Schlumberger - The History of A Technique

Download as pdf or txt
Download as pdf or txt
You are on page 1of 376

NUNC COCNOSCO EX PARTE

THOMASJ. BATA LIBRARY


TRENT UNIVERSITY
SCHLUMBERGER
The History of a Technique
Digitized by the Internet Archive
in 2019 with funding from
Kahle/Austin Foundation

https://archive.org/details/schlumbergerhistOOOOalla
SCHLUMBERGER
The History of a Technique

LOUIS A. ALLAUD
MAURICE H. MARTIN

A WILEY-INTERSCIENCE PUBLICATION

JOHN WILEY & SONS


New York • London • Sydney • Toronto
T ^0 V\ wo'S»

Copyright © 1977 by John Wiley & Sons, Inc.

All rights reserved. Published simultaneously in Canada.

No part of this book may be reproduced by any means,


nor transmitted, nor translated into a machine language
without the written permission of the publisher.

Library of Congress Cataloging in Publication Data

Allaud, Louis A.
Schlumberger.

“A Wiley-Interscience publication.”
1. Electric prospecting—History. 2. Oil well
logging, Electric—History. 3. Petroleum engineering—
History. 4. Schlumberger Limited. I. Martin, Maurice H.
joint author. II. Title.

TN269.A4913 622'.18'282 77-23566


ISBN 0-471-01667-5

Printed in the United States of America

10 987654321
To the memory of
Conrad and Marcel Schlumberger

&S1S43
Foreword

This book presents and explains the science-based techniques


which, over the last half-century, have so vastly enhanced the pros¬
pecting for and the production of petroleum. The volume is a fitting
tribute to the prominent part played by the ideas and work of
Conrad Schlumberger in the development of these techniques, and
in the creation of a large service company ranking at the very top of
the specialized organizations devoted to the needs of the petroleum
industry.
Between 1912 and 1926 Conrad Schlumberger, at first alone
and then with a small, slowly growing number of associates, in¬
vented an original mineral-prospecting method based on the possi¬
bility, which he verified, of gathering from electrical measurements
taken at the surface of the earth applicable information on the
geometric and physical structure of subterranean geological forma¬
tions. His success required not only a clear understanding of the
physical phenomena and their potential utilization, but also a deep
sense for experimentation in the laboratory and, above all, a genius
for observation in the field. From the observation of diverse and
complex facts he knew how to draw valid interpretations which
basic research on necessarily simplified models, or even experi¬
ments based on this research, could only suggest. It is noteworthy
that from the beginning his appreciation of the reasons for field test
failures made such failures almost as valuable as successes.
Foreword

Yet for Conrad Schlumberger there was a purpose broader


than merely establishing principles for new methods of prospecting:
what he wanted was to prove their practical value under the most
diverse conditions and so to pave the way for their worldwide
application. To this end, he promptly recognized the need for
designing sturdy tools with which delicate electrical measurements
could be made quickly in the field, even under difficult conditions.
In 1919 Conrad’s younger brother Marcel joined him, contributing
his own remarkable inventive mind and his special skill for solving
mechanical problems.
The establishment in 1926 of a small company, the Société de
Prospection Electrique, marked the beginning of industrial de¬
velopment. A crucial factor in shaping the Company’s future was
the association with the Schlumberger brothers, in the same year, of
Henri Doll, a young engineer. Within a year he was taking a large
and gradually leading part in a new orientation of the research
program, aimed at the employment in the oil wells themselves of
the principles of electrical surface prospecting. In a few years this
new field of electrical prospecting was to expand rapidly under the
name, first, of “electrical coring” and, later, of “electrical logging.”
The great technical obstacles encountered required the develop¬
ment of highly sophisticated and dependable electromechanical
tools; ensuing improvements in turn led to the delineation by
electrical measurement of additional parameters such as borehole
temperature or inclination. The recording of all these parameters as
a function of depth constitutes logging in its broadest sense.
World War II seriously disrupted the Company’s operations.
Only in the United States, under the management of Henri Doll,
could there continue to be any significant progress. One achieve¬
ment of this period was the successful extension of the recruitment
policy which had prevailed from the beginning of the work in
America: the association, particularly in the research teams, of
French and American engineers with widely different backgrounds.
After the war, research and development activities were con¬
centrated in the United States under the continuing leadership of

viii
Foreword

Henri Doll; at the same time, close contact was maintained with the
activities being resumed in France. During the postwar years, Henri
Doll brought about a growing use of electronics in logging equip¬
ment and advances in the interpretation of measurements. One
result was the electromagnetic induction method, which he had
already thought of before the war. It came into standard use after
1949, and was to be the last original logging method introduced by
Schlumberger.
I would like to take this opportunity to express my high regard
lor Henri Doll, whose outstanding human and professional qualities
I was able to appreciate in 1943 when, as a visiting professor at
Columbia University, I was invited by him to pay a few short visits
to Houston to consult on certain aspects of the nonprofit organiza¬
tion established by Schlumberger in support of the Allied war
effort.
Reading this book, scientists concerned with the penetration of
the scientific spirit into the whole range of human endeavor will be
happy to see that, even in the rather specialized field which is the
book’s subject, this spirit need not yield to the sense of business. Nor
have Schlumberger’s leaders lacked a keen business sense. Al¬
though the book is primarily addressed to engineers, who will find it
a source of valuable information, I venture to state that readers who
are not technically trained but who are eager for a better insight into
the reasons for the prestige of certain industries will also find this
modern epic unusually attractive even if they pass rapidly over the
more technical paragraphs.

Francis Perrin
Member of the Institut de France and

Professor Emeritus of the Collige de France

IX
Acknowledgment

We express our deep appreciation to Annette Gruner-


Schlumberger. It was her idea that a book be written about the
history of the Schlumberger technique. Without her guidance, en¬
couragement, and support, this book could never have been pub¬
lished.
We also thank Marcel Schwob for his translation of the French
text, and Frank Davidson, W. J. Gillingham, and R. R. Rieke for
their very helpful suggestions, revisions, and corrections in the
preparation of the English text.

Louis A. Allaud
Maurice H. Martin

Paris, France
Contents

, Introduction 1

Mineral exploration in 1910 4

The electrical resistivity of rocks 8

PART ONE SURFACE ELECTRICAL


PROSPECTING

Phase One: Before 1914 13

First laboratory studies 13

Electrical model studies. The bathtub.

First field tests 17

Vdl-Richer (Normandy). Choice of direct current. Technique


of the potentials map.

First surveys of geological structures 20

Sassy, Fierville-la-Campagne, Soumont, Sees (Normandy).


Contents

The exploration of conductive ores 30

Induced polarization. Spontaneous polarization. The


grounding method. The situation at the eve of World War I.

Phase Two: After 1919 35

Technical Developments 35

An overall view 35

The concept of apparent resistivity 37

The quadripole AMNB. Depth of investigation. AB profile


and dragging techniques. Resistivity map. Vertical electrical
soundings.

The potentiometer and its accessories 46

Basic design. Switching box, shunt, and Bowden. Millivolt


cartridge and induction corrector. Leaks. Pulsated-alternating
current. Pulsator. Skin effect.

Toward increasing depths of investigation 55

Exploration of deep structures. Electrical sounding at Vitre


(Brittany).

Prospecting for metalliferous ores 57

Spontaneous polarization, equipotential curves, and


resistivities. The differential magnetometer.

First studies based on the anisotropy of formations 58

Description of the phenomenon. Plotting of equipotential curves


with the source on the surface or inside a borehole. The
electromagnetic method: the spire.

xiv
Contents

Laboratory work and a bit of mathematics 62

Vdnous devices studied in the bathtub. Mathematical treatment


of the problem of horizontal layers. Charts for the
interpretation of electrical soundings. Dissemination of
information.

Operations

From 1920 to 1929 67

Short-term missions in many countries. The beginnings of


petroleum exploration: Rumania, United States. Potash
exploration in Alsace.

Electrical prospecting in the U.S.S.R. 77

Grozny region. Baku: the first offshore geophysical survey.


Development after 1932.

After the establishment of the Compagnie Générale de


Géophysique 82

History of C.G.G. Continued prospecting in Alsace.


Underwater survey in the Algiers roadstead. Corrosion of buried
pipes. Prospecting by telluric currents. Other developments.

PART TWO THE ORIGINS AND DEVELOPMENT


OF LOGGING 93

Introduction. Petroleum exploration. Rotary drilling.


Geological surveying and well completion. Sand-shale
series and limestone fields.

Phase One: 1927-1932 101


XV
Contents

Pechelbronn: A first resistivity log 101

Principle of the measurement. The operation of September 5,


1927. Description of the equipment. Subsequent field tests.

Developments in basic equipment 108

The first tricables. Measurement of insulation and location of


leaks. Winches. Hand-operated recorder.

The spontaneous potential curve (S.P.) 116

A first significant diagram at Pechelbronn. The hypothesis of


electrofiltration. Simultaneous recording of resistivity and S.P.
Tests on formation pressures. The discovery of electrochemical
S.P. at Baku.

Current output sonde and normal and lateral sondes 123

Differentiation of thin layers. Principle of the current output


sonde. Relationship between reserves, true resistivity, and
apparent resistivity. The MN-below sonde. The Courrieres
diagram. The AM sonde. The 1932 sonde. Generalization of
the normal sonde.

Other measurements in boreholes 134

Teleclinometer and electromagnetic dipmeter. Resistivimeter and


location of water intrusions. The electrical resistance
thermometer.

Field operations 143

Pechelbronn. Venezuela. United States. U.S.S.R. Indonesia.


Rumania.

xvi
Contents

Phase Two: 1933-1940 159

An overall view 159

Measurement techniques and interpretation methods 160

The case of oil-base muds. House documentation. Short normal.


The third curve (deep lateral). Long normal. Lateral sounding
curves. Studies on the true resistivity -saturation relationship.
Research on electrochemical S.P. and I.P. Problems of shaly
reservoirs and limestone fields.

Cables' winches, and miscellaneous tools 170

The quadricable. Problems of braid and gas attack. Sondes and


first head. Layout of trucks. Sheave and support.

The advent of photographic recorders 173

Shortcomings of hand recording. An attempt to use a


servomechanism and a high-speed mirror galvanometer. Two
and three galvanometer models. Optical system. First attempts at
single-conductor cable system.

Temperature, well drift, and formation dip 179

Thermometric cementing control. The photoclinometer.


Introduction of the electromagnetic dipmeter. S.P. dipmeter.

Sidewall sampling and casing perforation 182

The importance of these two operations. Test of a lever sampler.


Sampling gun. Plugged bullets. Disposable socket bullets. The
yields: sand-shale series, friable formations {sand bullet), hard
rocks. Testing of a punching and a drilling perforator. Bullet
perforator. Annular chamber gun. Lined bullet. Unit
perforator.

XVI1
Contents

Operations 191

United States, South America, Rumania, Europe, Far and


Middle East. Working conditions of the crews.

Phase Three: 1940-1945 195

France 195

The Paris center. The vihrationproof galvanometer. Studies of a


remote transmission system: the “sequential” system. The
combiné. Study on the injection of cement rings. Basic
research on electrofiltration.

United States 198

The Houston center. Its operation. Research on S.P. Work for the
War Department. Mine detector. The beginnings of
quantitative interpretation.

PART THREE: AFTER 1945

Phase One: 1945-1957 207

The situation and the prospects at the end of World War II 207

Resistivity. Perforation. Auxiliary operations. Condition of


equipment. Prospects.

The administrative reorganization 211

United States, South America, Europe, eastern hemisphere.


International status of engineers.

xviii
Contents

The series 500 and 700 trucks 215

Concept and choices. Chassis. Winch. Six- or seven-conductor


cable. Tension gauge. Depth measurement. Offshore Skid Unit,
Type C (O.S.U.-C.). Technical comparison of single-, six-, and
seven-conductor cables.

The postwar progress in interpretation 227

The two phases of interpretation. Main publications. Training


courses and schools.

The new resistivity logs 236

The microlog technique 237

Description. Commercial development. Improvements and


prospects. First powered sonde.

Induction logging 243


General problem of true resistivity measurement. Development of
induction for oil-base muds. Extension to water-base muds.
Basic advantages of the system. First field tests. Expansion.
Combination with short normal.

The laterolog technique 254

The problem of thin beds. The case of salt muds. Limestone sonde.
Origin of laterolog 7. “Electrical plug.” Technological choices.
Difficulties and solutions. First tests and development. Guard
electrode log and laterolog 3. Technical success.

The microlaterolog technique 261

Adaptation of Laterolog to the pad. Technical realization. First


tests. Intrinsic value of Microlaterolog. Notions of quantitative
interpretation. Maximum producible oil index. Operational
difficulties.
Contents

The nuclear logs: Gamma rays, neutrons 267

Gamma ray logs 267

General. First realizations: Well Surveys, Inc. Depth control.


Radioactive markers. Schlumberger logs.

The neutron log 270

History. Theory of neutron log. Porosity measurement. Gas


detection. Schlumberger log. Worldwide development.

Auxiliary operations 275

Dipmeter surveys 275

Operation of correlation dipmeter. Limitations ofS.P. dipmeter.


Resistivity dipmeter. Continuous dip measurements. Microlog
dipmeter. Poteclinometer-dipmeter.

Thermometry 281

The borehole caliper 282

Sampling 283

Need for improving the sidewall sampler. The C.S.T. The


mechanical sampler. Fluid sampling.

Perforation 288

Bridge plugs. Production packers. Loading truck. Supergun.


Shaped charges. Perforation below tubing.

Phase Two: 1957-1974 295

XX
Contents

A general view 295

Establishment of Schlumberger Limited. Administrative,


commercial, and technical evolution. New prospects.

The modem tools 297

Resistivity tools 297

Induction. Combinations. Laterolog. Dual Laterolog.

Nuclear tools and techniques 303

New gamma ray detectors. Evolution of the neutron technique.


Density logging. Thermal neutron decay log (T.D.T.).

The sonic log 307

History. Initial developments. Interpretation. Amplitude


measurement. Cement bond log. Other applications. Borehole
televiewer.

Miscellaneous techniques and tools 314

New dipmeter surveying. Sampling. Perforation. Production


logging.

A Note on offshore drilling 319

Presentation and interpretation of logs 320

Field interpretation. Quick-look methods. Computerized


interpretation. Transmission of data and results.

Conclusion 329

XXI
Figures

1 Schematic cross section of an anticlinal structure.


2 Diagram showing the uplift of strata by a saline surrection.
3 Equipotential curves at the surface.
4 Conrad Schlumberger performing a field measurement.
5 Paraphernalia of Conrad Schlumberger.
6 Experiment at the Val-Richer.
7 Experiment at Sassy.
8 Experiment at Sassy: enlarged detail.
9 Equipotential curves around ground electrode A.
10 Test at Fierville-la-Campagne.
11 Sain-Bel: map of spontaneous potentials.
12 Measurement of apparent resistivity (schematic).
13 Variation of the depth of investigation according to the position of MN between A
and B.
14 First resistivity profile.
15 Hypothetical cross section of an anticline and corresponding resistivity profile.
16 Potentiometer.
17 Measurement with potentiometer.
18 Operation of the pulsated-rectified current (schematic).
19 California: San Joaquin Valley.
20 Schematic view of the spire.
21 Resistivity map of the Aricesti, Rumania, salt dome (1923).
22 Electrical prospecting in Louisiana swamps.
23 Meyenheim anticline (1926).
24 Hettenschlag Dome (1926).
25 First known example of offshore geophysical prospecting.
26 Telluric current survey of the Hettenschlag area.
27 Sand-shale series and limestone field.
28 Schematic cross section of a completed rotary oil well.
29 Schematic view of electrical coring.

xxiii
Figures

30 First electrical coring diagram.


31 Early electrical coring: winch, cable, and pickup truck.
32 Truck and winch.
33 Electrical coring measurement with potentiometer.
34 First hand recorder.
35 First diagram where S.P. distinctly indicated a permeable bed.
36 Currents and S.P. curve opposite a permeable bed.
37 Distribution of current emitted by an electrode.
38 Facsimile of a sketch of the current output sonde by Conrad Schlumberger.
39 Vertical cross section of a formation invaded by mud filtrate.
40 Two resistivity diagrams through a coal seam.
41 Normal sonde and reverse or lateral sonde.
42 Equipotential curves surrounding an electrode in a homogeneous and anisotropic
dipping formation.
43 Part of mechanism inside the inclinometer.
44 Doll calibrating a teleclinometer in Baku.
45 Facsimile of the cover page of the first article on electrical coring.
46 First electrical coring diagram outside France.
47 First electrical coring diagram in the United States.
48 An early job in the United States.
49 A log recorded with standard equipment: short normal, long normal, lateral, and S.P.
50 Faboratory device for the study of the electrochemical component of S.P.
51 Measuring cab on a truck.
52 Truck and winch.
53 Fog recorded with the high-speed mirror galvanometer.
54 Two-galvanometer photographic recorder.
55 Schematic view of the S.P. dipmeter.
56 Schematic cross sections of gun sidewall sampler.
57 First sample-taker used as power electrode.
58 Sample-takers and perforators.
59 Schematic cross section of the annular chamber perforator barrel.
60 Cross section of a unit of the so-called unit perforator.
61 Ridgefield research center.
62 Headquarters of Schlumberger Well Surveying Corporation in Houston.
63 Series 500 truck.
64 Marcel Schlumberger in a cab on a truck.
65 Cab with photographic recorder.
66 O.S.U.-C. unit for offshore wells.
67 S.P. diagram: current flow.
68 S.P. in limestones.
69 Microlog pad and sonde.
70 Motorized microlog sonde.
71 Principle of induction logging.
72 First induction log.

xxiv
Figures

73 Comparison between induction-electrical log and conventional log.


74 Schematic view of the limestone sonde.
75 Schematic view of laterolog 7.
76 Pad of the microlaterolog.
77 Poteclinometer.
78 Proposed needle-igniting system.
79 Formation tester.
80 Four inch load carrier.
81 Printed circuit and modular element of cartridge.
82 Principle of the double laterolog.
83 Schlumberger sonic log patent.
84 H.D.T. dipmeter sonde.
85 Schlumberger unit on offshore platform.
86 Computer-processed interpretation.

XXV
SCHLUMBERGER

The History of a Technique


Introduction

Schlumberger is a large concern belonging to what economists call


"the service sector" of the business world. Although engaged for some 15
years in marketing products of its own manufacture, Schlumberger re¬
mains essentially what it was at the outset—a company offering
specialized services to industry. The field to which these services are most
applicable is petroleum exploration and production; the discipline is
geophysics, a technology developed only during the last half century.
Comparatively recent developments in geophysical techniques in¬
clude improved procedures for the measurement of certain physical
properties of underground rocks. Great advances in mineral detection
have resulted from applying these methods to rock density, electrical
resistivity, sound transmission, radioactivity, and magnetism. By using
special equipment, the measurements can be made either on the surface
of the land or inside boreholes. In the first case, they help to outline broad
features of the subsurface; in the second, they provide information on the
characteristics of the formations penetrated by the drill.
Formerly called “electrical coring,”1 and later known as “logging,”

1 The term “electrical coring’’ (in French, carottage électrique) was adopted in 1927 because
of its anology to “mechanical coring,” which denotes the analysis of rock samples (“cores,”
in mining terminology) extracted from a formation by drilling. The French word carottage
(“coring”) was introduced as such into the Russian language in 1930 to designate the new
technique. The English translation “electrical coring,” under which the process became
known in the United States, was replaced in 1933 by the term “electrical logging,” now in
use throughout the petroleum world. The word “log” means a strip of paper or film, on
which measurements are recorded as curves (diagrams) in terms of depth. The word is
borrowed from nautical science, where it refers to the plotting of the data that determine
the motion of a ship in terms of time. After 1945 French-speaking oil people translated

1
The History of a Technique -

the measurements in boreholes are the most important of the services


provided by Schlumberger. It is essential, therefore, to grasp the meaning
of logging as it applies to the petroleum industry.
The first objective of a drilling operation is to obtain as much infor¬
mation as possible on the various (and varied) formations penetrated,
particularly those that may contain and yield oil; the second is to utilize
the drillhole to produce oil from a reservoir found to be oil bearing.
Logging is a convenient tool for obtaining much of this necessary subsur¬
face information. It is performed by lowering an appropriate and highly
sensitive device, by means of an electrical cable through which the physi¬
cal properties mentioned above (except certain magnetic ones) can be
measured and recorded in a detailed and continuous manner over the
entire length of the open borehole. Through an analysis of the informa¬
tion obtained, the drilled formations may be characterized, their limits
accurately determined, and the oil- or water-bearing levels identified. The
basic physical property which the original Schlumberger log sought to
delineate was the electrical resistivity of the rocks. Eventually, measure¬
ment of other properties increased the efficiency and scope of the
method.
Schlumberger owes its existence to the electrical measurement tests
applied to the surface by Conrad Schlumberger in France, a few years
before World War I. This was the first time that such tests had ever had
practical significance; they led to the development of innovative mineral¬
prospecting methods that came to play a major role in the application of
geophysical concepts.
In 1919 Conrad and Marcel Schlumberger, with a handful of assis¬
tants and very little capital, launched a small venture for the purpose of
applying these methods to the detection of metallic minerals, to geologi¬
cal surveys, and to petroleum exploration. Foregoing any idea of self¬
enrichment through personal speculation in promising mineral areas, the
founders committed themselves to providing contract services of the
highest scientific and technical standards. This commitment has remained
a basic tenet of the Company.
After having contributed to the birth of geophysics, Conrad

"logging” into diagraphie, a new word derived from “diagram.” In certain works (mostly
textbooks), the expression borehole geophysics appears in all languages where logging is
described. For the sake of historical accuracy, the term “electrical coring” will be used here
for the technique prevailing until 1933, and “logging” thereafter.

2
Introduction

Schlumberger in 1927 invented electrical coring. After 1931, when the


firm s surface-prospecting methods were taken over by the Compagnie
Générale de Géophysique, the Schlumberger group devoted itself prin¬
cipally to measurements and operations in boreholes. Then, in 1934, after
15 years of relatively slow growth, the company experienced a vigorous
expansion extending its activities to most of the oil-producing regions of
the world. In addition, in the late 1950’s, while continuing to diversify its
services to the petroleum industry, Schlumberger made broad inroads
into the fields of electrical and electronic measurement and instrumenta¬
tion.
At the present time this multinational enterprise ranks 148th among
world firms (excluding those of the United States). Its 17 companies, with
operations in more than 70 countries, employ a staff of 40,000. The group
operates 18 research centers; the 3 major ones are devoted to logging.
Such growth is comparable to that experienced by certain pioneers
operating in new countries. Indeed, the founders of Schlumberger were
pioneers; the territory into which they ventured was an almost virgin area
for mineral exploration. A continuous flow of creative ideas shaped their
work. Sparked by the desire to meet the challenges of a rapidly evolving
petroleum industry and to anticipate its requirements, the original ideas
have bred more and more refined and complex techniques. The Company
has been served by men who have efficiently organized and led industrial
development under the most diverse conditions of geography, climate,
infrastructure, law, and communications.
The purpose of this book is to provide an historical account of
Schlumberger techniques from their inception to their current applica¬
tion, and to describe where and how they were employed. Lengthy techni¬
cal descriptions will be avoided to spare nonspecialized readers.
The contributions made by Schlumberger in the field of surface
geophysical prospecting were developed during the first part of this
century, when the Company was functioning on a small scale. Today these
accomplishments are almost forgotten in spite of their far-reaching ef¬
fects: it was from them that logging, now in worldwide use, was de¬
veloped; and they included enough scientific discoveries and inventions to
justify a rather detailed historical review in the forthcoming pages.
For its first 15 years the company resembled a family concern with a
modest staff; in many respects this was the most attractive period of its
history. The founders had a hand in every detail; everyone knew everyone

3
The History of a Technique

else personally. No two pieces of equipment manufactured for the same


purpose were ever identical when they left the workshop. As Conrad
liked to say, everyone was involved “for better or for worse.” More space
will be devoted to details of growth and development in this period than
to subsequent periods when the staff grew to hundreds and then
thousands, and when services multiplied and spread throughout the
world. After 1934 the emphasis on personal recognition will necessarily
diminish because the list of those who contributed to the common
endeavor becomes too long—this is not to imply that these later
associates deserve any lesser acknowledgment than those who formed the
initial team between 1920 and 1933.
Although Schlumberger has held first place in a discipline of its own
creation, other companies have offered similar if not identical services to
the petroleum industry. Many engineers, geologists, and geophysicists in
industry and in the universities have undertaken studies on specialized
aspects of logging, particularly interpretation. The results have often
provided Schlumberger with useful insights for the development and
application of its processes. Since a listing would be too long for inclusion
in this book, the reader may wish to refer to the specialized publications.2
To indicate the originality and importance of Conrad’s first studies,
this introduction will outline the field of mineral exploration as it existed
at the beginning of the century. Some data will also be provided on the
electrical resistivity of rocks in the crust of the earth. The measurement of
this parameter is basic to electrical surface prospecting, and it constitutes
one of the truly characteristic features of logging.

