumenten is
nuttig bij het verkrijgen van nauwkeurige inversieresultaten, (2) er is op dit
moment geen literatuur voorhanden van andere auteurs die een brug slaan
tussen toegepaste fysica en loginterpretatie. In het overzicht zijn de respon-
siekarakteristieken van alle belangrijke soortelijke weerstandsinstrumenten
voor het eerst systematisch vergeleken door hun logs in hetzelfde referen-
tiekader te analyseren. De rekenkundig meest efficiënte modelleringsme-
thoden voor soortelijke weerstandsinstrumenten worden op een rij gezet met
als doel dat de lezer dit proefschrift kan gebruiken als handleiding bij het
schrijven van eigen modelleringsprogramma’s. Het proefschrift is geschreven
vanuit een pragmatisch-historisch perspectief met het oogmerk te demonstre-
ren hoe specifieke problemen geleid hebben tot verbetering van zowel het
ontwerp van instrumenten als het wiskundig modelleren en de methodologie
van interpreteren.
De elektrische soortelijke weerstand is een van de belangrijkste gesteen-
teparameters en wordt gebruikt om de hoeveelheid koolwaterstoffen in een
reservoir te bepalen. Bij het elektrisch loggen worden elektrische stroom-
bronnen in een boorgat neergelaten en de aanwezige soortelijke weerstand
van een gesteenteformatie wordt verkregen uit de gemeten elektrische span-
ning op regelmatige diepte-intervallen. Simpelweg, hoge soortelijke weer-
stand wijst op de aanwezigheid van olie of gas in gesteenteporiën, omdat kool-
waterstoffen niet geleidend zijn. Aan de andere kant duidt lage soortelijke
weerstand op de aanwezigheid van water.
Omgevingseffecten maken de bepaling van de werkelijke soortelijke weer-
stand van een formatie met een specifieke omvang moeilijker. Loggingsin-
strumenten meten de gemiddelde soortelijke weerstand van een formatie
over een relatief groot gebied als geheel en verstoringen veroorzaakt door
aangrenzende lagen kunnen aanzienlijke gevolgen hebben. Enkele van de
meest voorkomende effecten worden veroorzaakt doordat: (1) het boorgat
gevuld is met boorvloeistof, (2) gebieden rond het boorgat doordrongen zijn
366

van boorvloeistof, (3) aangrenzende lagen afwijkende soortelijke weerstand


hebben.
Omdat het noodzakelijk is deze omgevingseffecten te evalueren en om-
dat de elektrische instrumentresponsie zeer niet-lineair kan zijn, is wiskundig
modelleren altijd noodzakelijk geweest bij het interpreteren van weerstands-
loggen. Vanaf de jaren 1930 tot 1980 werd een elementaire vorm van “in-
versie” gebruikt door het invoeren van logdata in correctiekaarten; dit werd
aanvankelijk handmatig gedaan, later met behulp van computeralgoritmen.
Correctiekaarten werden gemaakt aan de hand van computerprogramma’s
die op analytische wijze de vergelijkingen van Maxwell oplosten om de instru-
mentresponsie voorwaarts te modelleren in 1D configuraties. 2D problemen
werden geinverteerd met behulp van sequentieel gebruik van 1D informatie.
In de jaren tachtig traden belangrijke verbeteringen in rekenkundige
mogelijkheden op, die echte 2D (rotatiesymmetrische) modellering en in-
versiesoftware mogelijk maakten. Echter, de inversiemethoden waren een
beperkt commercieel succes in boorgatlogging, omdat ze traag waren en
niet altijd eenduidige resultaten opleverden. In het midden van de jaren
tachtig werd door de auteur van dit proefschrift de iteratieve voorwaartse
modellering met een grote verscheidenheid aan parameters voorgesteld ten-
einde bovenstaande problemen op te lossen. Iteratieve voorwaartse modelle-
ringsmethoden worden nog altijd vaker gebruikt dan andere inversiemetho-
den, omdat het gemakkelijker is om formatieparameters te begrenzen door
het introduceren van plaatselijke voorkennis en informatie over soortelijke
weerstandsverdelingen.
In de jaren negentig werden instrumenten geı̈ntroduceerd die samenge-
steld waren uit meervoudige arrays van antennes. Instrumenten in array-
samenstelling hielpen om het probleem van eenduidigheid in verticale boor-
putten aan te pakken, maar tegelijkertijd leidde een toename in het gebruik
van horizontaal boren tot tragere 3D modellering en inversie. De rotatiesym-
metrische antennes van weerstandsinstrumenten voor verticale boorputten
gaven onvoldoende informatie om nauwkeurige 3D inversie uit te voeren.
Door de toename van horizontale boortechnieken werd de aandacht geves-
tigd op het verrassend grote effect van soortelijke weerstandsanisotropie. In
verticale boorputten werd anisotropie gezien als een secundair effect, vooral
omdat conventionele soortelijke weerstandsinstrumenten alleen gevoelig zijn
voor de horizontale weerstandscomponent. In horizontale boorgaten gaven
de logs soms drie- tot viermaal hogere waarden aan dan verticale logs in
367