Mineral exploration in 1910

For many metallic ores (e.g., iron, copper, zinc, lead, manganese) the
exploration devices available in 1910 were much the same as they had
been centuries ago. These ores, whether in the form of lodes or intrusives
in older rock formations, are mostly eruptive; prospecting for them
basically consisted of searching for visible signs on the surface, either in

2 A substantial, if not exhaustive, bibliography can be found in “Well Logging Tech¬


niques,” by J. Riboud and N. A. Schuster, Transactions of World Petroleum Congress VIII
1971.

4
Introduction

already productive areas or in other areas with similar geological features.


Occasionally, thanks to geological experience, a deep, invisible deposit
was found; but since the distribution of such lodes is generally erratic,
their discovery was often due more to chance than to the prospector’s
educated intuition.
Other ore bodies (e.g., certain iron ores, coal, salt) have deposits
formed similarly to sedimentary rocks. These are found in layers in¬
terspersed among ancient formations. For a long time their exploration
depended on surface studies of the structure of the sedimentary series
which presumably contained them. Geological exploration gradually be¬
came more reliable and significant. From observations and measurements
taken where the subterranean layers outcropped, reasonable appraisals of
their underground structural patterns, such as folds, breaks, overthrusts,
and faults, could be made. The methods employed were those of tec¬
tonics, still a young science at that time.
On the other hand, actual physical inspection of the economic
mineral—the ultimate target of exploration—no longer necessitated the
digging of man-sized pits and tunnels until the time of exploitation. For
several decades it had proved faster and less costly to drill small-diameter
coreholes for cutting deep rock samples from the formations penetrated.
Methods of examining the samples had been perfected, and advances in
the new science of paleontology helped to determine the age of rocks,
thus narrowing the scope of exploration.
Despite the persistence of the widely shared belief that petroleum
gushed out of vast underground caverns (an idea popularized by many
publications, including highly scientific ones), most professionals had long
known that petroleum, liquid or gaseous, was to be found in the pores of
permeable rocks such as sandstones and limestones. Since the first oil well
was drilled by Colonel Edwin Drake in Pennsylvania in 1859, studies by
English and American geologists had verified the close relationship be¬
tween oil-bearing formations and deep geological structures. It was
known that oil, being lighter than water, may accumulate in the upper
parts of porous strata folded in the shape of arches or domes, and overlain
by impermeable layers such as clays or shales (Figs. 1 and 2). This
knowledge led to the “anticlinal theory,” which was to become a key tool
in petroleum exploration.3 To be sure, not everything had been clarified,

3 An anticline is a geological structure in the shape of a vault or an arch. A syncline has the

5
The History of a Technique

and many concepts held at that time were to be refuted or modified by


experience. Anyway, in 1910 the geologist was not too well regarded by
the majority of operators; the most popular way of prospecting for oil was
to drill a borehole near surface signs such as gas seepages, eternal fires,
asphalt lakes, and mud volcanoes.
Although the contributions of geophysics to the field of mineral
exploration remained limited, several centuries of observing the magnetic
and gravitational fields, and the more recent study of seismic tremors, had
enhanced the knowledge of our terrestrial globe.* * * 4 As early as the seven-

shape of a ditch. The word “dome” is self-explanatory. Many anticlines and domes have
resulted from the thrust of rock salt, which is relatively plastic as compared to other rocks.
Under the weight of the sedimentary overburden, wherever a passage is available, the salt
formation shoots up extensions, often nearly vertical, which take the shape of gigantic
ridges or mushrooms and either push their way through the surrounding layers (Fig. 2),
bend them back, or uplift them. Salt ridges and domes are prevalent in certain regions such
as Alsace and the Landes in France, the Gulf Coast (Gulf of Mexico) in the United States,
the Emba region in the U.S.S.R., and particularly parts of Rumania and Germany.
4 This area of learning was then called “geophysics” or “physics of the globe.” Later, when
measurement of the physical field would serve the purposes of mineral exploration, the
term “applied geophysics” was introduced. Nowadays “geophysics” refers mainly to the
scientific and technical studies involved in the exploration for minerals.

6
Introduction

Surface
Over-
N burden

Faults

Salt water
Permeable
beds

Figure 2. Diagram showing the uplife of the strata by a saline surrection.

teenth century, thought had been given to using the compass to detect,
among other things, magnetite bodies; yet it was not until the middle of
the nineteenth century that prospecting instruments of much greater
sensitivity and accuracy (Thalen and Tiberg compasses) led to important
discoveries in Sweden and the United States. In 1896 the Hungarian
scientist Roland Eotvos designed and built his torsion balance, an instru¬
ment sufficiently sensitive to detect the variations in gravity above the
cellars of his Budapest laboratory.5 However, its only practical applica¬
tion at that time was to measure the depth of Lake Balaton. Not until the
1920’s was the torsion balance used commercially in the search for salt
domes and oil fields.
In 1830 a British scientist, R. W. Fox, used a rudimentary apparatus

5 These measurements are the backbone of the gravimetric prospecting method.

7
The History of a Technique

to measure the electrical potentials generated by metallic sulfide deposits


in Cornwall.6 In 1880 an American, Carl Barus, published his observa¬
tions in the same field while studying mines in Nevada. Elsewhere, some
prospecting based on the electrical resistivity of rocks and utilizing high-
or low-frequency currents had been going on for almost a century; the
specialized literature fails to show, however, that these operations
brought about the discovery of any new ore bodies or the extension of a
known one.
In short, as late as 1910, geophysics had made scant impact on
mineral exploration. There appeared to be little appreciation of the role
that it might play in the study of geological structures; no attempt had
been made to apply it to the search for iron ore or coal. As for petroleum
exploration, which after 1925 would account for over 95 percent of
geophysical activities, those involved knew very little about the applicabil-
s

ity of geophysical principles. Professor de Launay at the Ecole des Mines


in Paris was one of the few who sensed these potentialities. After having
briefly appraised the efficiencies of the magnetometer and other instru¬
ments of that time for measuring the physical field, he wrote:

While one would think that the above-mentioned instruments would be of


such great value, their use . . . has been restricted to a few rare instances and
it does not appear that the spirit of scientific investigation, so active in every
direction, is here applied with an intensity commensurate with the practical
results that could be expected from a discovery.7

This, for geophysics, amounted to a statement of failure. There lay


ahead a vast and relatively unexplored field, open to those keenly inves¬
tigative and scientific minds that de Launay regretfully found wanting.
The field of mineral exploration was ready for the studies of Conrad
Schlumberger.

The electrical resistivity of rocks

Rocks encountered in sedimentary basins are composed of mineral


grains more or less cemented. Most often, these grains do not fill the

6 At that early date Fox predicted that methods of the electromagnetic type could help in
prospecting for metalliferous lodes.
7 De Launay, La Conquête Minérale, Flammarion, Paris, 1908.

8
Introduction

entire space; they are separated by pores which vary in size with the type
of rock.
In the finest sediments, such as clays and shales, the size of the pores
can be as small as 1/100,000, even 1/1,000,000, of a millimeter. In sands,
sandstones, and many limestones, sizes vary from 1/1000 to 1/10 of a
millimeter, and in gravels and some vugs they reach 1 or more millime¬
ters. Yet rocks are found with crystals so tightly joined that there are no
pores at all. This occurs mainly in deep-hardened sandstones or lime¬
stones, or in matter produced by evaporation in deep water such as rock
salt or anhydrite. Igneous or crystalline rocks are, in general, similarly
compact.
In nearly all subsurface rocks, the pores are entirely filled with water
varying'in saline content from pure water, produced by rain or snow, to
salt-saturated water (often encountered in oil fields). The highly aerated
formations close to the surface are an exception. On the other hand, in
hydrocarbon deposits the pores of the rock contain both water and oil,
both water and gas, or the three fluids together. With a few exceptions,
which will be considered later, mineral grains are electrical insulators. The
rocks they form allow electric current to flow because they are permeated
with water. Therefore the resistivity of a rock is of the electrolytic type; it
is proportional to the resistivity of the water and decreases steadily when
the temperature rises. The smaller the amount of water contained in a
given volume of rock—in other words, the less porous the rock—the
higher the resistivity. Resistivity is also contingent, although to a lesser
extent, on the shape and arrangement of the pores, rather than their size.
Thus there is only a loose relationship between the resistivity of a rock
and its permeability.
Rock porosity varies widely, from a few percent in consolidated
limestones, to 30 and 40 percent in sands and unconsolidated clays, and
up to 80 or 90 percent in the silt of lakes and oceans. Since the salinity of
waters also varies widely, the resistivities encountered underground vary
considerably too, from a fraction of an ohm-meter8 in unconsolidated
formations filled with salt-saturated waters, to several thousand ohm-
meters and on to infinity in compact rocks. No other physical parameter

8 The resistivity unit adopted by Schlumberger and used almost universally in electrical
prospecting is the ohm-m2/m or, more simply, the ohm-meter; that is, the resistivity of a
cube 1 meter on a side, the resistance of which, to a current parallel to an edge, equals 1
ohm.

9
The History of a Technique

(e.g., density, acoustical propagation velocity, magnetic susceptibility)


presents such a wide range of values.9
When exploration is undertaken in a limited zone, where each sedi¬
ment maintains a sufficiently uniform structure over a fairly large area,
and the salinity of the water does not change appreciably, resistivity
becomes a sensitive and accurate factor in identifying the formations.
Furthermore, when the water in the rock is partially replaced with oil, gas,
or both, the resistivity is much higher than if the pores are entirely filled
with water of even salinity. Depending on hydrocarbon saturation,10 the
resistivity of an oil-gas formation may be 5, 10, or 100 times higher than
that of a similar water-bearing formation. This property has far-reaching
implications in logging since it offers a sensitive and dependable indicator
for the differentiation of productive and barren zones.
Finally, certain ores, mainly metallic sulfides (e.g., pyrites), are made
of crystals that conduct electricity as do metals. This metal-like conductiv¬
ity (“electronic” is the term now used) is much lower than the conductiv¬
ity of ordinary metals, yet much higher than that of other rocks, even
those having the lowest electrolytic resistivity. A characteristic feature of
this type of ore, it has proved useful in prospecting.
When Conrad Schlumberger began his research in 1911, only very
fragmented data on the resistivity values of rocks were available to him.
Although he might have intuitively foreseen the significance of formation
resistivity for mineral exploration, there was no way he could have en¬
visioned the impact it would have on geophysics and, more specifically,
the petroleum industry.

9 The density ratio of the heaviest to the lightest rocks does not generally exceed 3; the
same ratio applies to the velocity of sound waves through the subsurface; and for magnetic
susceptibility there is a marked difference only between crystalline and sedimentary rocks
(with the exception of certain iron ores).
10 In the language of the petroleum industry, saturation designates the fraction of pore
volume occupied by a fluid: water, oil, or gas.

10
PART
ONE

Surface Electrical Prospecting


Phase One: Before 1914

First laboratory studies

September 1912. At the family estate, the Val-Richer, in Normandy,


Conrad Schlumberger was putting the finishing touches on lines drawn on
a large sheet of paper, with a few reference marks indicating lawns, ponds,
driveway, the outline of a mansion, and farms. The lines appeared to mark
topographic contour levels; however, instead of connecting points of
equal elevation, they connected points of equal electrical voltage and
were called equipotential curves. Conrad had before him the plotting of
the first on-site measurements he had made to test, under actual condi¬
tions, his ideas about the exploration of the subsurface through the
transmission of an electric current.
These field tests were the result of many months of meditation,
theoretical work, and small-scale experimentation. During that same
summer he had developed, almost entirely by himself, the apparatus and
methods that would enable him to put into practice his theories. How¬
ever, since there was no mineral wealth buried in the grounds of the
Val-Richer, his tests served merely as a feasibility study. The results were
strikingly clear; the measurements ranged neatly along regular equipoten¬
tial curves, whose shape and spacing were easily explained by the ground
relief. Although others before him had sent current into the ground, they
had become bogged down in insurmountable difficulties; no one else had
succeeded in making such precise, cohesive, and significant observations:
a method had truly been “discovered.”
What were the underlying principles of this discovery, and how did
they give birth to a method? Looking back almost 10 years later, Conrad
would answer that the same circumstances that had earned him a profes-

13
The History of a Technique

sorship at the École des Mines in Paris* had later compelled him to devote
his holidays to researching the application of physics to the field of
mining, with particular emphasis on prospecting. It did not behoove him
to add that these “circumstances” were strongly helped by a rare intel¬
ligence, exceptionally gifted for the abstract and practical, a fertile
imagination, a varied cultural background, and a lively curiosity controlled
by a rigorous scientific approach. To a man so endowed, who was little
more than 30 years old, an academic career, however prestigious, could
hardly bring fulfillment. Teaching a course in physics at the École des
Mines was not, however, without its advantages; it provided a quiet
occupation, scientific surroundings, a vast library, and the stimulation of
contact with a young audience. This was the “environment” that became
the fertile ground of creativity for one whose mind was shaped by mathe¬
matics and physics. Thoroughly aware of the problems of the mining
industry, and having a keen interest in the sciences of the earth, Conrad
was inevitably attracted by the challenges of the field later known as
geophysics. Furthermore, his fondness for research and his desire for
orderly and efficient work, as well as his concern for economic and social
factors, led him toward efforts to improve the methods of mineral produc¬
tion.
The search for metalliferous deposits was expanding during this
period, and Conrad’s first thoughts were directed toward methods for
improving reconnaissance. The problem was to decide on a physical
parameter with which one could accurately differentiate the deposits
themselves from the rocks in which they were embedded.
According to information available at the time, electrical conductivity
appeared to be a more significant parameter for the exploration of metal¬
liferous deposits than the rather uncharacteristic density, or magnetic
properties, limited to certain types of iron ores. However, while a mag¬
netic mineral body may be revealed by the disturbance it creates in the
magnetic field of the earth, there is no similar natural electric field that can
be scrutinized in the search for conductive ores.
Conrad thought, therefore, of creating an artificial electric field. The
simplest way was to transmit a current between two electrodes (two series
of pegs) driven into the ground, and to plot the equipotential lines
reflecting the distribution of resistivities of the subsurface.

* Graduate institute of mining engineering. (Translator’s note.)

14
Surface Electrical Prospecting

In the case of a homogeneous underground, the path of the current


lines and the shape and the arrangement of the surfaces of equal potential
were well known. A simple calculation shows that the current does not
flow inside the narrow channel connecting the two electrodes, the most
direct route (as certain experimentalists believed), but spreads in breadth
and in depth across the subsurface. The surfaces of equal potential inter¬
sect with the plane surface of the ground, forming equipotential lines (Fig.
3). Should an ore body with a higher conductivity than the surrounding
formation (always assumed to be homogeneous) occur, these lines would
be distorted, as revealed by the plotting.
It appears that this reasoning was somewhat oversimplified because
the strata around a mineral bed are never homogeneous. Certain difficul¬
ties, which Conrad hoped to overcome in future experiments, had to be
expected. Still, no serious theoretical objection could be raised against the
apparatus he had in mind, whereas the system unsuccessfully operated a
few years earlier by other researchers (Brown and MacClatchey)1 could
not withstand close criticism.
Early in 1911 Conrad conducted a study of the various elements
involved in his initial thinking. The laboratory of the École des Mines
afforded him the opportunity to examine ore samples and to measure
their electrical resistivities in order to supplement, correct, or confirm the
published data. He also extended his observations to geological forma¬
tions of sedimentary origin, especially argillaceous sands in recent alluvial
deposits. Small-scale devices were set up in the basement of the École des
Mines. At first, wooden crates were filled with a mixture of sand and clay
impregnated with water that varied in saline content. In these crates a
current flowed between two current electrodes (hereafter referred to as A
and B), and its effect was explored by means of two measure electrodes
(hereafter called M and N), with which the carefully leveled surface was
scanned. The alternating current used was of audible frequency. Silence,
or reduced noise, in a pair of earphones connected by loose wires to
electrodes M and N indicated that they were at the same potential.
Step by step, Conrad plotted grids of equipotential curves, checking
that they confirmed the theory. Other experiments using various electrode

1 Brown and MacClatchey had been concerned solely with measuring resistance through
the ground between two ground electrodes conducting the current. They were unaware of
the fact that 95 percent or more of the resistance is concentrated very close to the current
electrodes and remains the same whether or not an ore body is buried between them.

15
The History of a Technique

16
Surface Electrical Prospecting

arrangements followed, but the wooden crates were no longer adequate.


A watertight, metallic tank was needed to hold water of various salinities
and to set a zero potential. The small copper bathtub that had been used
by his children approximated these conditions and became the indispens¬
able item of equipment in Conrad’s experiments. To this day the term
"bathtub" is used in the idiom of “La Pros” (a nickname of the Company
originated by French engineers; see p. 37) to designate a metal tank
through which the passage of a current reveals a variety of complex and
informative data.
With these items of equipment it was possible to answer a question
that was basic to the practical application of the method: would a deeply
buried body of metalliferous ore reveal its presence by sufficiently distort¬
ing the equipotentials? Conrad devised experiments wherein a paral-
lelepipedic slab of potter’s clay was placed horizontally inside the clayey
sand of a crate and the effect on the curves observed. Repeated on several
crates, each one with its slab of clay at a different depth, these seemingly
simple experiments required meticulous attention to accuracy and detail
in order to achieve two media of natural materials that were sufficiently
homogeneous and had a resistivity contrast such that the measurements
were representative.
These experiments showed that the suggested method would work:
it was possible to detect a buried object provided that it was large enough
and its resistivity contrasted sufficiently with that of the surrounding
formations. The idea of adjusting the distance between A and B (hereafter
termed current electrode spacing) gave rise to a special advantage of the
current transmission method: the possibility of altering the depth of
investigation to meet the desired objective.

First field tests

After a few months of preliminary testing, done on a relatively low


budget during free time between teaching assignments, Conrad had de¬
veloped his initial ideas to the extent that he felt ready to undertake
full-scale tests.
In 1912, a summer vacation provided the necessary leisure, and the
park of the Val-Richer a field for experimentation (Fig. 4). The undertak¬
ing was not an easy one. It was no simple task to shift from laboratory-

17
The History of a Technique

Figure 4■ Conrad Schlumberger performing a field measurement (Normandy, 1912).

18
Surface Electrical Prospecting

scale models to actual field conditions. The earphone method had proved
disappointing. Mutual induction between the current-emitting (AB) and
measure (MN) circuits was negligible in the laboratory, but in the field,
with lines of several hundred meters, it became strong enough to jam
observations.2 Conrad then switched to direct current; experience
confirmed its superiority.
Within a few weeks he procured, tested, and adapted his instruments
and perfected the operating methods. Several months later, in the early
part of 1913, he summarized this first experiment on the same sheet of
heavy drawing paper on which he had originally plotted his mea¬
surements. He described the crude cables and reels borrowed from the
army, the small generator driven by the engine of the truck, the ground
stakes afid the way they were set into the soil to get good grounding, the
practical maneuvering of the electrode system, the trumpet signals used
for field communication between himself and his assistant, the deviations
of the needle, and finally the galvanometer itself. After testing several
galvanometer models, he finally found, in Frankfurt, one of adequate
sensitivity, accuracy, and convenience, made by Hartmann and Braun. He
would add to this galvanometer the components (batteries, resistances)
that would enable him to measure potential differences by the balance (or
null) method (see p. 47).
Another difficulty, arising from the use of direct current, had to be
solved. If M and N are plain pegs of copper, for example, electrochemical
phenomena take place at the point of contact of the metal and the water
present in the soil: the ground electrodes are said to “polarize.” As all
ground electrodes do not polarize in an identical manner, this imbalance
generates an electromotive force of up to several tens of millivolts, more
than enough to drive the needle of the galvanometer out of scale, making
any measurement impossible. Conrad conceived and built a nonpolarizing
type of ground electrode that was well suited for fieldwork (Fig. 5).
The subsurface of the Val-Richer testing ground was simple: a thick
layer of moist, uniform, clayey soil covered the ancient rocks, which were
embedded too deeply to affect the measurements. Electrode A was staked
at the center of a well-leveled lawn and connected to remotely located

2 Conrad has explained by these disturbance effects the failure of the earlier attempts of
L. Daft and A. Williams, who were sending alternating current into the soil and measuring
the potentials with earphones. Various other tests employing the same technique, espe¬
cially in Sweden, had not been followed up.

19
The History of a Technique

Figure 5. Paraphernalia of Conrad Schlumberger. 1 : Tripod with compass, a military instru¬


ment used for topographic survey of measuring stations. 2: Nonpolarizing electrode, the handle of
which features a coil used in the measuring circuit and the wooden box crating the potentiometer.
3 to 7: Electrodes of various lengths. 8: Horn to call the helpers. 9: Earphones. 10: Stake and large
coil for power circuit. 11 and 12: Current generators. 13 and 14: Storage batteries. 15: Geologist
hammer. 16 to 19: Commercial ammeters and voltmeters.

electrode B. Thus B had practically no effect on the shape of the curves


plotted around A; it was as if B had been removed to infinity. On the lawn
the equipotentials formed almost concentric circles, as stated in the
theory; to the northwest, where they spread out and tightened again, the
pattern indicated the presence of a slope (Fig. 6).
The test was conclusive, and the technique dependable.

First surveys of geological structures

The second experiment took place slightly to the south of Sassy,


Calvados. The main features of deep geology in this region are known,

20
Surface Electrical Prospecting

thanks to numerous iron mines. The subsurface is made up of steeply


dipping, occasionally almost vertical (Silurian) formations. The thick
basement is composed of three types of rocks: Armorican sandstones,
Calymenae shales (with a sedimentary seam of iron ore between them), and
May sandstones (actually a stack of sandstone and shale beds). These
ancient formations outcrop in places, depending on their hardness and on
the relief of the ground, but they usually remain hidden beneath younger,
softer, and nearly horizontal Jurassic marls and limestones. Conrad’s
purpose in undertaking the study was not to detect the iron ore, made of
hematite and carbonate and having no particular electric conductivity;
what he wanted to ascertain was whether the resistivities of various types
of rock were sufficiently distinct to reflect their shapes through the equi-
potentiai lines, and, if so, to what deductions this might lead (Fig. 7).
As previously noted, Conrad had, at that time, little information on
the resistivity of the rocks constituting the ancient basement of the un¬
derground terrain. Laboratory measurements on samples had revealed
much higher values than those for recent alluvial sands and clays. Fur¬
thermore, it was known that sandstones are harder than shale and marly
Jurassic limestones. This fact was confirmed by the outcropping of
sandstones south of Sassy (Fig. 8). Shales covered with a thick limestone,
on the other hand, showed that during a certain era they had been more
intensively eroded. The Armorican sandstones, in particular, formed a
slight, east-west oriented hillock, at the foot of which ran the trace of
their contact with the limestones, and, below, their contact with the
shales. Therefore it could be assumed a priori that such differences in
hardness would be reflected electrically, and that the resistivity of the
sandstones ought to be higher than that of the other formations. Unlike
the Val-Richer experiment, the one at Sassy involved a subsurface com¬
posed of several formations with different resistivities.
The measurements were carried out over a few days in October, and
the dependability of the technique was confirmed. In Conrad’s experi¬
ments, the position of A varied, whereas B was fixed at a remote location.
The tests showed that the equipotendal curves no longer appeared as
concentric circles: they clearly reflected the subsurface formations and the
locations of their contacts. To analyze their arrangement even more
accurately, Conrad resorted to a method that geophysicists were to employ
in the future: a model composed of several formations was devised. It was
simple enough to lend itself to mathematical treatment, and similar

21
\£-CXC-L£^U^ £y f\Ji t • £-t <■ C*, . ^ s/t\ , h(X-dx, 0*.L

YaJ. - Ph'c/vi* f CAl^/^h^ J . /hCJ- /Jt 2_ ^

£cbt c£e -7

fUM-~'t~y>-

/
Oe jyiifa***. $C.g.*«y,.*f*/ S^:.?J&.‘f..r? X^', _g.4—A—-A? J1AÀ1-W*—, ,
^'i»û j Jo*. aad-J.«-A • 2 efan*iye~f y_«< /t»^^*f***AU. ï ' a-i-* /{**. *,. •t' « '"if1'- “—4 fn.i** *»»-'' 1 <•
f j. *£, ^«A- ^ -X4U.A &**»ifae*/t' - * fa*-» $■ /# *" •**■ •»**• , *■«»«« •/ fa*3 <î *,/„ r^. //,/^ . "J a..-{/*,.
fa aJ~ ti*- Url» rt**j~A\a . fixa*. /> *faa»-*a* 4 fm/*- fa- fi 'a-i d* /w. .•-<■ ••- — --■/■ (5 <£>/» a*-*-* fa*. n..». *» .« . <J <i A-».*. X*/C.
^6 ( f'ffa* "bar** c«^> 4--- ' —^ • A/; i*-
Crt*f**» '-/-• fa*. <fa -L fa*.^
t ' fa*.
' •'■ .. faytA>J~ ïm ^ 'ù»4ufaLfa.- <*/*a't.m4rt
-
]J. OaU2 ^TfaA******* fa^**». , /£> fayrt- £ * r*d£**.X J i— ^ Z~ Ca*-ét» . 7U, r» . « aA fi £J\* {^ ,/tt*J~
£ C*f**.d* ) j fa*» fat* c***J» {fax*»**J- <r*fa ^ A> ti./, / ftt/ Aa c**»Jt ftu-»/ ÿ /Vi **.»& 'Vi Pi'xadtfa. hm*'^ d*du/»
^1 £—Vt^U A CAO-—, / AæAA» Aw/u y ^AAi U«t^i t ( U4a^ M {*-• ^1 4- In A J fia*» “* ^ cada
twj, i.^ ^/4y>^ <^» £«•' ioi , e**fax»nn d» a ** Cu ^ ka— /a/ia*a» ^uiî *Wru*^ ^ .
Toad»*** ^tu t1 I ., ^ C»u»-,A/«iu« 1,-v bA^^j aa, AW., . A f *.., -/» ■ *- AW* .
4)^ . Mjl t ft —V,. i/* '/A«—. /i Ai^yL, /* vfaj/at. »~ »t**jr, ( ’.At, _ *J? /. fa/t/x*r»-Xn.
' *''-^
faj. ( »v fa»!n*** <■ fay^tdS, di/tfa.» d* JtMi t UAA-: *-« « d'i*»e»X* ^-~.»^A~ {fa-* /wfe^X.. WXaV4^( **r-4 jfat+M
Û H—^itAAA , S» C fa-t-t. Aav*.Am w '-^-/ IVU,A^—<Xa- /",<A *S- far
rùr-ir» t<A*’ fca—*~faia , fa*-*, f.** ' *J **'y n fal*. . (iw. </"£ ' at-**» {fact 4**** .

°*+*i e»'- / sulctfa ***. farL. a***, fx^ <*4 CMt^xfa ->r*S f e//», Ze /^JT
Jf ‘fa‘S*fa* (l*P—... ) . -4 Cr-I**fa >~t* fafaJcfa fat. A WVt^icfa re/) A**-Sc*0 » 4 /*V«fa
& far-flf AC 4 Cn<fal r*X ***r W fa*. *. fa.Xt*/m*tA ^ t, fal~** A g)f ^ f'catfa A t^Z> , fa**,
-fa O-C , /r»‘e. tK«i, *-*faXe»J\-*v, /rrv, c~*fa dc fax»»S* J ■ ft»*» 6» e*2 J» 2) , & 4«A *. A- Cnt»£L.Jt", #»v-»
4 cA**^ct* crix***. £■**» ^/iAa uaTAi ( Awa-4 *-xX fa»*4-4* ^ CAs\ cut" ^ ^ cc\* -$/* A*, i***» s*Y4» /—"fa*
Ml/’ 'I II J ! J~t" 'Z / 'H*^v*‘ fatn**. C» cr^-t*jk*S~ . ^ e*n*-Si, If**» far*- . J~n* . <A« 1a/a/V< , — a * A^, < n »V
^ . Vt*^t {jt 'u**/y*»-*4' fav+»**' fa»». &. /Mti t 4 fifS. 4* crvtnot** 4 <»*-, ^aiAa*^ w
fUs+**Vu»»J~ /* 4 d±**r,1t fa crx*t»tx*r O*..y», i_»« «4 x t!$ el**./» 0^i***4-i /"**-* faeh-' UuAi <»4y.., /.,/, ^
hy*»X tf»» {fa *l£*4^ €^>->r**J~. A Ctri»' fa* cfa~ *fafa*4-, . * 4*. y — A/ 4 A a.,
fl*». d» , ■< fair~L* it»riJ* fa £XJc' ffaM**. fat*.—.»£*+' 4a**-. {t *nrC. / i*a_f* d* 4 ftv{a ,/ An... A vr./t*» J.
I X^fatfaf^Arx. C *^r ***** far, fa 4v*y —y ti-afai»^**^ fa t *- <• fttJdt. ) . Za C*n~J.
% tfa{r* A** Yxtr **» far*i*J- n»S'rm»-*t f*~, tj <*n**j»*t4~ e*4 An*** 0 j yo * , 'A-
-fain**X i. it*. £rvn— f i**-, » 'mW fa / 4 2 *** i***4\a ^ 2 . *—/ ** ' tj, 4*<Ai (\ t < *» '?.<*-*
• **■ •/*’” / JlfviSVrv. a f a.» *i* t i *tA^a*»4 f~ ’n. t~+ fa*. »y *' t t.fa*.cfa <4
^ /“* ^/*****^M fan»» fa xU W- x*^*fa-u. fact./ f*** «' ^ *♦• A'ntrt*. fattX ~*4 i <5»
-f, JfatXJin. **4- l*»*£it *4 e fa H» ✓wA < ?«fcrv*AA £, it i /*■ * n fatnh & Ma**J* it, it../ m.. A4A # t\yi*** 'da st-

.->>■ ^4n..h fa**, '-n - - x‘ {*-» G, /fa 4 Jc*h cfann Aa***k da, i"rt**A», a»**. X*
fa d** irfarx*. fa»**A da uaruxa. fc**» ^ dsi*fji » C t* *^r- a**J\* A, {**- r
4» An Jt**4- fat*» Aj«— / a
t,« l* % fafa/f CJtJt**- dt *4
ÿn!» /» ' ^ dldaJtfa*'
n* cri t**i Ae***, t fai.'
£n fa fatarr, o/xfa y ft4. it i »W fa»».»***//».»*-*.i ^/- /t**j»6, *

Figure 6. Experiment at the Val-Richer.

23
24
UA^C CAO, f

fa* (faut <?fauZ « On^cJui caA^Uaa,

de

~ 01-*. cy>My\^ (rC-J~ ^C*ji*yiAA) y&otA&A. ctAt-ty^

A'*

fàpi-. c «V- <£ u^ILefa^ <Pr'*^*ÿ‘'Ç*^tJ>U <£ U*+l/4AtjU / KtxaJT & i*-tM^tW+-* a /j+'
tJ-i^J-uâ. &+*■ V~&4-~ . £jxeu+Ji&> CüuaI^ ejo . (e{/A**Jx*'U++A J 4* dj fi*x+ iU‘* J-y
(nsi dcust IcasxA*^*-* (1 ) W.1^* Vvi Jbtat: t^' Jxi*-^. vr-MtM J)

fi**A (1 ) . 4 fa+tyjj/V **+- 4 f*À />•“ ''z qV) / ^


oJJûucha^J' «> >*»->■ ve+*, v*++rf' ; 6 (?/*-* 'uf*x^M /*-» <Au c+ie**l /' V ^
I Uaaa^Ya^a Liy^yVxJ*S ' a-'JCf, ^ At^rwt, f <»* *-»—*-» l~*A^vx^S*^x. fyx*- uj Ua'b ' * '/, ‘ '

$ - - - -1- - * /"a-. - A/ « /^uuhAti- 4VTm ^ iA u/AalJj f^u >Miyu< >»t-> ^


Ca-iée J ^ S, 4 ca£*-ji^J uji, . Ct i CiU “FLa »

^ c ***~ t zty^L^n A«4 *


'■£t t* liutu. (. {^ i+/- c*i**k K»*! /*-» «-m-^
/I ^Ü»i *~-A ^7 U JlAAA^/(+ ) 6f *tt, b « ^V. 4 ^4 Z /Uz«^ /*«• ^46.4 o_> '<i7> 1 c*iHL> (*f 1
Ùtt/UA*ujaa}- .
feJzZ /im t &v4Aa< ûàiM4 i-**c*
4tJv*AA*~ CV! fo-x-^-^a !-<-x+ >m4A-<m/»v. ^fuwut.y AcoV***^' 1

^ U-^t- 'c tA*A Caax^aA^ C^Vtt^itt^t,«


^ A ü» Aai ^
4 £*-."£-<• tVvtVL-t-A S^aU* c&„ f*i,
1. j kl
I#
4 A4m< ty,*Xi(**-* 6&+~<A , â| r^n «-, |’|.4>|/,. yfj 1 {, [ i !
liiWIII z
At, &. C*\~th~. e-A C~£-A* /£, (i) x^*j(j-f 4 «a. M+ aAj cC ( t ) gj_
&, Jx^i.*. <4 ^ »*• /Srv*v ^ 4^» fan' * *• f+*4 4 Cn^l^ ^4 . Wtof
Cti cUlft^A -*i ^ •^, +f!e^* Â, /tA'A, W- Cf, ttc/ct^vs zt-4,,,
/4^ $Ù~4 CcujiÏ-c, fu*—A ^ hfAifaity+i & Ci^xj, ? J <0 JtfasxAJvtn*
4 csl+aM ^WiW-iA ^0 i*/u^4.tA 4,' 4 ^ ^
*Tat*> Maa^, (1 ) Jfi» 4»*-* ‘<>j^ ^ 7 T - e/fy ,>vuA- caà, /.. Z.,^
^ /a^t’ </^ OtttA. (^ r^U+Aj, 4l'«n.^ Af Mm'9%^ ! (IJ
~*^hà- ,“■' ***7^- » ■ ?«*, ^ a. c^A^.
—-4-A^? ^/*•" *“ 4 4/ «S • ^
kaix* ^T^*) o***»4aiztAiA4.ttt4c/', ‘1 ^ ^ A^cV e^~
y^u</rx. P~4Alw i^4
^1*£u*aJ vr^/^-X*, ^^jU^J- 1r^, c^UA^U^^J. AAA^.

« ^ .4rf^^4« a- 4»}^ ^., ,^,'SZT 4 >44c*^

<rxy £ 1//xJa+~ J*As4-,~\ Cyt-A^ ^l, kvvA ^It-t, ^»U f tr^ a ttt< ’'•'.Li^t/jJ .

«-* /*^aa
/* *** /'*' * ^ ^ -U^AcZJr., 4^ c+x^ff^ f d
exJ^rx* _ a“ . & /:- ‘'j f . /',-Z'/^«.
*V4 ^i- A-- ^ A-^-ttZ 4 Z. 16^. -« - /_ _ /f Æ ^ ., . ^ ^ Cfl^f *» At«^/ #/4^L /Uxx.

>-a_ =, .err <-


!tZZ£
<» Ovc-A^
SA-ryU. «^TTCr
iC -<rv^ 4*t, ^tAtcZ,
__ ^a^aTm.:^
•VlA^U **<*' M, cl

Figure 7. Experiment at Sassy.


25
The History of a Technique

Figure 8. Experiment at Sassy: enlarged detail.

enough to the actual structure of the subsurface so that a plot of the field
measurements could be unequivocally explained by means of the cal¬
culated model. In Figure 9, the drawing on the left shows the calculated
equipotentials, separated by a vertical plane at the contact of two
homogeneous formations, one resistant, the other conductive. The sketch
on the right refers to a conductive medium between two resistant ones.
The curves plotted on the map appeared quite similar to those of the
drawings. The large equipotentials were elongated in an east-west direc¬
tion, that is, alongside the trace of limestone and shale layers, which are
softer and electrically less resistant than the flanking sandstones. Finally,
the breaks they showed at the contact between sandstones and lime¬
stones provided experimental verification of the refraction phenomenon as
predicted by theory.
Thus the potentials map can be informative even in the absence of
metallically conductive ore bodies, inasmuch as it can reveal certain
features of underground geological structure, such as the strike of the
strata and the contacts between formations. It is true that Conrad was
experimenting at Sassy on a known structure, visible over almost the
entire surface—a condition necessary to check the accuracy of the
method; yet it was becoming apparent that similar studies could be at¬
tempted in regions where the deep structures were completely concealed.
Such was the case in Normandy, where iron ore exploration was based on
the identification of contacts between generally hidden sandstones and

26
Surface Electrical Prospecting

Figure 9. Left: Calculated plot of the equipotential curves around an electrode A (with B very
remote, practically infinite) and close to the vertical plan separating a more resistant medium to
the left from a more conductive medium to the right. The curves are being refracted when crossing
the contact. Right: Approximate image of the behavior of the equipotentials when A is located in
the middle of a layer, the conductivity of which is higher than that of the flanking formations.

shales. At that time a survey could be made only by drilling coreholes, and
it appeared logical that the potentials map could complement the data
obtained by drilling, if not become a partial or total substitute for it. After
a few weeks of tests, prospects for industrial application were in sight.
The third experiment took place in January 1913 at Fierville-la-
Campagne, in the Calvados iron ore basin (Fig. 10). The composition of
the subsurface there is similar to that in Sassy, that is, nonoutcropping,
very steep sandstones and shales totally overlain by 60 to 90 meters of

27
The History of a Technique

M
Jurassic 'I. T_ i—i-'
~r-~i
limestone X '—I— -i 1^5

Feldspathj rian
sandstones black shales
Silu
Armorican sandstones
LC!1 '^^^'^'May sa
sandstones

Calymere shales

Figure 10. Test at Fierville-la-Campagne, Calvados, 1913■ (Reproduction from C. Schlum-


berger, Etude sur la prospection électrique du sous-sol, Gauthier-Villars, 1920.)

28
Surface Electrical Prospecting

marly Jurassic. Seven drillholes were required to outline the ore body and
thoroughly explore the formations. As in Sassy, the strike of the ancient
strata was clearly shown by the elongated shape of the equipotentials
surrounding electrode A, which was planted straight above the shales,
while B was placed a sufficient distance away so that it would not influence
the shapes of the curves. However, the contact between sandstones and
shales appeared less clearly, because of the blurring effect of the thick,
overlying formations.
The flexibility of the electrical method was now to compensate for
this lack of precision. As shown in Figure 3, homogeneous ground pro¬
duces very uniform potentials in the center of the drawing: the electric
field, that is, the potential drop by unit of length, remains constant. But if
any discontinuity should appear in the soil between A and B and separate
two beds of different resistivities (e.g., the contact between shales and
limestone strata), the electric field would vary at the level of this separa¬
tion; it would be weaker above the more conductive formations, and
stronger above the more resistant ones. This led Conrad to test what he
referred to as “field profiles.” Electrodes A and B were placed on a line
perpendicular to the strike of the strata, each about 1 kilometer away
from the presumed trace of the contact, and the electric field between two
very close electrodes, alongside three profiles parallel to AB, was mea¬
sured. The result was clear: each profile showed an abrupt change in the
value of the field, making it possible to locate the contact within an
accuracy of 50 meters.
Another new feature was apparent on the right-hand side of the map:
the equipotentials surrounding an electrode planted directly above the
May sandstones were similarly elongated, and ran in the same direction as
the curves plotted above the shales. Therefore the current from A flowed
more easily alongside the strata than across them. This occurred because
May sandstones are composed of thin, alternating layers of varied resis¬
tivities, affecting the current with what is known as an anisotropic
medium. The curves plotted on the map confirmed the theory that in such a
medium the equipotentials are ellipses elongated in the direction of the
strata. This was the first empirical demonstration in the field of the electric
anisotrophy of sedimentary layers.
Somewhat later, during 1913, the Soumont survey was made in the
same basin with the same Silurian structure hidden by Jurassic. During
that study a practical mining problem arose: the ore body under produc-

29
The History of a Technique

tion was obstructed by a fault that had thrust its extension outside the
working face of the heading. It was decided to use the electrical method
for relocating the ore.
After having plotted a few large-diameter equipotentials and several
field profiles, all of which confirmed the conclusions reached at Fierville,
Conrad tested two new variants of the electrical method that enabled him
to determine the slippage of the fault. These procedures lasted for several
days and kept the small team of five very busy. Eugène Léonardon, a young
graduate of the École Polytechnique and one of the first prospectors to
join the firm, was a member of the group. He and a young worker named
Jules Carré were to become popular figures of “La Pros.”
The Soumont success, which Conrad still called a test, in face repre¬
sented authentic mineral prospecting. It was later followed by another
study in the same region near Sees, Orne, from which additional valuable
information was extracted and in the course of which the techniques and
analytical processes were refined.
By the middle of 1914, before World War I interrupted his work,
Conrad had succeeded in demonstrating that physical measurements
made on the surface could provide useful data for mineral prospecting,
even when the ore could not be detected directly from the surface. By a
thorough study of the structure of the ancient sedimentary rocks of the
subsurface, it became possible to track down deposits of geologically
related minerals. Although at the time only iron ore was involved, a few
years later geophysicists all over the world would apply the same principle
in their search for a mineral that had eluded any instrument designed
solely for surface exploration: petroleum.

The exploration of conductive ores

While pursuing his experiments in direct prospecting of high-


conductivity metalliferous ore bodies, Conrad had noticed that, after the
current sent into the soil was switched off, small but measurable potential
differences continued for some time. It seemed to him that the ground as
a whole was polarized by the direct current and then discharged. A local
incident led to a comment that he jotted down on the potentials map of
the Val-Richer: straight above an iron water-supply pipe buried close to
the surface, the equipotential curve had moved approximately 2 meters

30
Surface Electrical Prospecting

when the current was reversed. He reasoned that, because the pipe had
become polarized, it affected the current differently, depending on the
direction in which the current circulated. From this observation he as¬
sumed that a metallic mass, or a metal-like conductive ore, could be
distinguished from the surrounding rocks by a particularly intensive
polarization. A new prospecting method became feasible: send direct
current into the ground, interrupt it, observe the effects of the polariza¬
tion, and then look for the spots where they were abnormally high.
The idea led to little experimentation at the time, and when it was
taken up again at a much later date it was not intensively pursued because
efforts were directed toward new projects that seemed more attractive.
By 1913 the induced polarization tests3 were sidetracked after the
discovery' of another phenomenon that was much sharper and easier to
exploit. This discovery would lead to a prospecting method widely
utilized in the geophysical industry: spontaneous polarization.
In April 1913, near Sain-Bel, Rhone, Conrad was developing a
potentials map of an area above a pyrite lode in the shape of an elongated
lens and about 100 meters deep. Before any current had been sent
between A and B, he noticed, in the vicinity of the ore, potential dif¬
ferences between nonpolarizing electrodes spaced less than 100 meters
apart. These differences reached hundreds of millivolts above the axis of
the lens (an amplitude substantially higher than that usually observed in
such tests), and faded away as the electrodes were moved farther from the
ore. Multiple observations confirmed the phenomenon: the measurements
would accurately reproduce themselves. A new type of map was drawn
(Fig. 11); with potentials becoming negative directly over the lens, the
map not only indicated the outline of the pyrite already mined in the area
but also showed the location of an unknown lens whose existence was
confirmed a few years later. This was an instance of a spontaneous phenome¬
non occurring in the ground without any artificial stimulation. The pyrite
lens acted like a huge metallic mass, whose upper parts had become
oxidized by the circulation of aerated surface waters,4 while its lower parts
had escaped alteration. The pyrite lode, once polarized, generated electric
currents that flowed around it from bottom to top, creating a center of

3 Induced polarization has now become an efficient process, widely used since 1950.
4 Such circulation and the resulting oxidation are enhanced by the seepages into mine
works; they can be sufficient to generate a measurable phenomenon for a virgin lens,
provided that its top is close enough to the surface.

31
The History of a Technique

negative potential at the surface above it. The validity of the assumption
was demonstrated through imaginative laboratory devices and mea¬
surements. New studies at Vaux, Rhône; Saint-Félix-de-Pallières,
Gard; Herrerias and Campanario, Spain; and Bor, Servia, all confirmed the
Sain-Bel observations and demonstrated that spontaneous polarization
(the term was eventually used to designate the mineral-prospecting pro¬
cess itself) characterizes all formations containing pyrites. As a result of
the survey made there in the early summer of 1914, the Bor copper mine

32
Surface Electrical Prospecting

was expanded considerably. This was the first time in history that
geophysics had been instrumental in the discovery of a nonmagnetic
deposit.
During an earlier experiment at Bor, in September 1913, Conrad had
succeeded in discovering a variant of the artificial current process: if a
transmission electrode is planted in the mass of pyrite inside a mine
heading, the whole ore body is raised to the same potential and becomes a
huge electrode, accurately outlined by the equipotendals plotted on the
surface. This so-called “grounding’’ method was subsequently used on
several occasions.
By 1914, in the limited time available between his obligations as a
professor at the École des Mines, Conrad had been able to advance the
research work far enough to have in his possession an operational tool
that had proved of value in several industrial surveys. However, he
considered such studies to be of an experimental nature only and had
sought no compensation from the mining operators who had benefited
from his work. He continued to regard his experience, although encourag¬
ing, as too narrow for expansion into the commercial field. Only after
World War I, in 1920, did he publish the first work describing his
methods of electrical prospecting of the subsurface.5 But he had been
prudent enough to protect his inventions, and his first patent, Procédé pour
la Determination de la Nature du Sous-sol au Moyen de l’Electricite (“Pro¬
cess for Determining the Nature of the Subsurface by Means of Electric¬
ity”), had been filed in France on September 27, 1912. Several additional
patents were to follow.
Out of a belief that nothing should be left unexplored, Conrad
meticulously investigated the findings of other scientists and
scrutinized whatever published material was available. He gave special
attention to the experiments of the Swedish researchers Gunnar
Bergstrom and Carl Bergholm, who had worked on the detection of
metalliferous deposits through the transmission of current and the plot¬
ting of equipotendals by the earphone method. They had experimented
for several years without producing any conclusive results, and only after
1914 did they contribute effectively to the field of mineral exploration.
Moreover, it was a long time before they could achieve the accuracy

5 C. Schlumberger, Etude sur la Prospection Électrique du Sous-sol, Gauthier-Villars et Cie,


Paris, 1920.

33
The History of a Technique

assured by the use of direct current. Additionally, the phenomenon of


spontaneous polarization was unknown to them. In this respect Conrad
called attention to the work of R. W. Fox, described in an article in
Annales des Mines (1837), and to the writings of Carl Barus (1880) and
R. C. Wells (1914), whose field and laboratory observations he found most
instructive. Yet nowhere could he find any account of their application to
mineral exploration; in fact, in 1914, the director of the U.S. Geological
Survey, George O. Smith, wrote in a preface to a work by Wells:

It should be emphasized that the results thus far obtained afford no adequate
basis for any method of electrical prospecting nor any promise of the
development of such a method which would make possible the detection of
an ore body by electrical measurements.

This was a few months after the Bor discovery. In all fairness to the
director, however, it should be added that information about this discov¬
ery had not yet been published.

34
Phase Two: After 1919

Technical Developments

An overall view

During World War I some progress was made in the application of


physical measurements to the search for minerals. Advanced mag¬
netometers (inclinometer and Schmidt balance) became tools for mineral
exploration. In addition to the contributions of the Swedes in the field of
electrical exploration, Frank Wenner, a physicist of the U.S. Bureau of
Standards, announced in 1915 the development of the four-electrode
device bearing his name, for the measurement of formation resistivity.
The possibility of using the Eotvôs balance to locate potential oil-bearing
structures had been recognized; however, thorough experiments con¬
ducted on salt domes in Czechoslovakia (then a part of the Hapsburg
Empire), Germany, and Rumania had not resulted in any discovery. As to
seismic methods, 5 years in the German artillery gave Ludger Mintrop the
opportunity to meditate on the propagation of shock waves in the ground,
but only after the war was he able to take out the patent that became the
basis of the refraction method. Similarly, Conrad Schlumberger, then a
French artillery officer, meditated on his side—whenever a lull gave him
the opportunity—about his prewar studies, to which he could relate
certain features in the military technique of ground telegraphy.
In 1919, with World War I at an end, Conrad decided to resume his
work, placing a heavier emphasis on its application to industry. He
teamed up with his younger brother Marcel, who brought with him
practical experience in the mining business, as well as a creative mind and

35
The History of a Technique

a talent for mechanics. The two brothers, with the moral and financial
support of their father, established the Company that bears their name. In
1923 Conrad resigned from his teaching post at the École des Mines to
devote himself fully to the new endeavor.
The Company premises at No. 30 rue Fabert in Paris consisted of five
rooms. Three were converted into offices, while the other two became the
workshop and the laboratory. Surrounded by maps, diagrams, and charts,
the two brothers shared a desk, with a drawing board nearby. They
supervised every aspect of the operation: conception and design of
methods and instruments, manufacture of equipment, seeking out cus¬
tomers and drafting contracts, purchase orders, analysis of supplies, and
invoicing. They directed and monitored the work of what they called the
missions: the crews (usually consisting of two prospectors and a few
assistants) in charge of carrying out the field surveys. For years they
occasionally took part in the manual work required for the measurements:
handling cables, generators, and electrodes, reading the potentiometer,
and so on. They interpreted the results, discussed them with their clients,
drafted the reports, drew up the budgets, and controlled the expendi¬
tures. Frugality was the rule.
In the beginning they had the help of a single engineer, Lëonardon,
who had rejoined them after the war. In addition, fieldwork had the
unflagging assistance of two young men trained in vocational schools,
Pierre Baron and Jacques Gallois, who were to have long and brilliant
careers with the Company. In Paris J. Carré was the storekeeper, and
R. Jacquin the mechanic and driver.
Thanks to the eagerness and dedication of its initial team, and also to
the growing interest of the French mining industry, the small Company
was able to expand, despite the unfavorable conditions of a period of
economic reconversion following World War I. New collaborators gradu¬
ally joined Conrad and Marcel. In 1922 Edward Poldini, a young Swiss
geologist, became their part-time consultant. Priority was given to the
recruiting of graduate engineers with university-level training. Mission
personnel was provided mainly by the various écoles des mines1 and the
École Centrale,2 with top consideration for grades received in geological

1 Graduate institutes of mining engineering. (Translator’s note.)


2 Ecole Centrale des Arts et Manufactures, a multidisciplinary graduate institute of
technology. (Translator’s note.)

36
Surface Electrical Prospecting

and electromechanical training. The Écoles des Arts et Metiers3 and the
Ecole Supérieure d'Électricité4 provided the candidates for the engineer¬
ing office, but there was never any sharp discrimination in this respect. In
fact, prospectors available between missions could be assigned to labora¬
tory work or instrument design. In any case recruiting in the first years
drew on widely diversified backgrounds, from the Institut Agronomique5
to the Ecoles des Maîtres-Mineurs.6 In the early years of the Company
several employees were doubtful about its future and left, but in 1924,
when an expanding world economy brought in long-term contracts, espe¬
cially in Rumania and the United States, and stabilized the development
rate of the Company, some of them rejoined. Also, new staff now joined
with confidence. In the summer of 1926, after a few months of part-time
collaboration, Henri Georges Doll, Conrad’s son-in-law and a graduate of
the Ecole Polytechnique and the École des Mines, joined the Company.
With his genius for solving electrical problems, he soon clarified and
improved measurement procedures. During his long career he was re¬
sponsible for a continuous series of new developments in both instrumen¬
tation and methods. Félicien Mailly joined the Company in 1925 and for
20 years was the mainspring of its electrical research laboratory.
On July 1, 1926, the Société de Prospection Électrique (S.P.E.) took
over all of the Schlumberger brothers’ operations, including personnel,
equipment, and premises. In Paris and in the field as well, everybody was
already calling the enterprise “La Pros.’’ One year later the staff had grown
to 16 engineers or equivalent technicians.