dezelfde aardlaag. Modellering toonde aan dat deze verschillen toegeschreven


kunnen worden aan anisotropie. Modellering voor homogene, isotrope me-
dia liet zien dat instrumenten steeds gevoeliger werden voor de verticale
soortelijke weerstandscomponent naarmate de as van het instrument meer
van de boorgatas afweek. Alhoewel deze inversie-algoritmen goed werkten in
het midden van dikke lagen, gaven ze onnauwkeurige antwoorden bij grens-
gebieden van lagen, alsmede bij dunne lagen.
Anisotropie komt vaak voor in sedimentaire aardlagen. Veel vaste deeltjes
zijn plat of hebben een langwerpige vorm, gewoonlijk georiënteerd evenwijdig
aan de laag waarin ze zich bevinden. Dit resulteert in een poreuze structuur
die gemakkelijker een elektrische stroom geleidt evenwijdig aan het opper-
vlak van de laag dan in de richting loodrecht erop. Anisotropie hangt ook af
van de schaaleffecten. Formaties die bestaan uit een reeks van isotrope la-
gen met verschillende lithologie, zoals gelamineerd zand en schalie, gedragen
zich ook anisotroop als het loggingsinstrument aanzienlijk langer is dan de
dikte van de laag. In verticale putten wordt in deze gelamineerde gebieden
de aanwezigheid van koolwaterstoffen gemaskeerd omdat soortelijke weer-
standsinstrumenten slechts gevoelig zijn voor de horizontale weerstand, die
gedomineerd wordt door de lagere weerstand van samengeperste gelaagde
klei. Meting van de verticale soortelijke weerstand, gedomineerd door kool-
waterstof bevattend zand met een hogere weerstand, zou een betere aanwij-
zing zijn voor de aanwezigheid van olie of gas.
Ervaring heeft aangetoond dat 2MHz inductieinstrumenten, die gebruikt
worden bij logging tijdens het boren, in horizontale putten licht gevoelig zijn
voor anisotropie. Vanwege deze gevoeligheid werd het 2MHz instrument
gekozen om de parametrische inversie te testen in gelaagde, anisotrope me-
dia. Het doel van een parametrische inversie is het bepalen van de horizontale
en verticale soortelijke weerstandsverdelingen binnen elke laag, gebruikma-
kend van de weerstandslogdata. Het inversiealgoritme is een iteratieve, op
Gauss-Newton gebaseerde methode met een kwadratische kostenfunctie. De
methode is gebaseerd op ingeperkte minimalisering waarbij boven- en onder-
grenzen gesteld worden aan de te inverteren parameters. Het voorwaartse
model dat voor de inversie wordt gebruikt is een frequentiedomein (AC) code
die de responsie berekent van willekeurig geplaatste magnetische dipolen in
anisotrope, gelaagde media.
Vervolgens werden er gesimuleerde 2MHz logs met het voorwaartse model
berekend en daarna geinverteerd, zowel met als zonder toegeroegde ruis. De
eerste iteraties voor horizontale en verticale soortelijke weerstanden werden
368