The concept of apparent resistivity

Two essential elements contributed by Conrad and Marcel enabled


“La Pros” not only to survive but also to expand its operations: the
concept of apparent resistivity, and the potentiometer. At this point it is
appropriate to describe briefly the methods and techniques that evolved
and were developed during the 1920’s from these two elements.

3 Technical schools emphasizing the practical training of mechanical engineers. (Trans¬


lator’s note.)
4 Graduate Institute of Electrical Sciences. (Translator’s note.)
5 The highest agricultural institution in France. (Translator’s note.)
6 These schools produced originally mining foremen, but were upgraded to schools of
mining engineering with emphasis on practical training. (Translator’s note.)

37
The History of a Technique

To this day the notion of apparent resistivity has remained a basic


concept in electrical processes used on the surface and inside boreholes.
Although the potentiometer has been largely replaced by the photo¬
graphic recorder since the 1940’s, it is still used in certain kinds of surface
work.
Before 1914 all the tests in which current was emitted into the
ground had consisted of plotting curves of equal potential. The electric
field profiles were nothing but a variant of this method. The surveys,
resumed in 1920 for the reconnaissance of ancient, sharply dipping for¬
mations, showed that the potentials map technique was limited in its
application. First, the electrical field becomes very strong when mea¬
surements are made close to the electrodes (Fig. 3); this phenomenon can
distort the interpretation of the map when the medium is not homoge¬
neous, as is almost always the case. Second, when the survey extends over a
large area, a single map of potentials, referred to (stationary) electrodes A
and B, becomes inadequate. The measurements must be repeated while
moving line AB, and it is no easy task to link such a succession of readings.
Finally, since the differences in the voltage between successive equipoten-
tials are proportional to the intensity of the current transmitted into the
ground, they reflect, with only approximate accuracy, the resistivity vari¬
ations. They do not give any absolute values, even approximate ones, and
therefore are not comparable from one area to another.
At this point it was decided that the reference parameter would be
called apparent resistivity. In a configuration of four ground electrodes
ABMN, where AIN is located between A and B, a current is emitted into
AB and the corresponding potential difference is measured between A1
and N. The apparent resistivity equals the ratio between the potential
difference and the intensity of the current emitted, multiplied by a factor
that is a function of the respective distances between the four ground
electrodes. This coefficient is easy to calculate and is such that, if the
system is placed in a homogeneous medium, the apparent resistivity must
be equal to the resistivity of this medium.
The apparent resistivity reflects an average value, depending on the
resistivities and spatial shapes characterizing the various formations in¬
cluded in the total volume of the subsurface being measured.7 The vol-

7 The word “apparent” is often omitted in the professional language without any resulting
ambiguity. “Resistivity” designates a precise parameter, which characterizes certain mat-

38
Surface Electrical Prospecting

Figure 12. Measurement of apparent resistivity (schematic).

ume itself is not sharply delineated, because the density of the current is
not suddenly reduced to zero at a certain depth, but decreases and
disappears gradually in all directions as the distance from the electrode
system increases.
Figure 12 shows a current being emitted along the line AB and, in a
vertical plane, lines of the current expanding as a function of depth. These
lines radiate around electrodes A and B when in the zones close to them.
If electrodes M and N are close to A (Fig. 13), the voltage difference of
two rather shallow equipotentials is measured. As the electrodes move
further away from A, the equipotentials become deeper and deeper. In

ter (e.g., a rock) when measured in the laboratory on a representative sample. The same
word also designates apparent resistivity, that is, an average, when the measurement takes
place with an AMNB configuration stretched over the ground.

39
The History of a Technique

Figure 13■ Variation of the depth of investigation according to the position of MN between A
and B.

other words the depth of penetration of the system is in direct proportion


to the remoteness of MN from A. This depth reaches its maximum, then
remains practically constant when MN is within the central third of AB,
and decreases again when MN moves closer to B. In the central third the
equipotentials are nearly vertical surfaces (the current lines being approx¬
imately horizontal). Originally it was assumed that, when MN is within the
central third of AB, the slice of ground below AIN to a depth equal to
ABl4 had a prevailing action on the value measured.
Thus the depth of investigation became quantified. Granted, this was
a convenient and provisional approximation, reasonably correct only in a
homogeneous medium. It was recognized that in real formations consist¬
ing of layers of different resistivities the distribution of the current lines
would be more complex than indicated in this ideally simple case. It was a
combination of empirical evidence acquired in the field and scientific data
computed by mathematics and laboratory research that gradually defined
the relationship between apparent resistivity and its various components,
and increased the accuracy of depth of investigation.

The initial system consisted of a line several hundred meters long.


Electrodes Al and N were located 20 to 50 meters apart on line AB, and a
series of measurements was made by displacing AIN within the central third
of AB (this was called the AB profile technique). The cables connecting A

40
Surface Electrical Prospecting

and B to the source of electric current were laid on the ground. Once a
profile was completed, the AB configuration was moved either sideways or
in the same direction, depending on the type of problem being solved. In
the latter case, care was taken to achieve continuous measurements. A
slightly more complex method first used in Rumania (hence the name
Rumanian method ”) consisted of making four series of measurements
aroundA by moving MN into four different azimuths for each position of
AB. The same was done around B. It would be tedious to describe other
configurations used, though they clearly illustrate the difficulty of selecting
the best technique.
Later, a more expeditious method was adopted. It employed a regular
network of measuring stations with the same AB spacing, while electrodes
M and N, always on the same alignment and within the central third, were
symmetrical about the center point of AB.8 Thus all measurements were
made with the same average depth of investigation. The stations were
equidistant from one another and positioned alongside alignments as
straight as possible. The AMNB configuration stretched along the same
alignments and was moved from station to station by a distance equal to
MN, thereby resulting in adjoining measurements. Since the operation
required that the cables be dragged on the ground, this resistivity profiling
was called the “dragging” technique at “La Pros.” Until about 1927 the AB
spacing never exceeded a few hundred meters, and the rather lightweight
equipment could easily be carried on a pickup truck.
Resistivity profiles were used for the first time in the fall of 1920,
during a survey of an iron ore-bearing syncline at May St. Andre in
Normandy. The problem was similar to those that had prompted the
above-described tests in their early stage: reconnaissance of tilted Silurian
formations, predominantly shales and sandstones, concealed by a nearly
horizontal Jurassic caprock of marls and limestones 20 to 40 meters thick.
However, the area to be surveyed was substantially larger. The survey
involved a series of nearly parallel profiles, about 2 kilometers long, each

8 This configuration was very similar to that mentioned in 1915 by Frank Wenner, who
had introduced the term “effective resistivity” to designate the same concept as “apparent
resistivity.” The latter has come into common use. In the Wenner configuration, however,
MN is always equal to one third of AB regardless of the length of AB. Wenner’s
measurements were limited to quite shallow depths, and he operated with very short lines
(a few tens of meters), using alternating current. However interesting, his studies remain
mainly theoretical.

41
The History of a Technique

with a south-north orientation, that is, perpendicular to the general strike


of the Silurian. The depth of investigation was such that the shape of the
Silurian formations was clearly revealed.
Figure 14 shows one of these profiles with the corresponding geolog¬
ical cross section: peaks appear above the sandstones, troughs above the
shales. By correlating the peaks of these profiles, it was possible to plot
the trace on the ground of the various formations hidden under the
Jurassic caprock.
Shortly thereafter, a cartographic representation, clearly indicating
the values plotted, became the simple complement of the profiles. The
resistivity values plotted at each station were transferred to a topographic
map, and curves of equal value were drawn by interpolation. The result
was a “resistivity map” where the equal-value curves showed the limits
between various zones of apparent resistivity. The zones were marked by
conventional colors: red for higher conductivities, blue for higher resis¬
tivities, and so on.
When the purpose of prospecting is the reconnaissance of steeply
dipping ancient formations overlaid by a caprock, the AB spacing is

Resistivities
(ohms m2m)

1 syncline 1 200 m

Figure 14. First resistivity profile (May St. Andre, Normandy, 1920).

42
Surface Electrical Prospecting

adjusted commensurately with the thickness of the caprock, so that the


resistivity measurements clearly reflect the formations being surveyed.
When the caprock is rather uniform, the resistivity map looks like a
geological map plotted as if the caprock had been pulled off; this is called
“geological skinning." It is important to remember that in such cases the
resistivity map reflects only rather gross features and gives a more global
or less detailed description of the formations than does the geological
map. On the other hand, the method has the advantage of spotting
elements invisible on the surface: faults, stratigraphic contacts, and the
like. For the Normandy synclines the result was a sketchy outline of the
upturned Silurian strata at the beginning of the Jurassic period, when the
surface of the earth had been eroded into a peneplain, and an invasion of
the sea had begun to bury these strata under new sediments.
The application of the resistivity method to the study of low-dip,
nearly horizontal formations came a little later, in 1923, under the stimu¬
lus of oil exploration. At one time it may have been thought that oil could
be located directly through surface resistivity surveys; but this proved
impossible since, in spite of a volume reaching millions of cubic meters,
oil deposits represent only a minute fraction of the formations containing
them. On the other hand, it seemed reasonable to expect that the resistiv¬
ity map would furnish clues regarding the position of the subterranean
structures suitable for the accumulation of oil, and thereby provide indi¬
rect guidance for its discovery. This was also the objective of the
gravimetric and seismic methods, which at that time were undergoing
their first tests.
Figure 15 represents schematically a hypothetical vertical cross sec¬
tion of formations across the axis of an anticline. The porous and perme¬
able layer where oil may have accumulated—a potential reservoir—
belongs to a very thick sequence of mainly argillaceous sediments with a
rather low resistivity (about 5 ohm-meters), capped by formations of
higher average resistivity (30 ohm-meters). Since the surface is approxi¬
mately horizontal, the thickness of these formations above the top of the
anticline decreases. An AMNB configuration of appropriate length, when
moved on the surface in a direction perpendicular to the axis of the
structure, will record a minimum of apparent resistivity above the top of
the anticline, where the thickness of the resistant formations is lowest. If
now a series of profiles is plotted at suitable intervals, the result should be
a resistivity map on which contour lines of equal resistivity will reflect the

43
The History of a Technique

A M N „
Ai . iB

Resistive
formation

-Conductive
shale —

44
Surface Electrical Prospecting

isobaths (lines of equal depth) of the top of the shales, as well as the
general shape of the structure. The reasoning is identical if the deep
underground is more resistant than the shallow formations.
The analysis of sedimentary series is usually less simple. If the com¬
position and hence the average resistivity of the upper formations vary
laterally, this variation will affect the resistivity map and make its interpre¬
tation more complex. In fact, to distinguish between the respective influ¬
ences of the shallow and deeper formations, it is necessary to use two
different spacings for AB, corresponding to two different depths of inves¬
tigation. For example, by comparing the readings it is possible to determine
whether a portion of the resistivity map indicates a deep tectonic feature
such as an anticline, or merely a change in the lithological structure or
thickness of the shallow formations. As a general rule, the resistivity map
will be plotted only with a long AB spacing; the short spacing serves mainly
as an aid in its interpretation.
"Dragging" with two spacings, whether for low-dipping or for up¬
turned formations, has remained to this day one of the basic techniques of
electrical prospecting. Although it seems simple enough, in the 1920’s it
was such a valuable innovation that Conrad and Marcel, and particularly
Eugène Léonardon, endeavored for several years to keep it secret. In fact,
until protected by a patent, the very terms “resistivity map” and “apparent
resistivity” were carefully omitted from all communications, such as re¬
ports and publications outside the Company. Instead, vaguer terms such
as “electrical profiles” and “parameters” were used.
It frequently happens that geological data fail to indicate whether or
not a dip is steep. Usually greater accuracy in this regard may be obtained
from the resistivity map. A rapid variation of values, that is, a tightening of
the equal-resistivity curves, reveals a feature characteristic of two differ¬
ent formations making contact alongside a steeply dipping plane, be it a
stratigraphic boundary, a fault, or a sheer fold. Widely spaced curves are
indicative of slow variations and are generally evidence of formations that
have a smooth tectonic trend with very low dips. To complete and refine
the data of the resistivity map in such zones, a new method was intro¬
duced: “vertical electrical sounding,” a procedure that has remained basic
to surface electrical prospecting. Although the method is suitable mainly
for the reconnaissance of strata composed of horizontal layers, it is also
applicable wherever the dips do not exceed a few degrees. It consists of a
series of measurements carried out with increasing AB spacings, while the

45
The History of a Technique

center of AB remains in a fixed position. Thus each station records


apparent resistivities pertaining to slices of ground of increasing depth.
Eventually, the rather short spacings used initially reached several
hundred meters.9 The results of the measurements were then plotted on a
graph with apparent resistivities as ordinates and the successive values of
AB as abscissae. The coordinates were linear. The curve reflected the
depths of the formation boundaries, provided that the formations re¬
mained reasonably homogeneous horizontally, at least within the range of
the investigation.
In most cases vertical sounding curves are representative of an area
where the subsurface is divided into several thick layers, each distinguished
by its average resistivity. In fact, each of these layers is, geologically
speaking, a stack made up of many thinner layers. Whatever the contention
of certain researchers of that time, electrical soundings are insufficient for
identifying every single layer of a stratigraphic series.

The potentiometer and its accessories

As has been described, before 1914 Conrad employed a rather


rudimentary measuring apparatus for his testing and research. Now the
time had come when it was necessary to tool up for industrial operations
for which dependability, efficiency, and cost were as important as the
accuracy of the measurements. What was needed was a device simple to
operate, of limited weight yet sturdy and highly sensitive; it also should
have a rather short time constant and adequate electrical insulation.
Whatever the terrain or the climate, the measuring tool had to be easily
transportable by man or vehicle, without risk of damage that would inhibit
a rapid resumption of the readings.
Once the exact specifications had been drawn, the Hartmann and
Braun firm was selected to manufacture the device. It took only a few
months to produce the famous “black box,” which for the next 20 years
was to remain the basic tool of prospecting and is still used for certain
types of surface work. Compact, lightweight, and watertight, the box
screwed onto a folding tripod and could be carried on the shoulder like
any topographic surveying instrument. The galvanometer it contained was

9 After 1927 the spacings were extended to several kilometers.

46
Surface Electrical Prospecting

as sensitive as a laboratory instrument,10 but once locked (at the touch of


a button), it could be moved about safely.
The measurements determined a potential difference between two
ground points through electrodes, the resistance of which usually changed
from one station to the next. The reading of such differences on the dial
of a galvanometer, serving as a millivoltmeter, would have required the
insertion of a high additional resistance into the circuit to render negligi¬
ble the resistance of the ground electrodes. This, however, would also
have reduced the sensitivity of the circuit to a point incompatible with a
precise reading. For this reason, the instrument included a potentiome¬
ter,11 which made it possible to use the null method with the assurance
that the measurement was independent of the total resistance of the
circuit. In this method the potentiometer circuit consisted of an ordinary
1.5 volt dry battery connected through appropriate resistors in opposition
with the potential difference being measured. The resistors were adjusted
through knobs graduated in millivolts (one knob for tens and one for units)
until the two potential differences reached equilibrium and the reading in
millivolts thus indicated the value of the measurement. The potentiometer
had four sensitivity ranges.
Nonpolarizing electrodes of the kind built by Conrad for his prewar
tests were required for spontaneous polarization measurements. Since the
potential difference between these electrodes was very low, the needle
remained near zero when the station was far from the ore, provided that
there was no noise from stray or telluric currents (see p. 85). When
approaching a deposit, the spontaneous currents deviated the needle; by
bringing it back to zero, the millivolts could be read as indicated above.
Nonpolarizing electrodes were first used for resistivity mea¬
surements; later, they were replaced by copper pegs whose polarization
was compensated for by a small auxiliary instrument. Under such condi¬
tions, with no current flowing between A and B, and on the assumption
that the noise level was weak enough not to interfere, the needle of the
galvanometer would remain stable and near zero. When the emission of a
current generated a potential difference between M and N, it deviated the
needle, sometimes out of scale. The measurements were made by succes-

10 From 25 to 150 x 10-8 ampere per scale division.


11 As a matter of fact, the name “potentiometer" has always been extended to include the
whole instrument (“the pot” in “La Pros" language).

47
The History of a Technique

sive approximations. The circuits of current transmission and of the


potentiometer’s dry battery were closed and opened simultaneously and
repeatedly, the duration of each closing varying from a fraction of a
second to several seconds, until two successive current emissions, corre¬
sponding to two adjoining graduations of the millivolt knob, slightly
deviated the needle in opposite directions; the measurement was then
bracketed between the two graduations, the needle of the galvanometer
serving merely as a mark.
To obtain accurate resistivity values with an unreliable standard like a
portable flashlight battery, the technique consisted of using the same
potentiometer, and hence the same battery, to measure the potential
difference between M and N as well as the intensity of the current
transmitted, the apparent resistivity being proportional to the ratio be¬
tween these two quantities. A precisely calibrated shunt of 0.1 ohm was
placed in series in the AB circuit. To measure the current, the shunt was
connected with the input terminals of the potentiometer; from that point
on, the operation was the same as in the preceding case, the intensity of
the current being read on the knobs.
With the AB profile method the prospector moving the potentiome¬
ter from one AIN location to the next would also move away from the
generator (initially a motor-generator unit, but soon thereafter dry bat¬
teries) and from the switch. The latter was operated by an assistant who
closed and opened the circuit in response to horn signals from the
operator, who, in turn, synchronously closed and opened the potentiome¬
ter circuit. The switch was contained in a box, together with the shunt and
an ammeter, and placed on a tripod. At the beginning and the end of the
profile, the prospector measured the value of the current with his poten¬
tiometer by connecting the input plug to the terminals of the shunt;
meanwhile, the assistant checked to see that the current had not varied on
the ammeter, as occasionally happened with worn-out batteries. Later, the
operation became more practical when the box containing the shunt and
the ammeter was fastened on the right-hand side of the potentiometer;
the prospector could then control the current transmission from the very
point where he was making his measurements, and the horn signals could
be eliminated. The switch was operated by a button located next to the
button that opened or closed the potentiometer circuit. Through a clip
resting on both buttons and a small system of springs located under the
instrument (the arrangement looked like the hand control of a bicycle

48
Surface Electrical Prospecting

brake—hence the name "Bowden”), the operator was able to open and
close the two circuits simultaneously by a pressure of the hand.
Whereas nonpolarizing electrodes were essential for the measure¬
ment of spontaneous polarization, several years passed before it was
realized that they were not required for resistivity measurements. The
discovery was made in April 1925 in the Haguenau forest, where resistiv¬
ity profiles (“draggings”) were made with short spacings (250 meters for
AB, 50 meters for AIN). Conrad was visiting the team, and while taking
part in the measurements, he noticed that the needle deviated every time
Roger Jost, the young prospector, moved his foot in the vicinity of an
electrode.12 It might be, Conrad thought, that if a dry battery were
connected to the terminals of the potentiometer, the needle would return
within scale. A battery purchased from a local store was found to be
suitable. It was round and yellow and looked somewhat like a hunting
cartridge; hence the assembly (i.e., a few solders and resistors) was chris¬
tened “the millivolt cartridge.” This also became the name of the new
black box, now built with a circular rheostat on the left-hand side of the
potentiometer. The nonpolarizing electrodes were replaced by copper
pegs, and the electromotive force generated by their polarization was
compensated for by the cartridge. In addition to the advantages of simple
assembly and operation, this device substantially decreased the resistance
of the ground electrodes, thus increasing the sensitivity of the system.
Finally, the potentiometer (Fig. 16) was completed by the addition,
on a third side, of yet another black box containing an induction correc¬
tor. This was a small, adjustable transformer, with its primary winding in
series in the current circuit and its secondary in the measuring circuit. Its
purpose was to correct the mutual induction between AB and AIN cir¬
cuits, which, when long spacings were used, caused the needle to jump
every time the current was switched on or off. The fourth side belonged
to the operator; it had to remain open since it carried the input terminals:
the connectors plugged into the cartridge on the left and into the termi¬
nals of the shunt or into the double jack leading to AI and N on the right.
There is little to be said concerning the other accessories. Since no

12 Probably a phenomenon of électrofiltration: when more or less mineralized water flows


through a porous body, the motion generates an electromotive force proportional to the
pressure producing the flow. In the case above, when Jost moved his foot, the changing
pressure on the loose soil (humus) resulted in changing the electromotive force and hence
the stray current flowing through the electrode, the line, and the potentiometer.

49
The History of a Technique

Figure 16. Potentiometer with switch and shunt on the right, polarization compensator (mil¬
livolt cartridge) on the left, induction corrector facing. The Bowden cable can be seen between the
legs of the tripod.

polarization problem arose for the ground electrodes A and B, they were
made of wooden, iron-sheathed pegs, while copper was used for M and N.
They were driven into the ground with steel-hooped mallets. The cable
was made of stranded copper wire, later reinforced with steel wire, and
came in 500 meter spools. It was dispensed by hand-operated reels
mounted on tripods. Convenient as the equipment was, its operation
sometimes caused difficult problems. Roads were not always available for
the pickup trucks. More than once the cables had to be unwound by hand
across woods, dunes, or marshes, and batteries and pegs had to be carried
by the team in all kinds of weather. In humid or rainy weather, insulation
defects were irritating because leaks in the AB circuit had an ill effect on
the measurements.

50
Surface Electrical Prospecting

For years the effect of leaks could be only roughly estimated, and
since nobody knew exactly how they occurred, it was impossible to
obviate them. One of Doll’s first contributions was to analyze these leaks
and to show how to identify and to counteract them. He particularly
stressed the importance of connecting the metal box of the potentiometer
to the potential of the shunt when the current was being measured; this
prevented leaks and electrostatic effects on the winding of the galvanome¬
ter. He also shed light on the effects of the resistance of the ground
electrodes on the errors arising from the leaks. A small special device,
called relay resistance, was connected in series next to each ground elec¬
trode A and B; by reversing the current, it was possible to multiply the
resistance of the ground electrodes by an average of three to five times. The
necessafy control was provided by transmitting the current in both direc¬
tions and comparing the measurements.
Taking the readings themselves often proved difficult. With very low
resistivities and very long electrode spacings, the potential differences
could become as low as a fraction of a millivolt and require the highest
sensitivity range of the potentiometer. The total resistance of the measur¬
ing circuit could be so high that at the closing of the current the deviation
of the needle was hardly visible. A tiny microscope fixed above the dial
helped in such cases. Additionally, should the needle become erratic
because of stray currents, it was very time consuming and difficult to
achieve an accurate reading (Fig. 17).
In the case of long spacings and low resistivity formations, another
complication arose from the phenomenon known as skin effect. Since at
the closing of the circuit the current concentrates near the surface before
spreading in depth, the potential difference between M and N was tem¬
porarily higher than after stabilization. As a result the needle deviated
beyond any possibility of correction, and a delay of several seconds was
necessary before equilibrium could be reestablished and the reading
taken.
When a survey takes place close to industrial installations such as
electric railroads or power stations, stray currents can reach such an
intensity that they cause erratic needle fluctuations which preclude any
measurement with direct current. On the other hand, alternating current
of the sinusoidal type is also unsuitable because of the mutual induction
between AB and MN and, in the case of long spacings, because the skin
effect reduces the penetration of the current into the ground and, hence,

51
The History of a Technique

Figure 17. A painstaking measuring with the potentiometer (1928).

the depth of investigation.13 As a remedy Conrad devised a variant of the


direct current technique, consisting of rapid periodic reversals,14 and the
first “pulsator” was born.
Basically this instrument was made of two sections operating syn¬
chronously, one in the current emission circuit, the other in the measur-

13 The reduction becomes larger as the frequency of the current increases, the length of
the transmission line increases, or the resistivity of the ground decreases.
14 The idea was proposed at the same time (1926) by O. H. Gish and W. J. Rooney.

52
Surface Electrical Prospecting

ing circuit. Switched off and reversed 15 to 30 times per second, the
direct current was sent into the ground electrodes A and B, thus creating
interrupted and reversed differences of potential, superimposed on those
generated by the stray currents circulating between M and N (Fig. 18).
The pulsated reversed component being rectified, the potentiometer
received 15 to 30 continuous impulses per second, separated by pauses
too short to be perceptible; and since the parasitic stray component
reached the potentiometer with the same pulsated-reversed frequency,
the result was only a very slight wobbling of the needle about a fixed
position: the inconvenience of an erratic needle was eliminated.
In practice the pulsator was made of two cylindrical collectors
mounted side by side on the same shaft, and connected by brushes to
the emitting and measuring circuits. The first model designed by Conrad
was faultless in terms of physics, but not with regard to its mechanical
functioning. Two bronze bearings, in particular, required a most delicate
and rather precarious adjustment. This made the model impractical for
handling in the field, and as a result it had a short life span.
Marcel then tackled the project, on the basis of one of those simple
and practical ideas that were his forte. What was needed was a device with
a double collector that rotated much faster than its driving crank. He
thought of a tool available in any hardware store, namely, a hand grinder.
All he had to do was substitute the collector for the grinding wheel, with
the necessary brushes and wiring. The result was an instrument that
served for many years in surface prospecting; over several decades it was
gradually improved and became part of logging equipment until it disap¬
peared around 1955 to I960 with the advent of electronic instruments.
A rather long idle time between phases had been provided in the
measuring part of the pulsator to eliminate the effect of the transitory
phenomena occurring at the closing and opening of the current pulses.
This was satisfactory for surveys with relatively short spacings on rather
resistant ground, but it became inadequate with long spacings and on
conductive ground where the transient from the skin effect can last for 1
second or more. At a frequency of 15 to 30 cycles per second, the current
no longer had time to penetrate to the depth it would have reached under
a steady state. In other words, the depth of investigation was substantially
reduced to about the same extent as if a sinusoidal current of the same
frequency had been used. For this reason, surveys with long spacings were
restricted for many years to the direct current technique with all its

53
The History of a Technique

Dead
Intensity i segment
of emission
current

Time
-—►

Voltage
Dead
between
segment
M and N

Voltage
received
by potentiometer
i \

— —

Time

Figure 18. Diagram of operation of pulsated-rectified current. Center: Solid lines show voltage
created between M and N by the current emitted. Dotted lines show voltage from stray currents,
with variations very long as compared with emission frequency. Bottom: Solid lines show rectified
voltage sent to potentiometer. Dotted lines are pulsated-inverted stray voltage, producing a mere
vibration of the needle.

54
Surface Electrical Prospecting

shortcomings. For the prospector this meant spending a long time at each
station, bent over his microscope and pressing the Bowden as many as 20
or 30 times, with 2 or 3 seconds for each closing, while endeavoring to
nonce some acceleration or slowing down of an erratic needle—all of this
under broiling sun or in freezing cold. In heavy rainstorms standard
equipment for each team was a large umbrella. The current transmitted
was pushed up to 2 amperes and more. The batteries, designed for high
voltages and low weight, polarized rapidly under this treatment: voltage
was lower at each closure, the current had to be measured several times,
averages had to be made, and so forth. Not until 1940, after adapting the
photographic recorders for telluric surveys, would Marcel be able to
improve the operation by recording the measurements on film. This
subject will be discussed later in connection with tellurics (p. 90).

Toward increasing depths of investigation

By 1927 the resistivity measurement techniques had already been


seasoned by years of field experimentation. Although these techniques
remained basically unchanged, some new developments resulted from a
broadening of their application to the study of deeper geological struc¬
tures.
Until then, the location of salt ridges or domes had been determined
by the red (conductive) areas of the resistivity map, which were supposed
to indicate the zones where the upthrust of the salt had thinned out the
superficial formations (alluvium) overlying the older and more conductive
marls. For this purpose the use of a short spacing of the current electrodes
(400 to 500 meters) was adequate, except that the image of the objective,
namely, the top of the salt, was rather distorted. Furthermore, the red
areas could be indicative of something other than the upthrust of the salt:
as the overlying alluvium becomes more shaly, its conductivity increases.
On the other hand, the alluvium even on top of the salt is sometimes so
thick that such short spacings do not reveal anything. For these reasons it
became important to increase the depth of investigation in order to obtain
resistivity maps indicating the presence of the salt itself and to “see” it, at
least at its structural highs. Since salt is electrically insulating, such highs
would appear as blue (resistant) areas on the map. The tops of most salt
domes lie at hundreds rather than tens of meters under the surface,

55
The History of a Technique

necessitating a corresponding increase in the spacing of the current elec¬


trodes.
The same problems existed in regard to the reconnaissance of anti¬
clinal structures without a salt core. In a stratigraphic series assumed to
contain a petroleum deposit, at least one formation or system having a
structure in conformity with that of the assumed reservoir had to be
identified. This formation or system would then serve as a marker; it had
to be rather thick and to have an average resistivity that differed signif¬
icantly from that of the overlying formations. Yet the top of such a
marker is often hundreds of meters deep, requiring systems that permit
deep “seeing."
In theory there seemed to be no limit to the depth of investigation;
the current electrode spacing could be increased indefinitely. The
difficulty lay in the handling of very long cables and in the limited
sensitivity of the measuring instruments. The greater the spacing between
A and B, the smaller the potential difference between M and N. Efforts
were made to use voltages as high as possible in the AB circuit, but safety
precautions, as well as the bulk and weight of the batteries, restricted the
AB circuit to 500 volts. Bulk and weight also imposed a limit on the
possible reduction of the ohmic resistance of the circuit; further reduction
would have required many long stakes and large-diameter cables.
Under these conditions the spacing could be increased to 2000, then
4000, and even 6000 meters in 1929, a record for that time. There was,
however, one exception, a rather bold and entirely original experiment
undertaken by Conrad and Marcel in the spring of 1928 near Vitré in
Brittany. The experiment dealt with an electrical sounding with a current
electrode spacing ranging from 2 to 200 kilometers, and a spacing be¬
tween potential electrodes M and N of up to 20 kilometers. The sole
purpose of the brothers’ efforts was scientific research in exploring the
thickness of the earth’s crust on the assumption that beyond 10 or 20
kilometers the rocks become semifluid and much more conductive than
the crust. Telegraph lines were made available by the Post and Telegraph
Administration. The current was reversed for periods of up to 4 seconds
for the long spacings to counteract the difficulties arising from telluric
currents, mutual induction, and skin effect, and thereby obtain a depend¬
able average reading. A high-capacity dry-cell battery could deliver a
steady 2 amperes for several minutes. Since the subsurface in Brittany has
a very high resistivity, the potential differences to be measured, even with

56
Surface Electrical Prospecting

a 200 kilometer spacing, were compatible with the sensitivity of the


potentiometer.
The result was an electrical sounding curve. The remarkable regu¬
larity of the plotting attested to the dependability of the method. Resistiv¬
ity increased sharply with depth without decreasing at any level; even with
this 200 kilometer line, the conductive substratum was not felt. Another
discovery the experimenters made was that it was possible to explore a
huge slice of the earth’s crust with only a few hundred watts of power—
no more power than is needed to light a 200 square foot room.

Prospecting for metalliferous ores

No substantial change was made in the spontaneous polarization


technique used by Conrad in his prewar tests, even though the findings of
further experimentation proved valuable in fieldwork and in the interpre¬
tation of measurements as well. It was confirmed that formations made of
conductive grains embedded in a resistant gangue do not produce spon¬
taneous polarization. Experience showed that anthracite beds, as well as
pyrites, can be the sources of spontaneous potentials, and that one had to
beware of premature enthusiasm when observing the beautiful reactions
of certain formations that happened to be devoid of any mining pos¬
sibilities, for example, graphite-impregnated schists. Mineral reconnais¬
sance was generally conducted by combining the plottings of spontaneous
polarization and equipotential measurements, the latter being soon re¬
placed by the resistivity map. The technique proved most effective in the
search for all conductive ores: pyrites, copper sulfides, galena, pyrrothine,
and so on. No conclusive result was achieved by various induced polariza¬
tion tests.
In the early 1920’s, the fledgling years, there was keen competition in
prospecting for metalliferous ores, and Conrad and Marcel thought of
extending their operations to include exploration for ores—mainly iron
ores—which, in addition to being electrically conductive, had magnetic
susceptibility. Good magnetometers, particularly the Schmidt balance,
were on the market, but, as Conrad often said, “You don’t develop an
industry from what already exists.” Therefore he endeavored to make a
personal contribution to the advancement of the magnetic method.
Although of high quality, magnetometers were difficult to adjust and

57
The History of a Technique

to handle. They were subject to drifting, and measurements were rather


time consuming. Conrad thought of tackling the problem from the elec¬
tromagnetic end and decided to build a prototype. Even though little
came from the project, it is worthy of a short description. The instrument,
conceived by Doll and engineered at “La Pros’’ by Chatelier and his team,
was of the differential type and able to detect the gradients of the
terrestial field. It was made of two circular frames 1 meter in diameter,
each wired with the same number of turns and driven by a crank around a
horizontal shaft. All the materials were nonmagnetic: wood, plywood,
Bakelite, aluminum, and so on. The first “field” test took place in Paris, in
front of the rue Fabert headquarters. The adjustment of the instrument
was a delicate one that required an absolutely invariable magnetic field. A
crowd soon gathered to examine the bulky rotating machine and to watch
the white smocks scurrying around it. One cyclist in the crowd became
quite fidgety, and each twitch of his bicycle displaced the needle of the
galvanometer, which was already jumping wildly at the passing of every
subway train. It was hopeless. The experiment ended with the arrival of a
policeman, who dispersed the crowd. The tests were resumed at Marcel’s
estate at Cormeilles-en-Parisis (which for some time served as a testing
ground for experiments in the open), but after further field studies the
matter was dropped. As occurred often in the history of “La Pros,” other,
more urgent developments demanded attention.

First studies based on the anisotropy of formations

From the very beginning, Conrad had observed the effect of electric
anisotropy of sedimentary formations. An electric current sent through
the formations flows more easily in a direction parallel than in a direction
perpendicular to the strata; the strata are thus said to be anisotropic with
respect to electric conductivity.
What is the origin of this phenomenon? A sedimentary series often
consists of a sequence of thin layers of limestones, shales, sandstones, and
marls of uneven resistivity. The current is impeded in crossing the succes¬
sive resistant layers frontally, and therefore finds it easier to flow
alongside the more conductive layers parallel to their stratification.
Moreover, the mineral grains which generally constitute the sediments
are nearly flat or oblong and, in most cases, settle parallel to the strata.

58
Surface Electrical Prospecting

This is clearly seen in clays, where the microscopic crystalline particles


look like flattened leaves. This disposition of the grains adds to the
tendency of the current to flow parallel rather than perpendicularly to the
strata.
In an area where the underground is a steeply dipping sedimentary
series like the May sandstones in Normandy, the equipotential curves
plotted in the vicinity of one ground electrode (the other being at a
distance sufficient to make its effect negligible) have an elongated, nearly
elliptical shape. This phenomenon is explained by the mathematical
theory which, when applied to geophysics, maintains that the equipoten¬
tial surfaces are flattened ellipsoids with an axis which passes through the
ground electrode and is perpendicular to the strata. Their intersection
with the surface of the ground is an ellipse whose major axis is parallel to
the stratification. Furthermore, experience shows that this effect can still
be felt on the surface if the uplifted formations are hidden under an
overburden which is rather uniform: not too thick or too conductive in
relation to the substratum.
A systematic application of the phenomenon was made for the first
time at May St. André, in conjunction with the survey in which the first
resistivity profiles were plotted. However, without conclusive informa¬
tion regarding the direction of the dip and its value, the method was of
little use. In his report of 1920 Conrad had noted that to obtain these data
the source of the current would have to be placed at a certain depth within
the sedimentary series; this would require drilling, thus considerably
limiting the applicability of the method. One of the first operations
involving a borehole took place at Velsen in the Saar coal basin during the
winter of 1921. A few more were carried out in the spring of the same
year at Molières-sur-Cèze in the Gard coal basin. There were not many
others before the 1930’s.
To detect the direction of the dip through surface observations, an
electromagnetic method was devised. The basic idea is as follows: on the
assumption that the surface is flat and horizontal, and the underground
homogeneous and isotropic, a current is emitted between two ground
electrodes, A and B, connected by a rectilinear wire. The electrical circuit
is then the equivalent of a vertical frame, the surface wire being one of its
sides. The result is that the average direction of the magnetic field created
by the current is horizontal. If, however, the subsurface is made of
dipping, anisotropic strata, and the wire AB runs parallel to the stratifica-

59
The History of a Technique

don, the frame tends to tilt in the direction of the dip, and the magnetic
field straightens up accordingly. With alternating current, the direction of
the dip can be detected through a receiver reacting to induction, pro¬
vided that the dip is steep enough without being vertical.
A first version of the receiver was a coil about 1 meter in diameter,
fitted with several turns of insulated cable and connected to an earphone.
This device, however, was not employed beyond the experimental stage
(Fig. 19). Another version, extensively used for several years, made it
possible to employ the conventional equipment for resistivity mea¬
surements without any special addition. The receiver was a spire of cable
of a simple, radially symmetrical shape: a square or an octagon of 50 to
200 meters with one diagonal placed alongside the emitting line. The
ground electrodes were symmetrical about the center of the spire (Fig.
20). When a pulsated current was sent into AB, the electromotive force
induced in the spire was rectified by the pulsator and observed on the
potentiometer. When the device was above dipping, anisotropic layers
and the AB line was running straight along the stratification, the magnetic
field had a vertical component that induced a signal in the spire. Bending

Figure 19. Experiments with an electromagnetic system for the determination of formation dip
(Jan Joachim Valley, California, 1928).

60
Surface Electrical Prospecting

Figure 20. Schematic view of the spire.

the cable AB symmetrically about the center in such a way as to form an


isosceles triangle modified the reading. It could be nulled by finding the
proper size and orientation for the triangle. The side of AB where the tip
of the triangle happened to be indicated the direction of the dip, and the
area of the triangle was approximately proportional to the dip angle.
These findings were shown on the topographic map by means of a vector.
In the frequent case where the direction of the stratification was not well
known, two sets of measurements were made at each station with two
perpendicular AB directions, and the sum of the two vectors was plotted.
In practice a reciprocal system was used in which the functions of AB and
the spire (as transmitter and receiver) were exchanged. This gave the same
results, but the sensitivity of the measurement was increased.
This simple and convenient method, requiring only lightweight
equipment, aroused great hopes. In the San Joaquin Valley in California,
where it was applied for many months, the dry climate and the ideally flat
and bare terrain, over which vehicles moved unimpeded, together with a
good operational organization, led to high performances: up to about ten
1000 meter stations per day. Unfortunately further experience showed
limitations in the applicability of the method. It worked only on very flat
ground, and the resistivity of the anisotropic layers had to be rather
uniform throughout the tested field; the presence of thick enough and
more resistant layers could upset the distribution of the electromagnetic
field to the point of giving misleading measurements called “antidips.”

61
The History of a Technique

These would result in lines on the map where the inversion of the
direction of the vectors looked exactly like the traces of anticlinal axes,
whereas they were merely resistant layers outcropping under the caprock.
For these reasons the spire technique was dropped after a few years.

Laboratory work and a bit of mathematics

While methods and techniques were undergoing field tests, interpre¬


tation was being clarified and refined through laboratory observations and
theoretical studies. In 1919 Conrad had resumed his model studies at the
Ecole des Mines in Paris. Perhaps it was an exaggeration to designate as a
“laboratory” the collection of a few instruments and apparatus in a room
of the rue Fabert building. At any rate, after a few years an expanded
laboratory was transferred to a former barroom located at the corner of
rue Saint-Dominique. The galvanized sink of the bar was kept for ex¬
perimentation purposes, as well as the children’s bathtub that had served
for Conrad’s research before the war. Later, larger and more complex
tanks were used; but, as stated earlier, in the idiom of “La Pros” they
continued to be called bathtubs.
Though of modest size, the bathtubs at that time proved to be
valuable in experimental studies, and many models were built. By packing
clay on one side and sand on the other, two formations with different
resistivities and a variable dip were represented. With scaled-down
AMNB configurations, measurements could be made alongside profiles
perpendicular to the trace of the contact between the two formations.
Thus, for each electrode spacing, an apparent resistivity profile could be
plotted and the effect of all factors involved in the measurements assessed.
By filling the bathtub with water, a homogeneous formation could be
simulated and its resistivity adjusted by adding the necessary amount of
salt. Plates of insulating material (usually ebonite) were placed across the
tank to represent resistant layers of variable thickness and dip. This ar¬
rangement resembled sandstone layers embedded in shales. By dipping the
upper edge of the plate below the surface of the water, an overlay conceal¬
ing ancient formations was simulated.
The fact that the possibilities for experimentation were multiple and
highly flexible led to partial clarification of a phenomenon that had been

62
Surface Electrical Prospecting

observed in the field. When the AMNB configuration is moved across the
stratification and ground electrode A nears the trace of a high-resistance
layer, the current emitted is impeded in its forward flow; the current
density increases toward the rear, and so does the apparent resistivity
recorded between M and N. On the contrary, once A has crossed the
resistant layer, the current is impeded in its backward flow and the
measured value decreases. The profiles show a peak followed by a trough,
at a distance of the layer equal to AB/2; a similar effect occurs when
ground electrode B crosses the resistant layer. The crossing of a conduc¬
tive layer by A and B produces similar but inverted pictures. The result is
two “electrode kicks" superimposed over the other peaks and troughs of
the profiles, which appear when MN is on top of a geological feature. If
the feature involves an alternation of several resistant and conductive
layers, it becomes difficult to distinguish between the effects produced by
the electrodes and those caused by geological features.
To increase the reliability of the interpretation, a variant of the
‘‘dragging" technique was worked out, involving two pairs of MN ground
electrodes, adjoining the center of AB. At each station the measurements
had to be repeated with two positions of AB, designated front and rear,
and offset by a distance equal to AIN. If either ground electrode A or B
nears or crosses a bed of contrasting resistivity during the offset, the two
measurements are different; otherwise they are the same. This technique,
called “repetitive dragging,” facilitated the unscrambling of the effect
produced by the “electrode kicks.” However, it also made the mea¬
surements more time consuming and more expensive.

Conrad had a definite gift for physics. His could have been the motto
that Pavlov had engraved on the door of his laboratory: “Observe; always
observe.” Marcel, on the other hand, would say, “Take material and make
it into something that works.” Neither man was attracted by purely
mathematical constructions because both saw them as dangerous over¬
simplifications of geophysical problems. Tongue in cheek, Conrad wrote:

The calculations dealing with these problems are all taken directly from the
theories of electrostatics .... There is plenty of room here for budding
mathematicians to enjoy themselves. Either the questions address them¬
selves to the high school level, or they are so unnecessarily complex that

63
The History of a Technique

they get lost in the clouds of still more complicated formulae; then the wisest
thing to do is to turn the page.15

Doll, the best mathematician of the three, always maintained in


his research a sensible perspective and a well-controlled appreciation of
the value of mathematics.
In the early stages the results of the measurements made in the
bathtubs were supplemented by calculations that, far from being “un¬
necessarily complex,” were modeled on extreme simplicity: a vertical
layer crossed by an electrode setup placed perpendicular to it.16 Later,
however, once his research required a more complete theoretical back¬
ground, Conrad enlisted the assistance of Alfred Liénard, the deputy
director of the Ecole des Mines and a prominent mathematician. On
September 15, 1925, with Liénard’s help, he filed a patent claiming that,
when the subsurface is made up of two homogeneous beds separated by a
horizontal plane, the apparent resistivity can be calculated.17 However, at
that time he did not attempt applications involving the use of numerical
data, and until about 1927 the interpretation of electrical sounding curves
remained empirical, sometimes rather arbitrarily so. This was adequate
for shallow investigations, but when operating with spacings of 4000 to
6000 meters, Conrad and Doll realized that interpretation had to rest on
more dependable evidence, which only mathematics could provide. They
knew that in a medium composed of homogeneous horizontal layers the
potential created by the flow of a direct current emitted by a point source
could be calculated by means of Maxwell’s image theory.18 A young and

15 In Le Puits Qui Parle, No. 4, April 1921.


16 The basic calculation is derived from Maxwell’s image theory. At any point near an
electrode, within a space divided by a plane in two media of different resistivities, the
potential is calculated by introducing a fictitious electrode symmetrical to the real one
about the plane. By analogy with optics, the fictitious electrode is called an image. In the
case of a layer perpendicular to AB (two parallel planes of separation), this theory involves
four infinities of images, yet the calculation is not too long because after a certain number
of reflections the images are so remote that their effects become negligible.
17 He was unaware of the study of O. H. Gish and W. J. Rooney on this problem, which
was being published at the same time in the United States.
18 The theory had already been used by N. Hummel in Germany in dealing with this kind
of problem, where either one or two layers rested on top of a third one of infinite
thickness. Without going into higher mathematics, the image theory led to lengthy and
cumbersome calculations at a time when computers were not yet available. In the case of

64
Surface Electrical Prospecting

brilliant mathematician, S. Stefanesco, worked out a more refined, flexi¬


ble, and expeditious analytical solution, which for decades served as the
basis for a number of theoretical curves calculated by a host of researchers
and operators.
In early 1929 Stefanesco’s method was limited to the calculation of a
set of curves for two formations. This set was called the “chart of the
overburden.” At the same time the diagrammatic presentation was
modified: the values of AB/2 were the abscissae, the apparent resistivities
were the ordinates, and the logarithmic scale was adopted for the theoret¬
ical graphic presentation as well as the transfer of the field measurements.
At that time the interpretation of a sounding curve consisted of superim¬
posing over it a transparent overburden chart and finding the theoretical
curve giving the best coincidence. However, the method was dependable
only when the subsurface appeared as a series of two or, at most, three
clearly distinct layers. In this case hypotheses were possible which, com¬
bined with data from other sources (geology, drilling), led to acceptable
solutions. The treatment of a few typical configurations had shown that
the interpretation of electric soundings was often obscured when a strati¬
graphic series could not be reduced to a very small number of layers.
During the years that followed, some researchers went beyond the
overburden charts. In 1933 the Compagnie Générale de Géophysique
(see p. 82), under the direction of Raymond Maillet, began the system¬
atic calculation of curves for three formations. These were collected in a
book from which it was possible to select the curves most suitable for
interpreting specific measurements. The book contained more than 400
curves. Later, H. M. Mooney and W. W. Wetzel published a book with
2400 curves, of which 2000 are for four formations.19
There is at least one other method for the interpretation of electrical
sounding curves. Upon first examination it is possible to imagine a model
of horizontal layers, calculate their theoretical curve, compare it to the
field measurements, and retouch the model step by step until a satisfac¬
tory coincidence is reached.20 This approach with its cumbersome calcula¬
tions became practicable only with the advent of the computer.

more than three layers the mathematics became inextricable. Hummel’s calculations, as
well as those of L. V. King in England, had no practical application.
19 University of Minnesota Press, Minneapolis, 1956.
20 Theory has shown that there are infinities of combinations in keeping with a measured

65
The History of a Technique

Until 1926 there was no dissemination of technical information


among engineers in the form of service memoranda or bulletins. The
prospectors—and there still were only a few of them—were trained by
association. Sent to the field almost as soon as they were recruited, they
were directed by a senior colleague from whom they received most of
their knowledge. They also had the benefit of exchanges with Conrad and
Marcel when they stopped over in Paris between missions. Additionally,
they received letters from Conrad (his brother was not much of a letter
writer) commenting, advising, or drawing conclusions from the surveys
they were managing. It might be added that the head office was not
markedly eager to disclose information that could be of interest to enter¬
prising competitors, especially in the United States and Canada. This was
one reason why written documents concerning methods as well as in¬
struments and operations were sparingly distributed.
After 1927, however, technical memoranda became rather frequent.
The sum of theoretical and experimental data serving as a basis for
interpretation was extensively exposed in a letter of November 22, 1927,
and was circulated to the the entire staff. It shed a new light on a number
of things and helped to increase understanding of the meaning of the
measurements. Nevertheless Conrad and Marcel maintained their belief
that theoretical deductions led to oversimplification of hypotheses, and
that field observations were the ultimate and safest elements of decision.
This was the thinking behind the remark Conrad made to a prospector: “If
you find something in the ground that does not show up on the mea¬
surements, it’s better to believe the ground than the measurements.” A
quip indeed, but how realistic at a time when surveys dealt only with
shallow structures, and careful examination of the terrain would help in
the understanding and verification of the electrical measurements.