verkregen uit metingen in het midden van de lagen. De begrenzingslocaties


van een laag werden verkregen uit buigpunten op de logs. De resultaten
van de inversie waren in de meeste gevallen nauwkeurig, maar er waren
twee problemen: (1) de verticale (langs de boorgatas, de z-as) magnetische
dipoolantennes van het 2MHz instrument waren ongevoelig voor anisotropie
als de afwijkingshoek kleiner was dan 40 graden, (2) een afwijking van meer
dan twee inch in de begrenzingslocatie van de laag geeft grote fouten in de
resultaten van de inversie.
Om deze problemen op te lossen werd de inversie vervolgens toegepast bij
logs gegenereerd door een drie-assig inductieinstrument samengesteld uit on-
derling orthogonale x-, y- en z-gerichte dipolen. De x- en y-gerichte dipolen
hebben een juiste gevoeligheid voor de verticale soortelijke weerstand, zelfs
in verticale boorputten. Teneinde een grote radiële velddoordringing van het
signaal te verkrijgen werd gekozen voor de werkfrequentie van enkele tien-
tallen kHz. In de drie-assige inversie werden de begrenzingslocaties van de
lagen en de hellings- en inclinatiehoeken niet vastgelegd, maar opgenomen
in de inversieoplossing, d.w.z. alle formatieparameters werden als onbekend
aangenomen. Inversieresultaten toonden aan dat het drie-assige instrument
voldoende gevoeligheid had om nauwkeurig te inverteren voor de horizon-
tale en verticale soortelijke weerstandsverdelingen in alle lagen, de begren-
zingslocaties en de hellings- en inclinatiehoeken. In de toekomst zal on-
derzoek worden gedaan om de inversie uit te breiden tot een volledige 3D
anisotrope geometrie, waarin boorgat en effecten van vloeistofindringing wor-
den meegenomen.
About the Author
Barbara Anderson was born in Danbury, Connecticut, USA. She earned
a B.S. degree from Western Connecticut State University in 1963, ranking
first in her graduating class. After teaching mathematics at a local high
school and taking graduate courses in mathematics at Fairfield University,
she joined Schlumberger-Doll Research (SDR) in Ridgefield, Connecticut in
1966 as a numerical analyst–programmer. She is currently a Senior Research
Scientist in the Modeling and Inversion Applications Program at SDR.
Since joining Schlumberger, most of Barbara’s work has been in the area
of software development for modeling resistivity tool response. Her soft-
ware has been used in the design and interpretation of tools which provide
Schlumberger with a large percentage of their revenue. These tools include
the dual laterolog, dual induction, Phasor induction, array induction, spher-
ically focused log, electromagnetic propagation tools and 2-MHz resistivity
tools.
One of Barbara’s ongoing goals has been to minimize uncertainty in re-
sistivity log interpretation by integrating forward modeling directly into the
interpretation process. For many years, log interpretation was performed
using correction charts and simple algroithms because full-blown computed
inversion was too slow and often gave nonunique results. As a stopgap so-
lution for these inversion difficulties, Barbara wrote a paper in 1986 demon-
strating how she used iterative forward modeling to solve several previously
intractable 2D interpretation problems [3]. (She was motivated to write the
paper after discovering that it became easier to win hand-waving arguments
with interpretation gurus when she enlisted James Clerk Maxwell on her side
when explaining tool physics.) This paper led to an explosion of modeling
applications by others in the areas of geosteering, basic log interpretation,
and fast forward modeling algorithms. She was subsequently involved in the
development of the first commercial resistivity modeling package, ELMOD.
Her current research interests are 3D modeling and inversion, and anisotropy.
Barbara has authored or co-authored over 30 internal reports and over
60 technical papers in referreed journals and technical meeting transactions.
She co-authored the 1981 Society of Professional Well Log Analysts (SP-
WLA) Symposium Best Paper and the 1994 Society of Petroleum Engineers
(SPE) Best Paper in the journal SPE Formation Evaluation. She also re-
ceived the 1987, 1988 and 1999 Best Paper Awards from the SPWLA journal
The Log Analyst.
370

Barbara has been an active participant in SPWLA, serving on the Board


of Directors as Director-at-Large (1989-90), Vice-President Public Relations
(1990-91), Vice-President Technology (1991-92) and President (1994-95).
She was an SPWLA Distinguished Speaker and received the 1997 SPWLA
Distinguished Technical Achievement Award. She is currently serving on
the Board of Directors of the SPWLA Foundation as Vice-President. She is
also a member of SPE, SEG, IEEE, ACM and SIAM.
Barbara enjoys gardening and needlework in her “spare time”. She also
plays string bass in a local symphony orchestra, and electric bass guitar in
the SDR rhythm and blues band.
Her “Top Ten” most significant papers are listed below.

Most Significant Publications


[1] S. Gianzero and B. Anderson. A new look at skin effect. The Log An-
alyst, 23(1):20–34, 1982. (The first 2D depiction of induction response
functions with skin effect, along with a volumetric interpretation of
induction response. Won the 1981 SPWLA Symposium Best Paper
Award)
[2] W. C. Chew, S. Barone, B. Anderson, and C. Hennessy. Diffraction
of axisymmetric waves in a borehole by bed boundary discontinuities.
Geophysics, 49(10):1586–1595, 1984. (The first demonstration of the
speed and accuracy of the Chew–Barone semi-analytic method for sim-
ulating induction response.)
[3] B. Anderson. The analysis of some unsolved induction interpretation
problems using computer modeling. The Log Analyst, 27(5):60–73,
1986. (The first paper showing how iterative forward modeling could
be used in interpretation as a practical substitute for inversion. Won
the SPWLA 1987 Award for Best Paper in The Log Analyst.)
[4] B. Anderson, K. A. Safinya, and T. Habashy. Effects of dipping beds on
the response of induction tools. SPE Formation Evaluation, 3(1):29–
36, 1988. (Demonstrated the large error in net pay estimation caused
by dip effect in deviated and horizontal wells.)
[5] B. Anderson and T. Barber. Strange induction logs–A catalog of en-
vironmental effects. The Log Analyst, 29(4):229–243, 1988. (Used
modeling to explain 2D effects on logs that had been unsuccessfully
analyzed using 1D plus 1D interpretation. Won the SPWLA 1988
371

Award for Best Paper in The Log Analyst.)


[6] B. Anderson, T. Barber, J. Singer, and T. Broussard. ELMOD–
Putting electromagnetic modeling to work to improve resistivity log
interpretation. In SPWLA 30th Annual Logging Symposium Trans-
actions. Society of Professional Well Log Analysts, 1989. Paper M.
(Introduced the ELMOD modeling package as part of Schlumberger’s
commercial interpretation software.)
[7] B. Anderson, S. Bonner, M. Lüling, and R. Rosthal. Response of 2-
MHz LWD restivity and wireline induction tools in dipping beds and
laminated formations. The Log Analyst, 33(5):461–475, 1992. (The
first paper to document the magnitude of anisotropy effect on 2-MHz
logs in highly deviated wells.)
[8] T. Habashy and B. Anderson. Reconciling differences in depth of in-
vestigation between 2-MHz phase shift and attenuation resistivity mea-
surements. In SPWLA 32nd Annual Logging Symposium Transactions.
Society of Professional Well Log Analysts, 1991. Paper E. (Applied
Born response functions to analyze differences between 2-MHz phase
shift and attenuation measurements.)
[9] D. H. Davies, O. Faivre, M. T. Gounot, B. Seeman, J. C. Trouiller,
D. Benimeli, A. Ferreira, D. J. Pittman, M. Randrianavony, J. W. Smits,
B. I. Anderson, and J. Lovell. Azimuthal resistivity imaging: A new-
generation laterolog. SPE Formation Evaluation, 9(3):165–174, 1994.
(Described the integration of an azimuthal laterolog imaging array with
dual laterolog resistivity measurements. Won the SPE 1994 Award for
Best Paper in SPE Formation Evaluation.)
[10] B. Anderson, T. Barber, V. Druskin, P. Lee, E. Dussan, L. Knizh-
nerman, and S. Davydycheva. The response of multiarray induction
tools in highly dipping formations with invasion and in arbitrary 3D
geometries. The Log Analyst, 40(5):327–344, 1999. (The first paper to
analyze 3D effects on induction tools in deviated and horizontal wells.
Won the SPWLA 1999 Award for Best Paper in The Log Analyst.)
372
Acknowledgments
The author is employed by Schlumberger-Doll Research, Ridgefield, CT,
06877, USA. She thanks Schlumberger for giving her the opportunity to carry
out the research described here, and for granting permission to publish this
thesis.
The author also thanks:
- Dr. Michael Oristaglio for his encouragement to begin this thesis,
- Dr. Tarek Habashy, Dr. Hans Blok and Dr. Jacob Fokkema for their
helpful guidance throughout the preparation of this thesis,
- Dr. Tarek Habashy for his instruction in the “art” of inversion,
- Stan Gianzero, Tom Barber, Martin Lüling and William Batzer for
their constructive comments on the manuscript,
- Tom Barber for providing the template file used to plot tool response
in the benchmark formation,
- Joseph Doduy and Martin Lüling for providing the HRLA inversion
results shown in Figure 3.89,
- Jacques Tabanou for his collaboration on the descriptions of early lat-
erolog tools in Section 3.4 (taken from an unpublished monograph by
Tabanou and Anderson),
- Schlumberger Technology Corporation and Schlumberger Well Services
for allowing the publication of charts and diagrams from internal doc-
uments (credited as they appear),
- The Society of Professional Well Log Analysts (SPWLA), the Society of
Petroleum Engineers (SPE) and the Society of Exploration Geologists
(SEG) for allowing the publication of figures that originally appeared
in the journals and conference proceedings of these societies (credited
as they appear),
- The New Orleans and Houston Geological Societies for Figure 2.9,
- The Society of Economic Paleontologists and Mineralogists for Figure
2.10,
- PennWell Publishing Company for granting permission to publish Fig-
ures 3.1 and 5.1, which were originally published as Figure 1-3 (page 5)
and Figure 2-2 (page 23) in Essentials of Modern Open-hole Log Inter-
pretation, by John T. Dewan, copyright 1983.
374