Operations

In 1920 electrical prospecting, as well as all other geophysical


methods, was virtually unknown to the mining industry. Schlumberger’s
many explanations and demonstrations gradually introduced a new tech-

curve. To narrow the possible solutions of the problem, other data on the formations
studied and their resistivities are required.

66
Surface Electrical Prospecting

nique into the field of traditional mining operations: instead of digging to


explore the underground, a small array of electrical equipment (wires,
stakes, batteries, potentiometers) could be used to scan from the surface.
In addition to the traditional dowsers, quite a number of “inventors”
were peddling their allegedly “scientific” wares, claiming that they had ef¬
fected astounding discoveries. Some industrial concerns, unable to distin¬
guish between such eccentricities and well-established methods, remained
skepdcal. The terms “black magic” and “geomancy” gained popularity.21
The more realistic mine operators were less hesitant in making their
choices. In France several mining concerns, already aware of Conrad’s
scientific authority, were quick to express a preference for his methods.
Other qualified researchers either became involved in, or continued
operations in, geophysical prospecting: Lundberg, Mintrop, Ambronn,
Koenigsberger, Slichter, and Mason were the principal ones. From the
start, Schlumberger had to compete actively, not only in gravimétries and
seismics, but in the field of electrical prospecting as well. For the latter,
certain operators used high-frequency electromagnetism, but it was all too
easy to imitate the spontaneous polarization, the plotting of equipotentials,
and resistivity measurements with short spacings.

From 1920 to 1929

In the early 1920’s most of Schlumberger’s activities consisted of


short-term missions (generally a few weeks) scattered throughout a
number of countries: France, North and South Africa, Spain, Italy,
Yugoslavia, Bulgaria, Poland, the United States, Canada, and Japan. In
France outstanding operations included the reconnaissance of Silurian
synclines in Normandy, the location of faults in the Lorraine coal and the
Briey iron ore basins, the use of spontaneous polarization in prospecting
for anthracite at Saint-Michel de Maurienne, and the exploration for
lignite in the Landes.
Mineral prospecting by electrical measurements had been introduced
into the United States and Canada in the fall of 1921. The first experiment
took place in the copper ore basin of Ducktown, Tennessee, where the

21 In the French text of a 1919 patent, H. Lundberg, one of the best known geophysicists
of that time, refers to the measuring ground electrodes as "divining rods.” Perhaps
commercial reasons led him to adopt this rather odd analogy with the dowser’s tool.

67
The History of a Technique

shapes of the lodes at Mary Mine, Boyd, and Culchote were traced by
spontaneous polarization. Other tests in Pennsylvania, Michigan, and
Ontario led to commercial development, especially in Manitoba and
Saskatchewan. Outstanding results were obtained at the Noranda mines
in Canada, where an important deposit was discovered in the unexplored
part of the lease.
Even though the activities of Conrad were far from widely known, he
was contacted by companies wishing to license his processes. The first
contract of this type was negotiated with a French concern prospecting in
Ecuador. For 1 year and for the payment of 10,000 French francs, the
Schlumberger Company guaranteed to supply all equipment and operat¬
ing instructions. Under similar contracts drawn up during the following
years with mining companies in Rhodesia, Katanga, and Japan (Yokei area
and Nagamunu mine), extensions of already-known copper mines as well
as new deposits were discovered.
In a country like France, which, according to most geologists, had no
prospective petroleum wealth, electrical prospecting was limited to
geological reconnaissance about the small oil field of Pechelbronn, north
of Strasbourg. One of these explorations succeeded in locating the trace
of the great Rhenish fault under the Quaternary terraces near Lobsann
(1921). The news of these results spread to Steaua Romana, a French oil
company operating in Rumania, where some geophysical work was in
progress. A contract was signed in 1923, and except for an interruption
from 1927 to 1931 Schlumberger crews worked there for more than 20
years.
Initially, Conrad’s main thought was that the salinity of formations in
the vicinity of petroleum deposits ought to be very high.22 Consequently,
areas where the readings were lowest were sought out on the resistivity
maps. A few months after reconnaissance of the Prahova district west of
Ploesti, Rumania, was begun, the map showed a red spot that stood out
sharply near the village of Aricesti. Two boreholes had detected the flank
of a structure to the north of the red spot; the electrical method confirmed
its existence, and a contour map of equal-resistivity values showed its
outline (Fig. 21). Other boreholes, sited on the basis of these data,

22 In 1923 he tried to verify this assumption by asking Pechelbronn to make mea¬


surements on clays both near the producing zones and remote from them, but Pechel¬
bronn was unable to comply.

68
Surface Electrical Prospecting

30
Equiresistivity curve and value
of resistivity in ohms m2m.
Closer hatching for lower resistivity.

Scale

Figure 21. Resistivity map of the Aricesti, Rumania, salt dome (1923).

revealed a salt core and a huge reservoir of hydrocarbons. This was the
first time in history that an oil-bearing structure had been determined by
geophysics.
Although the conclusions drawn from the resistivity map were cor¬
rect, the theory on which they rested was later disproved. Experience
showed that saline formations were frequently present in oil-bearing
basins, but because they extended far beyond the immediate vicinity of
the producing zones, they could not be used as guides for oil explora¬
tion.23 It took several years before a new and correct interpretation of the

23 Another long-advocated idea was that salt domes diffused at the top and formed “salt
aureoles” that served as the origin of the conductive zones on the maps. This assumption,
too, was gradually disproved.

69
The History of a Technique

Aricesti map was made. The stratigraphic series contained a thick stratum
of Tertiary marls overlain by a substantially more resistant Quaternary
alluvium (sands and gravel). During the Quaternary the dome kept mov¬
ing upward (and continues to do so to this day), causing the alluvium to
thin out. The equal-resistivity curves reflected the topography of the top
of the marls, while a conductive spot at the center located the crest of the
structure.

In the United States petroleum exploration was beginning to apply


other geophysical methods. In 1921, in Oklahoma, J. C. Karcher and
W. P. Haseman ran the first successful tests of seismic reflection. The
gravimetric method was being introduced in California and Texas, and
seismic refraction on the Texas coast and in Mexico. In 1924 two spec¬
tacular discoveries of productive structures confirmed the value of these
methods: the Nash Dome in Texas, located by the Rycade Oil Company
with the help of the Eotvos balance, and the Orchard Dome, also in Texas,
discovered through seismic refraction for the Gulf Oil Corporation by
Seismos, a German firm.
The Royal Dutch Company’s American affiliates had undertaken an
intensive geophysical prospecting campaign. Informed of the success of
the Schlumberger processes in Rumania, they decided to test the resistiv¬
ity method in Texas. Accordingly, Marcel Schlumberger arrived in Hous¬
ton in June 1925 with his nephew Marc, and proceeded with his first
measurements (resistivity profiles with 400 meter spacing) on the Pierce
Junction Dome, the southern flank of which was producing 500 barrels
per day. Other tests followed in the same area (Blue Ridge, Humble, and
Goose Creek domes) and continued toward Beaumont (Spindletop, Fan-
nett, and Sour Lake). In spite of experimental difficulties with existing
pipeline systems and leaks from cables lying in the saltwater puddles
spilled by certain wells, the profiles showed a resistivity drop above the
producing domes. Continuing with the notion that saline formations
indicated the proximity of hydrocarbons, it was thought that electrical
prospecting detected the pay zones, whereas the balance and seismic
methods served strictly for locating salt domes, whether productive or
not. The electrical method seemed to be more efficient; its equipment was
easier to handle, and its operation much less costly. Thus the first contract
with Royal Dutch was signed on September 23, 1925, on behalf of its
affiliate, Roxana Petroleum, which was operating in Texas.

70
Surface Electrical Prospecting

Marcel had returned to France before the end of the summer, leaving
behind his nephew (later joined by Pierre Baron and Jacques Gallois) and
a few junior engineers—Paul Charrin, Gilbert Deschatre, Julien Breusse,
Marcel Jabiol, André Allégret, and Robert Roche—the first group to
settle in the United States, in Freeport, Texas. In the early part of the
following year, operations were extended into Louisiana (Fig. 22).
Few domes were known at that time, but geology led one to believe
that many others could be discovered. The active competition among the
various companies to acquire the most promising areas24 gave geophysical

Figure 22. Electrical prospecting in Louisiana swamps (1925). (Reproduction from La revue
pétrolière, 1935, p. 1608.)

24 In the United States the subsurface belongs to the owner of the land; therefore such
areas had to be purchased or leased.

71
The History of a Technique

processes a strong boost. Roxana was interested in the work of “La Pros”
but without great expectations of immediate results. At any rate, in May
1926 an office was opened in Dallas. Marc Schlumberger’s visa had
expired, and he was replaced by Charrin until October, when Léonardon
arrived.
Conrad, on a visit in September, was greatly interested in a conduc¬
tive area on the resistivity map of Bayou Serpent, Louisiana. Contrary to
the opinion of Roxana’s geologists, he suspected the possibility of a dome,
and after thorough consideration he took it upon himself to suggest the
location of a borehole. Roxana had other plans, however, and after several
postponements finally shelved the proposal.
Having used electrical prospecting for 18 months with no conclusive
results, Roxana figured that Schlumberger was merely experimenting,
often blindly. Indeed there was some justification for this thinking: opera¬
tions were still run with spacings which were too short, and no theory had
evolved that gave a clear understanding of what the measurements rep¬
resented. Moreover, their accuracy remained questionable. Giving pref¬
erence to balance and seismic methods, Roxana finally canceled its con¬
tract in the spring of 1927.
Meanwhile, however, another contract had been signed on March 21,
1926, with the Shell Oil Company; this brought a Schlumberger crew (R.
Nisse and R. Viry) to the San Joaquin Valley in California, in search of
anticlinal folds as a possible extension of the Lost Hills and North Bell-
ridge structures. No results had been obtained there in more than 2 years
with the balance, and the only discoveries had come from boreholes near
surface showings of asphalt.
The local working conditions were excellent: the region was a flat
semidesert; travel was easy and the climate dry. No electrical leaks were
expected. The spacing could be increased to 1000 meters. There were no
salt domes or conductive spots showing on the maps, and the anticlines
seemed to correspond to maximal resistivities. Conrad, who visited the
mission after his trip to Texas in the fall of 1926, thought that this was
indicative of rather hard sandstone beds marking the crests of the struc¬
tures.25 A second crew was stationed at Wasco, California, from the spring
of 1927 until the early part of 1930, when the mission ended. Meanwhile,

25 The same results could lead to different geological constructions, but they remained at
the hypothetical stage.

72
Surface Electrical Prospecting

the multielectrode resistivity method had been supplemented by the


inductive method referred to as the spire, particularly for studies of dips.
However, in the wake of the disappointment caused by the suspen¬
sion of work in Texas, Paris was inclined to give America second priority.
Slow communications prevented careful monitoring of the fieldwork, and
the market proved to be much tougher than in France. The trend was to
repatriate the surplus engineers since they could be fully occupied in
France. It required perseverance on the part of Léonardon to ensure that a
half-dozen available prospectors were assigned to Canada, where mining
exploration was in full swing. Changing from the Texas heat to the rigors
of the Great North, the prospectors traveled from mine to mine between
August and October; numerous vein and lode surveys by spontaneous
polarization and resistivity were made under conditions of severe hard¬
ship. Backed by the unrelenting commercial efforts of Léonardon, who
was now established in New York along with most of the mining com¬
panies, these activities continued, through thick and thin, until 1930.

Exploration for new deposits in the Alsatian potash basin of France


began in 1926. There were two potash beds, intercalated in a thick
formation of rock salt; therefore it was necessary to locate the areas where
the rock salt came closest to the surface. It lay directly beneath a thick bed
of uniform Oligocène marls, impregnated with highly mineralized water
and overlain by Quaternary sands and gravels of much higher resistivity.
Since this configuration was almost identical with that encountered in
Rumania, a similar operating method was adopted: a resistivity system
with rather short electrode spacing (AB = 350 meters). After a few weeks
of prospecting, a conductive strip about 7 kilometers long, running along
the Ill River, revealed a broad saline ridge: the Meyenheim anticline had
been discovered (Fig. 23). In the following fall another red spot marked
the Hettenschlag Dome, located north of Meyenheim (Fig. 24). Two
boreholes confirmed that the salt was 100 meters deep, as indicated by
vertical electrical soundings. These were momentous discoveries in the
history of geophysics.

By late 1927 the number of engineers employed by “La Pros” had


increased from 3 to 16. During the last 7 years some 50 missions had
taken place, many outside France. Most had yielded positive results. The
Schlumberger processes had become better and better known; some 10

73
i

Ensisheim

2km
The History of a Technique

publications (articles, brochures, and papers presented before various


congresses) by Conrad Schlumberger, Léonardon, Charrin, and others
had appeared in France and the United States. Mining companies in South
Africa and Japan had disclosed the results of the processes in the profes¬
sional journals of their countries. With regard to petroleum exploration, a
field in which Schlumberger had aggressive competition, a number of
decisive successes had proved the superiority of gravimetry and seismic
refraction over electrical prospecting. Conrad thought it preferable,
therefore, to focus the commercial effort on detailed geological studies
(uplifted beds, faults, etc.), where the superiority of the resistivity method
was well known and “La Pros” had a marked technical lead.
With the proposal of a dam on the Connecticut River, a new field of
applicability opened up: civil engineering. For this project electrical
soundings were used to determine the depth at which the dam should be
seated (April-May 1928). Twenty surveys of this kind followed in the
United States, Canada, and North Africa. In particular the preliminary
reconnaissance of a dam site on the St. Lawrence River near Morrisburg,
Ontario, in the spring of 1929 was among the outstanding achievements
of electrical prospecting. This method was also used to locate competent
rocks among sedimentary complexes and to study the depth of crystalline
substrata, problems that arose during the digging of tunnels at Bridge
River, British Columbia (summer of 1928), and in the Lièvre Valley near
Masson, Québec (1929). Also in 1929 Jean-Paul Mathiez carried out a
survey for several months on behalf of the American Zinc and Lead
Smelting Company, Joplin, Missouri, to plot the subterranean topography
of the roof of limestones covered by shales. The technical quality of the
survey was such that it earned recognition at “La Pros” as a textbook
example.
In the eastern hemisphere there was no need to pursue contracts and
fight tough competition. Surface work proceeded smoothly. In addition to
many small missions, long-term surveys included new studies of iron
ore-bearing synclines in Normandy (Soumont, 1928-1929; Halouze,
1929; Sées, 1931); prospecting large areas for salt domes in the Landes in
search of potash, a problem similar to that in Alsace (1929); discovery of a
pyrite body at Chizeuil, Nièvre (1929); and a big success at Benissaf in

Figure 23- Meyenheim anticline (1926). The equal resistivity curves east of the III River draw
the outline of the salt ridge.

75
The History of a Technique

Algeria, where a previously unknown bed of iron ore 50 meters thick was
uncovered by resistivity and magnetometry and confirmed by drilling
(spring of 1929).
Abroad, the team of Paul Charrin and Jacques Castel was active in
Spain between the fall of 1927 and the summer of 1928. The team had
won out over the competition of seismic refraction and gravimetry in the
coal basin of Villanueva-de-las-Minas in Andalusia, and then continued
quite successfully with the search for metallic ores (copper, lead) in the
Cordoba and Huelva regions, as well as in Spanish Morocco. The follow¬
ing year more missions operated in Africa: Ookiep (Union of South
Africa), where a mission discovered an anomaly that, years later, led to the
development of a copper mine; Upper Katanga, where the Union Minière
depended on Schlumberger’s technical assistance for the interpreta-

76
Surface Electrical Prospecting

don of surveys under license; and Minouli (Brazzaville area), where,


presumably because of the lack of ore, the survey turned out to be
disappointing.
The spring of 1929 marked the beginning of negotiations for an
important contract with the U.S.S.R. At the same time electrical coring
was introduced into Venezuela and was later tested in the United States.
The central facilities in Paris were reorganized and expanded. New build¬
ings were acquired at Nos. 40 and 42 rue Saint-Dominique; a new
workshop was built; the operating, equipment, and administrative depart¬
ments were restructured; and new staff was recruited. Such optimism was
fully justified during 1929: by the end of that year, the staff had grown
from 56 to 95, of whom 45 were field engineers.
The'Wall Street crash of October 1929 hit the mining and petroleum
industries hard. As early as 1930 exploration in the United States and
Canada had slowed down or stopped altogether, engulfing geophysics in
an acute crisis that was soon to extend to the whole world except the
U.S.S.R. After years of what seemed to be unending growth, the bottom
was falling out of everything. It would be several years before the oil and
mining companies recovered from the shock and exploration resumed.
In the spring of 1930, after 2 years of almost totally unsuccessful
prospecting, the Shell Oil Company terminated the California contract.
The ensuing disappointment and the tendency to blame the economic
depression could not alter the fact that the spire method, after prolonged
large-scale operations, revealed false anticlinal axes along the trace of
hards beds (see p. 61). This was also the end of operations in all other
fields, electrical coring included, in the United States and in Canada. The
last activities had been in Joplin, Missouri.

Electrical prospecting in the U.S.S.R.

In 1929, despite the fact that the U.S.S.R. had already established a
large scientific organization, which had conducted extremely valuable
theoretical research, the country was lagging in the development of
geophysics. Petroleum exploration by geophysics had been limited to the
torsion balance, used mainly in the Emba Basin north of the Caspian Sea,
where a number of salt domes, some with small production had been

77
The History of a Technique

located. Nothing had been done in seismic exploration at a time when first
refraction and then reflection methods were already in use in the United
States and elsewhere.
Geological considerations seemed to explain the lag in part. In the
United States the main areas of production were the Mid-Continent26 and
the Gulf Coast. The former was characterized by smooth structures and
stratigraphic series with compact beds (limestones) which offered good
reflecting horizons; on the latter, the formations were mostly associated
with salt domes, lending themselves to seismic refraction and gravimetry.
By contrast, most of the U.S.S.R.’s production depended on the deposits
of the northern Caucasus and Baku, located in generally soft and steeply
dipping formations hardly suitable for the seismic prospecting methods of
the time. Furthermore, since these deposits were found in anticlines with¬
out a salt core, gravimetry was practically useless. Such overall conditions
were not unlike those in California, where seismic and gravimetric activi¬
ties had been abandoned.
Yet there was a pressing need to undertake extensive geophysical
exploration, especially for Grozneft, the petroleum trust of the Grozny
region north of the Caucasus. The large prewar production of the “Old
Fields” was decreasing after 20 years of exploitation. Although the drop in
production was compensated for by recent discoveries of the same mag¬
nitude (the “New Fields”), additional reserves had to be found. With little
success from geological prospecting and after several disappointing
gravimetric surveys, Grozneft turned to electrical prospecting. On the
basis of its experience in California, Schlumberger seemed to be the
organization most qualified in this technique. A 2 year contract, signed in
July 1929, provided for the utilization of resistivities and of the spire,
together with the testing of electrical coring. An important provision was
that Russian crews were to be gradually trained by French instructors.
The mission arrived in Grozny in early September, led by Vahé
Melikian.27 Operations started in the immediate vicinity of the Old and
New Fields, marked by hills stretching in two long lines almost parallel to
the Caucasus chain. By November a first resistivity map already showed

2I> The region extending from the Rockies on the west, the Mississippi River on the east,
Iowa and South Dakota on the north, and northern Texas on the south.
27 A young engineer of "La Pros,” Melikian, a native of Baku, had had a leading role in the
conclusion of the contract.

78
Surface Electrical Prospecting

clearly the eastward extension and closure of the anticline. Moreover, the
spires traced the axis of the anticline and confirmed the steady dipping of
the formations. The Russian geologists, used to retiring to their offices as
soon as fall came, were dumbfounded when they saw the French prospec¬
tors—Raymond Sauvage, André Poirault, Jean Lannuzel, Roger Jost, and
Charles Scheibli—proceeding with their work throughout the winter in
worn-out vehicles, handling kilometers of cables over snow-covered hills
and steppes or through vast quagmires. Such tenacity, together with
Melikian’s skill and perseverance, was rewarded by an expanded con¬
tract, this time with Soyuzneft, the central petroleum organization. In April
1930 a second group of prospectors arrived in the U.S.S.R. Still others
followed; some, like Castel, stayed for several years. In fact, until 1936
almost the whole of “La Pros” took turns going to the U.S.S.R.
The spring and summer campaign of 1930 involved mainly the pros¬
pecting of the Terek steppe, between the river and the Manich Basin to
the north. From a borehole the thickness of the overburden had been
estimated at several hundred meters, a depth that first excluded the use of
the spires and then required resistivity equipment with a deep range of
penetration. A series of north-south resistivity profiles with 4000 meter
spacing was undertaken up to distances of 100 to 150 kilometers from the
Terek. The term “dragging” previously used for this kind of operation was
not appropriate here since, for the first time in the application of this
technique, the cable was no longer dragged from station to station, but
was laid and rewound in 400 meter lengths by means of truck-mounted
reels. Most of the survey area was covered by dunes, and travel was very
difficult; the sand gave high resistance to the ground electrodes, and the
sensitivity of the readings was sharply reduced. Additionally, electrofiltra¬
tion and telluric currents kept the needle of the potentiometer in con¬
tinual agitation. Whereas in a field offering easier travel and better electri¬
cal grounding a team could produce 25 kilometers of profiling daily, only
3 to 4 kilometers could be completed, and that arduously, in the dunes.
By the fall of 1930, however, a vast territory had been prospected in
the Grozny perimeter; the resistivity map showed a remarkable concor¬
dance with the geological map and had located all known anticlines.
Therefore, in Grozny and in Moscow, Marcel believed he could safely
state that, barring oil-bearing horizons beyond drilling range, the exis¬
tence of other anticlines need not be considered.
In Baku, where operations started at the same time, geological condi-

79
The History of a Technique

tions did not allow for reconnaissance surveys as extensive as those in


Grozny. The huge Kura Plain south of Baku, the proposed site of a search
for deep structures, covered a subsurface composed of loose sands and
clays impregnated with salt water, where there seemed to be no layer of
sufficient thickness and contrast to constitute a good electrical marker.
Hence renewed attempts at electrical prospecting in this plain yielded
results of dubious merit, and surface operations were confined to the
Apsheron Peninsula, where limestone beds contrasted sharply with the
surrounding clays.
One of these operations, offshore from the Bibi-Eibat field close to
Baku, constituted the first example in history of underwater geophysics.
The Bibi-Eibat structure, an anticline whose trace is marked by a lime¬
stone outcrop, was partially submerged in the sea. The sea had gradually
been land filled as the field was developed. It was necessary to know how
far this filling had to be extended—in other words, to determine the
closure of the formation under the sea by plotting the extension of the
limestone. The survey was made in a few weeks during the spring of 1931,
using equipment borrowed from the electrical coring operation with
spacings of AM = 45 meters and MN =10 meters. The winch was on
shore. For each profile the cable was pulled offshore perpendicular to the
coastline, and the measurements were made, point for point, while it was
being wound back. The plotting showed resistant bands outline the
underwater trace of the limestone bed (Fig. 25). When the operation was
resumed 3 years later, the Russians enlarged the resistivity map by a band
several kilometers wide bordering the peninsula. This they did by drag¬
ging the cables on floats towed by a boat carrying batteries and poten¬
tiometer; in fact, this was the primitive precursor of the system that was to
be used much later in underwater seismics.
In the Emba Basin some results were obtained by detailed study of
half a dozen salt domes, but the measurements required exceeded the
capabilities of the method because of the highly conductive formations
and the hardships of working year round in a semidesert with extreme
climatic conditions. It is no wonder that the accuracy, efficiency, and yield
were far behind those obtained in Alsace.
When Conrad arrived in the U.S.S.R. at the end of 1931, the overall
balance sheet of the operations, especially electrical coring,28 showed

28 Electrical coring in the U.S.S.R. is discussed in Part Two, “The Origins and Develop¬
ment of Logging.”

80
Surface Electrical Prospecting

Legend
1.5
Equal resistivity curve

-Off-shore electrical profiles


-Trace of limestones off-shore

Limit of limestones on land.