On the Covers
The front cover depicts superimposed real and imaginary 2D axisym-
metric response function for the ID (Deep Induction) tool, computed by the
author of this thesis in 1985.
The back cover depicts 1D vertical response functions for several pro-
posed ID antenna configurations, sketched by Henri Doll in 1947.
Index

2-MHz inversion results, 305 borehole correction, 265, 266


2D HRLA inversion, 166 borehole effect, 29, 40, 122, 264,
5FF27, 64 327
5FF40, 69 Born response function, 59, 73, 77,
6FF28, 76 90, 92, 205
6FF40, 70 boundary conditions, 47
Broadside array, 113
AIT, 87 bucking coils, 72
analytical methods, 185
bucking currents, 132
ANISBEDS, 289
butterfly charts, 270
anisotropy, 15, 36, 285, 294, 309
annulus, 30
Cartesian coordinate system, 47
anti-squeeze configuration, 69, 157
cartridge, 39
apparent resistivity, 4, 27
Cased Hole Formation Resistivity
ARC5, 105
tool, 182
Archie’s equation, 254
cave effect, 30, 75
ARI, 178
CDR, 101
Array Induction Tool, 87
chartbooks, 9, 263
Array Resistivity Compensated tool,
CHFR, 182
105
coaxial coils, 307
Azimuthal Resistivity Imager, 178
coils, 39
bed boundary effect, 316 compatibility relations, 43, 44
benchmark formation, 69, 70, 80, Compensated Dual Resistivity tool,
84, 92, 99, 104, 106, 111, 101, 208, 293
120, 126, 128, 135, 138, constant power, 146
142, 146, 152, 156, 165 constitutive relations, 43, 45
Bessel functions, 204 contrast source inversion, 274
BKZ (Russian)tools, 127 coplanar coils, 308
block Gaussian elimination, 223 cost function, 296, 298
boosting, 259 crossbedding, 38
borehole, 27 curl operator, 50
376 INDEX

cylindrical boundaries, 202 Formation MicroImager, 176


cylindrical coordinate system, 49 Formation MicroScanner, 176
Fourier-Bessel transform, 199
data transmission rate, 39 frequency domain, 44
de-deconvolution, 262 future plans, 326
deboosting, 262
deconvolution, 258 gamma ray tool, 6
Deep Propagation tool, 108 Gauss-Newton minimization, 300
Delaware effect, 148 Gauss-Newton solution, 281
delta function, 198 geometrical factor theory, 58, 186
density tool, 7 geosteering, 19
departure curves, 9, 263 glossary of computer codes, 247
depth of investigation, 63, 315 Green function, 192, 200, 203
dielectric constant, 108 grid generation, 218
dip correction, 85 Groningen effect, 149
dip effect, 272 ground loops, 57
dipmeter tool, 6, 175 guard electrodes, 137
dipping beds, 34 HALS, 151
distorted Born inversion, 274 Helmholtz equation, 192
DIT-E, 86 High Resolution Azimuthal Laterolog
divergence operator, 49 Sonde, 151
DLT, 143 High Resolution Laterolog Array,
DPT, 108 160
Dual Induction tool, 76 horizontal drilling, 19
Dual Laterolog tool, 143 horizontal magnetic dipole, 308
horizontal wells, 35
eccentricity effect, 264 HRLA, 160
effective length, 72 hybrid methods, 233
Electromagnetic Propagation tool,
112 ID, 76
ELMOD, 11 IES, 67
Endfire array, 113 IM, 76
EPT, 112 imaging logs, 37
imaging tools, 175
finite difference method, 215, 225 induction tools, 54, 55
finite element method, 215, 216 integrated radial geometrical fac-
FMI, 176 tor, 190
FMS, 176 integrated radial response function,
focusing, 64, 65, 131 60
INDEX 377

invasion, 28, 30, 33 mud filtrate, 53


inverse filtering, 82 mudcake, 30
inverse tool, 5, 125 mudcake correction chart, 171
inversion, 14, 22 mudcake effect, 167
iterative forward modeling, 274 mutual inductance, 196

Kaufman’s hybrid method, 244 neutron tool, 7


Newton minimization, 299
laminated formations, 36 nonlinear transformations, 303
lateral tool, 5, 124 nonuniqueness, 14
Laterolog 3, 136 normal tool, 117
Laterolog 7, 130 numerical methods, 214
Laterolog 8, 139
laterolog modeling, 212 OBMI, 180
laterolog tools, 53, 116 Oil-Base MicroImager, 180
oil-base muds, 29
least squares inversion, 274, 279
line search, 301 parametric inversion, 24, 274
LL3, 136 passive focusing, 137
LL7, 130 permeability, 9, 254
LL8, 139 Phasor processing, 80
log interpretation, 9 planar boundaries, 199
logging while drilling, 7, 39 point dipoles, 40, 287
Poisson’s equation, 212
mandrel, 39 porosity, 9, 32, 52, 254
maximum entropy inversion, 15, 94, propagation tools, 100
274 Proximity log, 171
Maxwell’s equations, 41, 43, 44, 50, pseudo-geometrical factor, 79
212 pseudo-laterolog, 145
MCFL, 166
MicroCylindrically Focused Log, 166 R-signal, 57
microinverse, 168 RAB, 179
MicroLaterolog, 170 RAD, 102
Microlog, 168 radial geometrical factor, 189
micronormal, 168 radial response function, 60
microresistivity tools, 166 reactive signal, 57
MicroSFL, 173 real axis integration, 197
mode matching, 234 regularization parameter, 302
monitoring conditions, 132 resistive signal, 57
MSFL, 173 resistivity, 3
378 INDEX

resistivity through casing, 182 transverse magnetic dipole, 21, 308


Resistivity-At-the-Bit tool, 179 triangular functions, 237
resistor network, 10 triaxial induction tool, 286, 307
RPS, 102 triaxial inversion results, 318
Russian induction tools, 95 two-coil sonde, 56

salinity, 32 ULSEL, 18
sands, 38
vertical geometrical factor, 189
saturation calculations, 256
vertical magnetic dipole, 307
semi-analytic method, 233
vertical response function, 60
SFL, 153
shales, 36 water saturation, 9, 52
shoulder beds, 28 water-base muds, 29
shoulder correction, 266 wireline logging, 39
skin depth, 61, 196
skin effect, 60, 190, 308 X-signal, 57
skin effect correction, 97
SLDM, 225
software focusing, 135
sonde, 39
sonde error, 86
sonic log, 7
spectral integration, 197
spectral Lanczos decomposition method,
225
Spherically Focused Log, 77, 153
spontaneous potential, 6
squeeze configuration, 68, 157
survey current, 117

tensor representation, 45
thin beds, 32
TI anisotropy, 45
time domain, 43
time-line, 56, 117
tool coefficient, 119
tornado charts, 269
transfer impedance, 134
transition zone, 30

You might also like