Cape '
Shikhov

Cape
Bailov

1 \
Caspian Sea

Figure 25. The first known example of underwater geophysical prospecting: tracing through
resistivity measurements of the limestone beds demarcating the Bibi-Eibat structure under the sea
(Baku area). (Reproduction from a reprint of Science et Industrie, nonserial issue, “La
technique des industries du petrole,” 1932.

brilliant success. One contract followed another until 1936. This collab¬
oration enabled the Russians to expand the use of electrical methods
both on the surface and in boreholes. In the 1930’s operations were
extended to Georgia, the Maikop region, and central Asia, and later to the
West Ural and Volga basins, to the Donbas (coal), and finally to the region
of Lake Baikal, Sakhalin, and Kamchatka. In addition, in 1935 Schlum-
berger played a decisive role in the launching of seismic reflection, which
ever since has been applied widely in the U.S.S.R. Today the crews
assigned to surface electrical prospecting and logging in the U.S.S.R.
number in the hundreds, and many geophysical laboratories and research
institutes can be found all over the country. In no other place in the world

81
The History of a Technique

does there exist such a massive utilization of electrical processes, while


elsewhere seismic reflection plays the leading role in petroleum explora¬
tion.
The success of Schlumberger in the U.S.S.R. was offset by a great
tragedy; it was during his return from a trip to Moscow in May 1936 that
Conrad suddenly died. His premature death deprived the Company of a
prestigious chief whom everyone loved and the geophysical industry of a
scientific mind that was one of the most original and creative of the era.
Marcel then took over the management of the Company.

After the establishment of the Compagnie Générale de


Géophysique

In 1929 the Banque Mirabaud and French mining companies had


become the licensees of the German Ambronn and Selfeld processes of
electromagnetic prospecting and seismic refraction, and had established
the Société Géophysique de Recherches Minières (S.G.R.M.). Hit almost
immediately by the economic depression, this company sought to make
an arrangement with Schlumberger, which led to the creation of the
Compagnie Générale de Géophysique (C.G.G.) in March 1931. The new
corporation combined the management of both partners’ operations ex¬
cept for the Russian contracts and electrical coring. Raymond Maillet, the
general manager of S.G.R.M., brought his highly scientific expertise to
the group.
The structure of C.G.G. was to change several times. In 1935 the
Société de Prospection Geophysique, a joint venture of Compagnie Fran¬
çaise des Pétroles and the Banque de Paris et des Pays-Bas, merged with
the C.G.G. group, contributing gravimetric prospecting with the Eôtvôs
balance. With the acquisition of the license of American seismic reflection
processes in 1934, C.G.G. was equipped to handle any geophysical
method. Later, having given up its managing function, it became an
operating company. After World War II, under the leadership of Léon
Migaux, assisted by J. Cunin, it became the second largest geophysical
company in the world. Seismic reflection was the key to its expansion, first
under French programs and then on an international scale. Schlumberger
withdrew from C.G.G. in the 1950’s.
At a time when electricity still accounted for the bulk of C.G.G.’s

82
Surface Electrical Prospecting

activities, some outstanding results were achieved, among them the dis¬
covery of the Bucsani anticline in Rumania (1933) and the geological study
of the Digboi anticline in the Assam jungle (1936). This period also
marked the dawn of the application of the electrical method to groundwa¬
ter exploration. Several important surveys were conducted under the
personal supervision of Schlumberger: continuation of the survey in the
Alsatian Plain, underwater prospecting in the Algiers roadstead, corrosion
studies of buried pipes, and measurement of telluric currents.

In 1931-1932 prospecting in Alsace involved resistivity profiles with


6000 meter spacings and stations 500 meters apart. This was a region
crisscrossed by roads, railroads, and canals and covered by farmland and
forests, where it was difficult if not impossible to stretch the cables in a
straight line as could be done in the semidesert plains of California or the
U.S.S.R. With consideration for the topographic constraints, the location
of the ground electrodes was determined on the map. Once the electrodes
had been planted, a truck-mounted automatic spooler connected them,
along the most practicable route, to each other, to the batteries, and to the
potentiometer. The stray currents created by the trolley lines often re¬
quired nighttime operations.
This prospecting was to elucidate the salt structure of the Alsatian
subsurface. Resistant bands on the map revealed four saline ridges
parallel to the Rhine River and a transversal ridge. With an accuracy that
was confirmed by drilling, electrical soundings assessed the depth of
the salt at about 500 meters. Simultaneously, Schlumberger undertook
balance surveys, especially around the Hettenschlag Dome, but the re¬
sults, although generally in accord with those of the electrical mea¬
surements, appeared less complete with regard to depth assessment. Also,
the balance proved slower than electricity. These resistivity mea¬
surements were remarkable with respect to salt tectonics and seemed to
be superior to results achieved by seismics on the Gulf Coast. One might
ask whether, under the circumstances, an opportunity to successfully
resume operations in the United States had not been missed. In all
likelihood, with gravimetry and seismics by then firmly rooted in that
country, the chance of developing a substantial resistivity market would
have been remote.

In 1932 a new underwater survey of about the same nature as that in

83
The History of a Technique

Bibi-Eibat was made. This survey was more difficult, however, because it
had to determine the thickness of the silt deposited on the bottom of the
Algiers roadstead, which lay 30 meters under water of a much higher
salinity than that of the Caspian Sea. Consequently, there was a risk that
this would reduce the sensitivity and the accuracy of the measurements.
However, the survey was facilitated by the fact that the sea and the silt
were homogeneous media of known resistivity, while that of the underly¬
ing rock was practically infinite. A special kind of electrical sounding was
required for this operation, whereby the cables were dragged on top of
the silt. Thanks to a chart of theoretical curves calculated for this configu¬
ration, the measurements could be correctly interpreted.

During the same year Conrad thought of using his methods to solve
the problems arising from the corrosion of buried pipes (e.g., liquid or gas
pipelines, or the sheathing of telegraph or telephone lines), which were
inadequately protected by the insulating coatings of the time. Thorough
studies, among them those of the U.S. Bureau of Standards, had revealed
the crucial relationship between the nature of the soil and corrosion.
Because of the mineralization of the impregnating waters, the metal was
chemically etched and eaten away in places. Even worse was autogalvanic
corrosion. Pipelines buried in soils more or less aerated, or containing
different salts, are the source of spontaneous currents. At the spot where
the current leaves the pipe, metallic ions are diffused in the surrounding
water, thus progressively eroding the pipe. A still more destructive action
is caused by stray currents. This action usually occurs in pipelines located
in urban areas near industrial installations which generate the currents.
The study of these phenomena had entailed many potential mea¬
surements with electrodes placed very close to the pipes, requiring the
digging of holes or trenches at a cost that would have been prohibitive on
a larger scale. Conrad tackled the problem with surface measurements.
From laboratory experiments he was able to refine and complete the data
already published, and to define the methods suited for the various phases
of the problem, a description of which follows.
The layout of a pipeline could be determined with the help of a
resistivity map plotted with very short electrode spacing (a few meters)
along a strip straddling the proposed route. Where possible, the pipe had
to skirt the most conductive, or ionically active, zones, and to avoid
crossing the boundaries between zones of different conductivities. The

84
Surface Electrical Prospecting

process was used in the U.S.S.R. for laying an oil pipeline 1000
kilometers long between Guriev, north of the Caspian, and Omsk in
southwest Siberia (June-November 1932). For urban piping systems the
electrical field of the stray currents was measured throughout the indus¬
trialized region with an apparatus similar to that used for the tellurics (see
p. 88). Since the resistivity of the ground was determined at the same
time, the value of the current density at each point of observation permit¬
ted a more effective planning of the pipe-laying operation.
Delineating the zones that were corroded or corrosion threatened by
the autogalvanic effect required that potential profiles be run straight
above the pipes between one nonpolarizing electrode and another fixed-
reference electrode. The potential peaks pointing to the zones of outflow¬
ing current indicated the need for adequate protection. Since this mea¬
surement of potentials could not apply to pipes exposed to erratic stray
currents, the current outflow zones were located through a specially
designed differential device. These methods were applied to a few urban
centers in the early 1930’s.

In geophysical language the term “tellurics’’ refers to irregular cur¬


rents that flow hundreds and thousands of kilometers through the crust of
the earth, usually in the form of nearly horizontal sheets, and reflect
electrical phenomena of the upper atmosphere. They are distinct from all
other currents flowing through the ground: the strays, often stronger but
spatially more circumscribed; the currents of spontaneous polarization
confined to the vicinity of the bodies where they originate; and localized
currents produced by electrofiltration in zones where, for example,
mineralized water has seeped upward toward the surface of the ground by
evaporation.
Until about 1920 the scientific work published had consisted of
measuring the potential differences created by the tellurics between
widely spaced (several tens of kilometers) ground electrodes, generally by
using out-of-service telegraph lines. With such long spacings the potential
differences reached several volts, and the measurements did not require
especially sensitive instruments. However, because they encompassed vast
areas, these measurements did not permit an adequate analysis of localized
soil resistivity effects. Conrad tried to determine these effects by using his
potentiometer and rather short electrode spacings.
In 1921 he instructed the crew operating near Lobsann in Alsace to

85
The History of a Technique

observe the telluric currents on both sides of the great Rhenish fault,
which brought conductive shales into contact with the resistant Vosges
sandstones. On each side of the fault and perpendicular to it, two lines of
the same length were connected to two galvanometers whose operators
made simultaneous readings. These very brief observations confirmed
that the tellurics varied constantly and that their average amplitude was
much higher for sandstones than for shales. However crude, this was the
very first indication of the effect produced by a geological feature on the
telluric field.
Other observations were recorded by Conrad in several unpublished
papers written in 1922-1923. One of these describes the observation he
made in the course of an experiment in the Cherbourg roadstead (May
1922) in order to locate metallic wrecks through the spontaneous polari¬
zation phenomena they produce.29 His conclusion had been that the
conductivity of the seas, high in relation to that of the continents or of
submerged rocks, could be assumed to play an important role in the
overall distribution of the telluric currents over the surface of the earth.
“We hardly can be wrong,” he wrote, “in stating that, up until this time, all
available documentation on this new aspect of the telluric currents (which
we shall call pelagic currents, since the term ‘marine currents’ has quite a
different meaning) does not go beyond the few observations we made
offshore the Cherbourg roadstead in May of the past year.” He was
considering measurements made at sea, from which oceanographers could
draw valuable data on submarine topography: shoals, continental shelves,
abyssal troughs, and so on.
The first systematic study of the phenomenon, as reported in another
paper, took place at Paray, Haute-Marne (June 3-9, 1922). The electrode
configuration consisted of two lines, each 150 meters long, connected at
each end with nonpolarizing electrodes and oriented west-east and
south-north, respectively. As in the Lobsann experiments, the readings
were made on two galvanometers. The plotting of the many curves
showed rather irregular pulsations with a period of 30 seconds to 1
minute, simultaneous on both lines, and superimposed over movements
of much longer duration (several hours). Vectors proportional to the
amplitude of the pulsations recorded between two instants on the two
lines were marked on rectangular coordinates, and the resultant gave the

A submerged metallic hull forms a huge pile because it is made of diverse metals.

86
Surface Electrical Prospecting

average direction of the tellurics, as well as a value proportional to their


intensity over the same time interval. Conrad had observed that this
direction had a tendency to line up with certain azimuths, varying accord¬
ing to the time of day. Furthermore, simultaneous measurements using
two identical cross-shaped configurations separated by about 2 kilometers
had shown concordance of the pulsations at both stations.30
Still another unpublished paper set forth the basic ideas of prospect¬
ing by tellurics. These currents, if they spread through formations of
uniform resistivity, would take the form of sheets of parallel lines. It was
as if they had been created by artificial A and B sources several hundred
kilometers apart, with the AB alignment and the intensity of the current
varying constantly. Since the soil is not homogeneous, its configuration
and consequently its average resistivity affect the magnitude and the
direction of the telluric vector at any moment and at any point of the
surface. However, since neither the length of the imaginary current
emission line nor the intensity of the current is known, no absolute value
of the resistivity can be obtained from telluric measurements; the study is
limited to its spatial variations. To this end the measurements made at a
mobile base moved from station to station across the area being pros¬
pected are compared with those recorded within the same time intervals
at a fixed reference base. The technique amounted to a resistivity map
made with very long spacings and lightweight equipment, like the one
used for the spontaneous polarization surveys. Although easy to carry out,
the method did not match the current emission techniques where the
spacings and the direction of profiles could be selected to adapt to the
characteristics of a particular structure. Since only operations with shallow
depth of investigation and spacings of a few hundred meters were then
under consideration, experimentation with tellurics did not go any further
at the time.
After several years of experience had demonstrated how difficult it

30 These were the manifestations of the tellurics proper. Variations would occur from
time to time, reflecting the meteoric origin of these currents: lightning (abrupt induction
effect on the lines), wind, rain, sunshine (change in the evaporation rate of subsoil waters,
and hence in the amplitude of the electrocapillary potentials). Conrad’s reflections on the
latter phenomena, their mechanisms and their behavior according to topography, tempera¬
ture, sunshine, vegetation, and so forth, actually touched the field of physics of the globe;
they supported the original assumption that electrofiltration might be the cause of the
negative charge of the terrestrial globe.

87
The History of a Technique

was to operate with long current spacings, the measurement of tellurics was
given fresh consideration in 1934. Marcel, always in favor of lightweight
and convenient equipment, furthermore perceived these measurements
as an expeditious way of conducting large-scale reconnaissance. The much
costlier seismic reflection method would then be used to detail zones
demarcated by the anomalies of the tellurics map.
As Conrad had indicated in his notes, the technique involved one
fixed and one mobile base. The hand recorders used were adapted from
the electrical coring. The readings on each mobile base took 10 to 15
minutes. Even over long distances (i.e., 10 kilometers), there was excel¬
lent correlation of the pulsations observed at the two bases. Since ex¬
perimentation had shown a linear vectorial relationship between these
recordings, it was convenient to scale the vectors plotted at each station,
taking the synchronous vector of the fixed base as a unit: the ends of these
vectors would define ellipses whose areas were proportional to the aver¬
age resistivity of the ground.
One of the most significant surveys of that time took place near
Hettenschlag in Alsace, the site of many previous resistivity measure¬
ments (Fig. 26). At each measuring station the vectors indicated the aver¬
age direction of the tellurics: they were channeled between the Vosges
in the west and the Black Forest in the east, flowing between the sur¬
face and the electric insulator constituted by the salt. The orientation of
these currents was nearly parallel to the Rhine River except in the vicinity
of the north-south saline ridges, which they crossed frontally.
This study and others showed that recording by hand was both
inconvenient and costly (four permanent operators), and that the future
success of the method required automatic equipment of the type already
devised in connection with the first photographic recorders used in log¬
ging. The underlying principles, the housings, and the optical systems of
the two recorders were the same, except that for tellurics the galvanome¬
ters were more sensitive and slower and the film was unwound by an
automatic timing device.31 These instruments were ready for field use
when World War II broke out.

31 This category of measurements was not based on the null method as with the poten¬
tiometer, but on the deviation of the galvanometers: the total resistance of the circuit,
ground electrodes included, was adjusted for each base to a common known value, which

88
Surface Electrical Prospecting

Figure 26. Telluric current survey of the Hettenschlag area. The solid black lines show the
outline of the saline ridges.

After the German invasion in the summer of 1940, part of the staff,
with equipment from S.P.E. and C.G.G., gathered at Saint-Gaudens,
Haute-Garonne. Only a few months earlier, a large gas-bearing formation
had been discovered in a well located nearby on the Small Pyrénées chain,
a line of hills parallel to the main chain and reflecting a series of closed
anticlines. As a result a new field was being developed near the village of
Saint-Marcet. A discovery of such importance led one to believe that the
region contained other deposits. Although a few structures were visible
from the surface, geophysics was required on the search for those that
were not. On the basis of the available geological data, Marcel assumed
that the region was favorable for electrical prospecting and proposed its

made it possible to calibrate the deviations in potential-difference values. The adjustment


of the circuit was possible thanks to an accurate electromotive force calibration standard
equal to 1 volt.

89
The History of a Technique

application, mainly by tellurics. Thus, as early as 1940, a large part of the


Aquitaine Basin was prospected by this method, which clarified several
structures and traced their outlines.32
Part of the region was also prospected by resistivities with long
spacings. The technique (basically like that used in Vitré; see p. 56) was
to reverse the current at regular intervals of a few seconds and to record
on film the potential differences between M and N with the instruments
used for tellurics. The measurements, each a few minutes long, were
plotted in diagrams against a background of fluctuations created by noise,
where sharp deviation appeared at each reversal of the current. The
average amplitude of these deviations gave the value of the potential
difference generated by the current emission. The “kicks” resulting from
induction and skin effect were easily eliminated by visual examination.
This was a safer and more convenient process than the one provided by
traditional measurement with the old potentiometer.
During the same years prospecting by tellurics was extended to the
Languedoc, Bresse, and Bas-Dauphiné basins, as well as to part of the
Rharb Plain in Morocco. It proved especially valuable in the detection of
uplifted structures hidden by overlaps: outcrops, faults, steeply dipping
anticlines, piercement folds, and the like, where seismic reflection often
fails.
In charge of surface prospecting and logging in southern France
between 1943 and 1946 was Eric Boissonnas, whose role in the develop¬
ment of tellurics was outstanding. Later he sought to apply his experience
with the method in the United States, particularly on the Haynesville
Dome in Texas, which had already been thoroughly explored by drilling.
Its shape, as outlined by the telluric survey, was in keeping with the data
provided by earlier surveys. In spite of these results, however, tellurics
did not succeed in gaining recognition in the United States as a tool for
petroleum exploration.

After World War II C.G.G. continued to improve the resistivity


method. Theoretical studies undertaken at the highest level included the
calculation of charts for multiple horizontal layers, the treatment of cer-

32 In southern part of the basin, the measurements were perturbed by intense stray
currents from the Bayonne-Toulouse railroad line, to the point of making the values
geologically meaningless.

90
Surface Electrical Prospecting

tain cases of dipping formations, and the use of computers for the in¬
terpretation of electrical soundings with electrode spacings reaching 10 to
12 kilometers. From 1957 to 1958 the investigation of saline ridges was
again resumed in Alsace and was extended to Baden and Wurtemberg. In
addition, C.G.G. undertook several telluric surveys, particularly in the
Parisian Basin (a detailed exploration of its eastern part was conducted in
1949) and in Algeria (Hodna Basin, 1948-1950). These surveys contrib¬
uted to the study of certain aspects of the phenomenon by mathematical
treatment, on laboratory models, and under specific field conditions. One
survey was conducted with bases thousands of kilometers apart
(France-Madagascar-Gabon), between which the recordings showed cor¬
relations (1945).
From the 1950’s on, surface electrical prospecting was used chiefly
outside the field of petroleum exploration. The resistivity method was
applied in studying the configuration of captive aquifers, in locating
contacts between fresh and brackish waters, in assessing the thickness of
alluvial deposits containing groundwater tables, and so on (France, Italy,
North Africa, and certain arid countries). The electrofiltration
phenomena were put to use for the application of spontaneous polariza¬
tion measurements in the reconnaissance of privileged water circulation
zones. Electrical methods entered the geothermal exploration field in
Larderello, Italy, and later in Kizilder, Turkey. One of the most important
spontaneous polarization surveys for metallic ores took place in
Mauritania (copper sulfide and magnetite).
Induced polarization, which had caught Conrad’s attention as early as
1912, was the subject of important theoretical research work and was
used in prospecting for pyrite, blende, and galena deposits. Finally, sur¬
face electrical prospecting was applied in civil engineering (dam founda¬
tions) and public works surveys.

91
PART
TWO

The Origins
and Development
. ' of Logging
INTRODUCTION

P etroleum and natural gas occur within the pores of certain perme¬
able formations (sands, sandstones, and limestones) of sedimentary
origin. Also contained in a sedimentary series are other formations such as
shales which, though porous, are practically impermeable, or rocks al¬
together compact like hard sandstones or limestones, gypsum, rock salt,
and flintstone, as well as all the intermediate types: shaly sandstones,
sandy clays, marly limestones, silicified clays, and so on. Vertically there is
a very great variety among all these sediments, whereas laterally, at least
within the same geologic unit, they may retain almost constant lithologi¬
cal characteristics over wide areas. Moreover, should these characteristics
change laterally—say a limestone bed becomes marly—it often happens
that the fossils remain the same.
Petroleum is produced by the transformation of certain live or¬
ganisms buried, together with very fine sediments, in (generally) marine
waters, and gradually compressed under the weight of further deposits
accumulating on top. Formed between mineral grains and then in large
part expelled by compression, petroleum travels through the openings
available—permeable layers, faults, fractures—until it reaches the sur¬
face, where its traces soon disappear unless stopped inside a permeable
layer by a barrier of impervious rock. If, for example, the permeable layer

93
The History of a Technique

were folded, all the petroleum that could penetrate it would rise above
the water toward the upper part of the fold or anticline. Such a configura¬
tion is called a structural trap. If there were also gas, it would move to the
top of the petroleum, in which case the reservoir rock would, from
bottom to top, enclose oil and gas. The real situation is more complex; a
certain proportion of water remains in the pores by capillarity even at the
very top of the structure. Structural traps are provided by other configura¬
tions: the top of a dipping permeable layer may be sealed off by clays
brought about by a fault, or it may butt against the sheer flank of a salt
dome.
On the other hand, a permeable layer may contain an increasing
amount of argillaceous material and become altogether impermeable at
the top, or it may taper off and disappear. Many petroleum deposits are
also found in sand lenses or other reservoirs completely surrounded by
shale. These are the so-called stratigraphic traps.
After surface geological and geophysical surveys have established the
presence of potential oil- or gas-bearing structures, the only way to
ascertain that these structures actually contain such deposits is to drill one
or several exploratory boreholes.1 The purpose of these boreholes is to
identify the productive zones, if any, and to provide overall geological
data on the formations penetrated. From the fact that the lithological and
paleontological characteristics of the formations remain laterally constant
it is possible to establish correlations among boreholes and to determine
the structural features of a deposit or a basin.2 Furthermore, exploration
may sometimes be guided toward the discovery of stratigraphic traps by
comparing borehole data from several wells, including rock lithology, dip,
thickness, and fossils.
Once reconnaissance has located a commercial deposit, a production
program is drawn up involving the drilling of a certain number of addi¬
tional wells, called development wells. Exploration holes in the discovery
area are usually converted to production.

In the 1920’s numerous wells, particularly in shallow areas of the


United States, were still drilled by the old cable tool method, in which a

1 Very recent developments in seismic reflection make it possible to detect, in certain


favorable cases, the presence of a gas deposit.
Data gathered from the observation of outcrops can be rapidly complemented and
refined by drilling small-diameter boreholes no deeper than a few hundred meters.

94
The Origins and Development of Logging

heavy steel ram was projected at the lower end of a cable. However, for
the past 20 years or so, the cable tool process had been largely supplanted
by the rotary method with its much higher drilling speed and ability to
drill much deeper.3
The drilling tool employed in the rotary method is called a bit. It is
rotated like a twist drill at the end of hollow drill pipes and is driven by
surface machinery. To begin with, the bit is screwed to the bottom of the
first drill pipe; as it penetrates the soil, a second length is screwed to the
top of the first, then a third, and so on to the depth desired. The derrick
—formerly a wooden structure, now steel—allows vertical handling
through a crown block and elevators of pipe lengths from 60 to 90 feet.
The drill pipe is rotated through a huge gear called the rotary table, which
is driven by steam, electrical, or diesel power. In the course of drilling,
water, charged with various ingredients and called mud, is pumped
through the drill pipe down to the bit; it then returns upward through the
annular space between the pipe and the wall of the borehole. This mud is
composed mainly of suspended clay, but also contains a number of min¬
eral and organic additives which, dispersed or dissolved, give to the mud
its desired qualities: density, viscosity, and colloidal and chemical proper¬
ties, according to the specific conditions of the borehole (depth, tempera¬
ture, diameter, nature of formations drilled through, etc.).
By strengthening the walls and applying hydrostatic pressure on the
formations, the mud makes it possible to drill over hundreds of feet
without the need of lowering a protective casing. It also lubricates the bit
and the drill pipe and carries the fragments cut out by the bit to the
surface. The density of the mud is adjusted in such a way that, barring the
unforeseen, its pressure at the level of each permeable horizon remains
higher than that of the formation fluids—water, oil, or gas. Thus the mud
prevents uncontrolled blowouts with their disastrous consequences, such
as destruction of installations and fire.
After setting a short section of large-diameter conductor pipe, drill¬
ing begins with a borehole about 12 to 16 inches in diameter. At a depth
of 800 to 1200 feet, a first string of casing (surface pipe) is lowered into
the hole and sealed against the wall by the injection of cement. This casing
is usually set below all freshwater sands to protect them from borehole
fluid contamination. Drilling is then continued with a bit diameter of

3 In the 1920’s the depth record was 7500 feet; it is now 30,000 feet, drilled in 2 years.

95
The History of a Technique ---

about nine inches to the depth determined from the geological conditions
or the location of the pay zones drilled through or hoped for, after which a
new string of casing is lowered and cemented. According to the circum¬
stances, drilling ends at this point or is continued to the final objective.
Before electrical logging was known, the only expeditious way to
identify formations was to monitor the rate of drilling, thus obtaining
some idea of the relative hardness of the rocks, to observe the mud (for
level drops, changing salinity, gas bubbles, or oil droplets), and to exam¬
ine the cuttings as they reached the surface. With the techniques then
available these procedures were vague and unreliable. The most accurate
one was mechanical coring. In this operation special crown-shaped bits cut
cylindrical samples a few feet long called cores, which are drawn to the
surface for examination. Although valuable, the data gathered by mechan¬
ical coring have shortcomings. The frequent retrieving, lowering, and
disassembling of the tool, as well as the whole string of drill pipe, make
drilling slower, more complicated, and more costly. Furthermore, the
cores are often crushed by the tool, and in very soft formations core
recovery remains incomplete. Also, cores coming from permeable rocks
are contaminated by the mud and, while the tool is being retrieved, lose
most of the oil or gas they may have contained, thus giving only a
distorted picture of the actual formation. Neither is it always easy to
obtain from the cores stratigraphic data which can serve as a basis for well
correlations.
These shortcomings were more acute 50 years ago, when the
techniques available for the handling and observation of cores had not
reached their current degree of refinement and accuracy. It may be added
that, to limit costs and risks, operators endeavored to core only where the
presence of key horizons or pay zones could be anticipated from geologi¬
cal data and drilling incidents. Such a procedure was often disappointing;
the cored lengths could be devoid of interest, or the cores could have
been taken too late, after the promising formations had already been
drilled. There remained room for much improvement.

When electrical coring was invented, most of the oil produced came
from fields in Tertiary basins made up of sands, sandstones, and shales
that were generally unconsolidated, with here and there a few compact
beds. Oil fields in Venezuela, California, on the Gulf Coast of the United
States, in the Caucasian provinces of the U.S.S.R., in Rumania, and in

96
The Origins and Development of Logging

Indonesia were of that type. These were also fields where electrical coring
rapidly became an efficient and economical auxiliary to exploration and
production (Fig. 27, left).
Another type of field, peculiar to the Mid-Continent of the United
States, Mexico, and the Middle East, has since represented an increasing
share of world production as new discoveries have taken place in these
countries and such others as the U.S.S.R. and Venezuela. What is found
there are thick and hard beds, mostly limestone, where in places the
porous, permeable, and usually fissured rock yields petroleum or gas,
often with spectacular outputs (Fig. 27, right). Whereas scant use had

97
The History of a Technique

Figure 28. Schematic cross section of a completed


rotary otl well. Such a layout was in worldwide use
up to the early 1930’s.

98
The Origins and Development of Logging

originally been found for electrical coring in this category of formation,


this situation changed much later when new techniques had improved its
capability and scope.4
When one or several horizons have been identified as pay zones, the
oil well is completed (Fig. 28). In the 1920’s only one horizon at a time
could be produced, usually beginning with the shallowest. The sequence
in this operation was, first, to lower and cement a string of casing reaching
to the top of the pay zone, so as to seal off the upper aquifers (water
shutoff), and, second, to position a perforated liner in the pay zone and
lower a small-diameter tubing. The mud column was then lightened by
swabbing the tubing or by injecting gas or water into it. If the formation
pressure was high enough, the oil reached the surface on its own; other¬
wise it had to be pumped. Most of the same procedures are still employed
today. Only later, presumably in the wake of electrical coring, did perfora¬
tion techniques appear. Perforated liners have been largely replaced by
casing perforations. As a result multiple pay horizons may now be pro¬
duced at the same time; if horizons are produced one at a time, the usual
order is from the bottom to the top.

4 According to prevailing terminology, these two types of fields will hereafter be called
sand-shale series and limestone fields.

99
Phase One: 1927-1932

Pechelbronn: A first resistivity curve

It will be helpful here to look back a few years. In March 1921


Marcel Schlumberger, Eugène Léonardon, Pierre Baron, and Jacques
Gallois were conducting a survey by plotting equipotentials in the Bes-
sèges coal basin near Molières-sur-Cèze, Gard, where a reconnaissance
borehole was being drilled. Marcel recognized there a unique opportunity
to determine in situ the resistivity of the subsoil formations for the
purpose of enhancing the interpretation of the surface measurements.
The same borehole was to serve a little later, in July, for a dip study.
Some 2500 feet deep and cased to about 1500 feet, the hole was full
of water. There are no detailed records of the operation except for the
fact that between March 20 and 22 several resistivity measurements (over
a few feet at the bottom of the hole) did reflect the variations in the nature
of the formations. This was no doubt the first operation ever of this kind
in the history of mineral exploration; it was not followed by such another
until 192 7.1 The time lag can be explained by the fact that every effort was
then being devoted to surface prospecting.

The Pechelbronn company, which for years had been contracting


with Schlumberger for surface work, was eager to know whether resistiv¬
ity measurements in boreholes could be of assistance to its geologists. The

1 Marcel had nonpolarizing electrodes lowered into an available ventilation shaft; one of
their components was . . . flower pots. How he managed their contact with the wall of the
shaft has not been recorded. This episode is mentioned only as an example of the
improvisations often required for the success of an experiment.

101
102
The Origins and Development of Logging

company was especially interested in the location of a key horizon (the


top of a bed of Hydrobiae marls), which drilling sometimes failed to
reveal. This suggestion was welcomed by Conrad, who for a long time
had had a similar idea in the back of his mind.2 This was the starting
point of a project from which electrical coring was to be born. If such
measurements proved feasible, he thought, they would contribute primar¬
ily to refining the geological cross section rather than actually detecting
the petroleum horizons, as scarce and thin as they were in that basin; from
these measurements, furthermore, data could be gathered on the resis¬
tivities of deep formations that would shed light on surface work.
In a note dated April 28, 1927, and entitled, Recherches Électriques
dans les Sondages (“Electrical Research in Boreholes”), Conrad outlined the
principl'e of the new method, which from then on was called electrical
coring (Fig. 29). Doll was given the responsibility for equipment design
and testing.
The actual invention consisted of proving by facts that open-hole
resistivity measurements were feasible and significant—something
which, at the outset, was far from obvious. Some thought that the current
would flow through the mud without penetrating the far less conductive
formations. Although this claim had seemed to be invalidated by the few
tests made at Molières-sur-Cèze, these tests had been too few and had
been undertaken under conditions too special to allow the drawing of a
conclusion. Now Conrad, aside from any experimentation, had demon¬
strated through simple reasoning that the current emitted by the elec-

2 In fact, it seems that the idea of electrical coring originated in Conrad’s mind before any
suggestion was made by Pechelbronn. During a conversation at lunch, in early spring
1927, a mining industry executive emphasized the difficulties encountered in recognizing
bottom-hole formations. Thereupon Conrad disclosed to P. Charrin the idea of using to
that end the recording of resistivities in boreholes, and gave him a sketchy description of
the measuring system. Charrin was entrusted with the preparation of the equipment, but,
because of an urgent mission abroad, he had to turn the project over to Doll. (Communi¬
cation from P. Charrin subsequent to publication of the French edition of this book.)

Figure 29■ Schematic view of electrical coring. Three electrodes. A, M, and N, are lowered into
the borehole. each at the end of an insulated conductor. The current emitted by A flows through the
mud and spreads across the formations. The voltage created between M and N is transmitted to the
surface and measured there. From this measurement and that of the current intensity it is possible
to deduce the value of the apparent resistivity. As in practice no borehole ever is strictly vertical, the
three electrodes are in contact with the wall.

103
The History of a Technique

trode ought to flow through the mud and then spread inside the forma¬
tions.3 Others, among whom were men of repute such as Professor R.
Ambronn, maintained that below a certain depth all geological formations
became altogether compact and hence infinitely resistant. Nothing of the
kind had been shown by surface measurements, but since they did not
penetrate very deeply, positive proof was lacking.
The first electrical operation was undertaken on September 5, 1927,
at Diefenbach in well No. 2905, rig No. 7, by Doll, assisted by Scheibli
and Jost. Nothing would be more eloquent than to quote an excerpt from
the description by Doll 32 years later of this memorable experiment.4

AM was three meters long and MN was one meter long. We made a sonde by
connecting four meter-long sections of Bakelite tubing by means of short
lengths of brass tubing, fastening them to each other with brass screws. The
electrodes were wired to the Bakelite tubes. We contrived a weight, or
plummet, for the bottom of the sonde, making it of one meter of brass
tubing, four centimeters in diameter, and filling it with lead pellets like those
used in duck shooting. It was plugged at both ends and weighed about 25
pounds. The whole assembly looked like a long black snake with five joints.
The cable, if you could call it that, was three lengths of rubber-insulated
copper wire, of the kind used on spark plugs in cars.5 It had a tensile strength
of about 80 pounds per wire. The wires weren’t spliced together, as was
done in later surveys, but were allowed to wind onto the winch drum loose
from each other.
The winch had an X-shaped wooden frame; the drum was made with
wooden flanges and the core of a large Bakelite tube. It was assembled by
long brass bars and nuts. To turn the drum, we had a big pinion connected to
a smaller pinion by a motorcycle chain. The moving axle was steel, with a
bicycle pedal mounted at either end. One of us would get on one side and
one on the other, and turn the pedals. There was a ratchet to keep the drum
from unwinding.
We had no collector. Instead, we had a plug, much like a common wall plug,
at the side of the winch flange. When the winch had to be turned, the cable
connection to the potentiometer was unplugged so the turn could be made.
Then the cable was plugged back in so that we could make the readings.

3 Several years later highly advanced calculations confirmed the correctness of his demon¬
stration, which in the meantime had been verified by thousands of measurements.
4 See Sonde Off, September 1959, p. 22.
5 This was actually the cable then used in surface prospecting.

104
The Origins and Development of Logging

The sheave6 was made of wood with an eccentric axle. It had a long tail as a
counterbalance. This served as our strain gauge.7 We were very worried
about the wires breaking; and by watching the rise and fall of the tail, you
could tell what kind of pull was being exerted on the wires. For depth
measurement, we had a counter on the sheave wheel like the mileage
indicator on a car. We planned to take readings at intervals of one meter.
We made our measurements with a standard potentiometer mounted on a
tripod like those we used in our surface exploration work.
It was a nice fall, a decent day. We drove out to the well—it was Diefenbach
2905, Tower No. 7—in an old station wagon that had been used in surface
prospecting and was completely worn out.
The well was about 500 meters deep which is about 1500 feet. We couldn’t
have gone much deeper, for we only had about 1800 feet of wire.
We began making our measurements. Someone had to unplug the connec¬
tor, someone else turned the winch, someone had to run up on the rig floor
to look at the counter on the sheave . . . there was a lot of running back and
forth. I wrote down the measurements on a pad, together with the depth
reading. Then it was unplug, roll up one meter to the next station, and plug
back in. Make the next reading. And so on, one meter at a time.
At first, we jumped around a good bit, but soon we got the swing of things
and before long, we were able to take about 50 stations per hour. At one
meter per station, that’s around 150 feet per hour surveying time.
The whole arrangement worked well. Everything went off as we had
planned, except for one incident. When we came out of the hole, and had the
sonde hanging in the derrick, we unfastened the weight to remove it from
the bottom of the sonde. We forgot that the upper tube, being hollow, had
filled with mud — which showered all over us — and we got thoroughly
messy. We ended the day by going into the nearby village to take a bath.

In preparing the experiment, Doll had thought of how to control the


electrical leaks that would occur, mainly in the cables after they had been
dipped in hot saline mud under a pressure of several tens of atmospheres.
To this end he had provided the sonde with a relay resistor not unlike the
ones used in surface work. The control consisted in making sure that,
when the current was reversed to actuate the relay and to introduce the
resistor into the circuit in series, the values measured did not vary; the

6 A grooved pulley at the wellhead redirecting the cable toward the winch.
7 This instrument has been named the Roman balance; the sheave was mounted on the
short arm, where the pull of the cable was balanced by the weight of the long arm.

105
The History of a Technique

operation showed that in this first experiment the leaks had not played a
significant role.
Back in Paris, Doll plotted his measurements on a strip of graph
paper and drew the first of the typical diagrams that were to become
familiar to the petroleum industry. A part of this venerable document is
reproduced in Figure 30. It shows a zone of rather uniform resistivities
corresponding to the Hydrobiae marls whose top served as a marker and,
above, a sequence of peaks and troughs where the geological cross section
indicated hard marls, conglomerates, and sandstones. As could be fore¬
seen, the resistivity dropped to zero at the lower end of the casing:8 the
depth at which the drop occurred was within 6 feet of that indicated by
the drillers. In view of the rather primitive metering of the cable, more
could hardly have been expected.
The tests were continued at Pechelbronn under generally difficult
operating conditions, mostly at night when drilling was usually suspended
(Fig. 31). Cave-ins or exceedingly heavy or viscous mud frequently
prevented the lowering of the sonde; and since boreholes are rarely
vertical, the dragging of the sonde and the cable against the wall when
being pulled out resulted in frictional stress close to tensile strength.
Contrary to the prevailing concepts of Marcel, heavier weights and
stronger cables were required. As access to drilling sites was difficult in
rainy weather, an odd suggestion came from Paris: to unload the truck
when stuck in the mud and push the winch like a wheelbarrow, with the
reel serving as the wheel. Needless to say, the first attempt converted the
whole mechanism into a huge mud ball.
In spite of these difficulties and, in particular, the inadequate lengths
of open hole available—only 60 to 130 feet—positive conclusions could
be drawn in the very first weeks of operations. Hard layers appeared on
the diagrams as peaks contrasting clearly with the soft and conductive
marls. From the similarities in the log features, accurate correlations could
be established between boreholes from which the configuration of the
formations could be determined throughout the field. Comparing the mea¬
surements with the cores in a sufficient number of holes allowed the
identification on the logs of most peaks and troughs. Thus electrical
coring gradually came to replace most mechanical coring.

8 Compared to the resistivity of the formations, that of the metallic casing is practically
zero.

106
RESISTIVITES OHMS M2/m

Figure 30. The first electrical coring diagram (Pechelbronn, Alsace, September 1927).

107
The History of a Technique

Figure 31 ■ The winch, cable, and pickup truck used in the beginnings of electrical coring at
Pechelbronn.

Developments in basic equipment

Initial results obtained at Pechelbronn warranted an effort to provide


more advanced equipment for electrical coring. Later, when the method
had proved its value in the large petroleum basins of Venezuela and the
U.S.S.R., it became the main objective of engineering and manufacturing
activities, with surface operations taking second place. This led the man¬
agement to a reorientation of the Company’s technical plans and goals.
This new method opened for Conrad an attractive field of activity;
yet, since it was merely a technical service performed at a day-to-day

108
The Origins and Development of Logging

operational level in support of the oil companies’ drilling activities, it


could not provide him with the same stimulus as geophysics proper, with
its much broader objectives at the very forefront of exploration. The
intellectual interest of electrical coring lay mainly in its multiple
technological and operational aspects, whereas Conrad’s preference was
for research work relevant to physics and geology. This explains why
(although often intervening in crucial decisions involving the orientation
of electrical coring research and techniques), he took personal initiative
only in a few special projects where he felt most at ease, such as the design
of the “guarded monoelectrode sonde”9 and laboratory work on spon¬
taneous potentials. However, in the rare moments not devoted to the
overall management of the Company, he followed surface work with a
keen interest, hoping that he could return to this favored activity once
electrical coring was well on its way and required less of his personal
attention.
The contributions of Marcel and Doll to surface prospecting were
important indeed; but to the former, with his mechanical genius, and the
latter, with his mastery of electricity, electrical coring presented a wider
and more varied range of problems awaiting solution through application
of their respective talents. Most of the technical developments of the
Company were theirs, and they knew not only how to recruit a highly
qualified staff, but also how to give it inspiration and guidance in the
design and manufacture of highly specialized equipment. Only its princi¬
pal components will now be reviewed.

Cables

It will be recalled that the initial operations in Pechelbronn utilized


the cables employed in surface work, tied every few meters with insulat¬
ing tape. The three wires became entangled when being rewound on the
winch, however, and it took great effort to unravel them when again
lowering the sonde into the borehole. Moreover, as it was impossible to
equalize the stress among the three wires, breakages were frequent.
Sometimes only the copper conductor would be ruptured under the
insulation, and then it was necessary to lay the cable flat on the ground to
allow a step-by-step inspection to pinpoint the leaks. To this end the

9 Also called “current output sonde”; see p. 123 ff.

109
The History of a Technique

operator had to slip his hand under the electrically charged cable until a
slight pricking located the point of discharge. All this was time consum¬
ing, clumsy, and inaccurate, and the ensuing frequent delays were not
appreciated by the Pechelbronn company. The prospectors considered
that a cable better suited to the task was indispensable for the successful
marketing of electrical coring. What they were demanding was a real,
industrially made tricable with good tensile strength and adequate insula¬
tion. Doll was promoting the same idea, and the first cable made by
French industry to Schlumberger’s specifications was put into operation in
June 1928. It consisted of three conductors, each made of about 20
stranded steel wires and coated by rubber insulation and tape. In turn,
these three conductors, with an appropriate packing, were stranded to¬
gether and enclosed in a sheath protected by a tarred cotton braid. The
tensile strength of this cable was 1 ton. With it field service improved
immediately as lost time decreased. Thereupon a greater tensile strength
was sought; in 1930 it reached 2 tons for the French-made, and 1.3 tons
for the first American-made, cable. Both had a textile-braid protection
(jute in France, cotton in the United States).
Nevertheless, in spite of tight specifications and detailed inspections,
after a few operations the electrical insulation of the cable was often
reduced to the point where errors in measurements became unacceptable.
Insulation control in the field and localization and repair of leaks there¬
fore became the focus of continuing studies, as witnessed by numerous
articles on this subject published in the early 1930’s in Proselec.10
To measure the insulation of a cable was a rather simple thing, but to
locate the leaks was much more difficult. Improvements in the latter
respect are worth reporting. In 1930 the Paris engineering office had
designed a rather complicated system based on the Wheatstone bridge
principle; its shortcoming was that it functioned properly only when a
conductor developed a single, relatively minor leak. Another process was
tried a little later. After the ends of the conductors had been insulated, the
electrically loaded cable was lowered into a borehole; as soon as a leak
reached the top of the mud, the circuit was closed, and the potentiometer

10 Proselec appeared first in mimeographed form as a technical publication for the use of the
Schlumberger staff only. Initially, it covered the whole range of activities of the Company.
From April 1931 on, it comprised two parts: Proselec Carottage (Proselec Coring) and Proselec
Surface. The latter part was replaced in early 1934 by Ceg'egec, published by C. G. G. Both
publications ended with World War II.

110
The Origins and Development of Logging

deviated. Unfortunately the effective functioning of this system required


sizable leaks, and it proved of little use. In Baku in early 1931, R. Sauvage
and G. Delamotte resorted to the extreme remedy of “arcing” the leaks.
The cable was dipped into a tank filled with water and then connected to a
2000 volt transformer; this ruptured the insulation at its weak points, as
shown by sparks, bubbles, and smoke. This was, without doubt, a radical
process, and not without risk to the operators and the overstrained cable
itself. Furthermore, few crews could find an appropriate installation locally.
Finally, the so-called guard ring method was devised in Paris by
Flenri Doll and Gilbert Deschatre. The operation took place when the
cable came out of the hole. While it was passing between sheave and
winch, its braid well soaked with water, three wire coils (called rings),
wound around a wet sponge, were held by hand against the braid. Each of
these coils was properly grounded, the middle one across the potentiome¬
ter. The cable was connected to a battery; whenever there was a leak
outside the coil system, the current flowed directly through the braid into
the ground without affecting the potentiometer. On the contrary, when a
leak came within the two outer coils, part of the current flowed through
the central coil, and the needle of the potentiometer deviated. This
system was simple, reliable, and accurate and made it possible to pinpoint
a leak within a few inches. It was tested by Doll in the U.S.S.R. during
the spring of 1932 and was utilized by all crews as long as textile-braided
cables remained in use.

Winches

The new winches were practically a personal achievement of Marcel.


The first improvement was to fasten a collector with four copper rings on
one of the flanges of the wooden, crank-driven reel used in the Pechel-
bronn tests. Three of the rings were connected to the ends of the conduc¬
tors wound on the reel. One brush was connected with the battery, and
two others with the potentiometer. The fourth ring was grounded as an
additional precaution against leaks that could occur between the ring of
the emission circuit and those of the measuring circuit.
When operating in boreholes deeper than those at Pechelbronn, that
is, when the cable was subjected to increasing stress, the need arose for
winches of greater capacity and durability. The trend was to build all-
metal reels with steel-plated flanges and an iron core. However, to reduce

111
The History of a Technique

to acceptable limits the induction phenomena occurring in the conductors


wound on the winch at each switching on and off of the circuits, a better
solution was adopted: metal flanges and a Bakelite core. In 1930 and 1931
there were winches of two tricable capacities: 4000 and 7000 feet. They
were electrically driven, with a gearshift and a clutch. When electric
power was not available at the drilling site, the winch was driven by the
engine of the truck. The arrangement was simple and rugged: jack up one
rear wheel of the truck, prop up the chassis, and connect through a flat belt
the tire of the lifted wheel with a pulley at the end of the winch drive
shaft. This was the first model of what Americans, with a smile, were to
call the “French power takeoff.” Someone even had the idea that the spare
wheel of the truck could be used as a pulley! Indeed, as early as 1929,
prospectors arriving in Venezuela had tinkered with a chain drive for the
winch made with a sprocket on the winch shaft and another fixed on a
wheelhub. This system was later to be improved with a powerful clutch
and special props to give firm support to the rear of the truck. The year
1930 also witnessed the first winding device to provide for the regular
reeling of the cable on the winch.
The winch problem illustrates the diversity of equipment then pro¬
duced in the Paris workshop. Such diversity was inherent in the variety of
operating conditions from one country to another (average well depth,
density and viscosity of muds, etc.); it also reflected Marcel’s ceaseless
drive to devise improvements. Thus, although by the end of 1932 there
were only half a dozen logging crews operating in the world (the U.S.S.R.
excepted), nine models of winches were already in operation. (Fig. 32).

Recorders

Doll has described (see p. 104) his first electrical coring operation, at
a time when the measurements involved the same reading techniques as
surface prospecting, that is, point by point, interrupting the upward
movement of the sonde at regular 1 meter intervals (Fig. 33). The varia¬
tions in resistance of electrode A were negligible as compared to those of
the emission line (as for electrode B, since it stayed at a fixed point
on the ground, its resistance obviously remained unchanged throughout
the operation). It was sufficient, therefore, to measure the current just
once at the beginning of the operation and then check at the end to see
that it had not changed: only the voltage, A v (see Fig. 29), had to be

112
The Origins and Development of Logging

Figure 32. Truck-mounted winch (1932).

measured at each station. As it became more and more urgent to seek


continuous measurements without interrupting the upward motion of the
sonde, Marcel designed the first hand-operated recorder, which was to
serve for many years with only minor changes.
In this system the potentiometer and the null method served as
before, but to avoid a situation in which the measurements were affected
by the polarization of electrodes MN and other parasitic effects, the
operation resorted to pulsated-alternating current (see p. 52 ff). The
potentiometer was enclosed in an aluminum casing, on the top of which a
circular plate with a handle could be rotated around a vertical axis.
Through a rather sophisticated transmission gear (a groove in the shape of
an Archimedes spiral milled in the plate was its main component), the
operation of the handle drove the unit and tens knobs of the potentiome¬
ter together with a pencil-holder, whose linear displacements were pro¬
portional to the measured values. Once the current was plugged in and
the pulsator rotating, the operator maintained the potentiometer needle
at zero as the sonde moved upward. The pencil drew a curve on a paper
strip which unwound mechanically in relation to the motion of the cable.
The curve was a continuous resistivity diagram drawn as a function of
depth. The scales selected for resistivity were obtained by adjusting the

113
The History of a Technique

Figure 33. Electrical coring measurement with potentiometer (1929).

intensity of the current and the sensitivity of the potentiometer; for the
desired depth scales the proper set of pinions had to be inserted.
The first recorder was fixed on a console containing the switches,
rheostats, and other accessories for the adjustment of the emission cur¬
rent and the control of the electric motor driving the winch. During a
survey this console rested on a trunk used to store and transport the
various tools (Fig. 34). The whole arrangement was set up between the
winch and the drilling rig. Under the console the cable passed between
two pulleys, one of which drove the cylinders with the paper strip. This
type of recorder was sent in 1929 to the crews in Venezuela, the United
States, and the Dutch Indies and somewhat later to Grozny in the
U.S.S.R.
With such a setup, it could happen that the cable would suddenly
become so taut that the whole apparatus would go flying into the air
before the astounded gaze of the operator. In 1931 an improved version
was introduced, in which the arrangement of the accessories was more
compact and convenient, and the tripod-mounted recorder was located on
the drilling floor; the motion of the paper was controlled by the rotation
of the sheave through a flexible drive shaft. In 1932 the recorder (or,
rather, recorders, since at that time the spontaneous potential was being

114
The Origins and Development of Logging

Figure 34. The first hand-operated recorder (1929). The cable, passing between two pulleys on
the side of the box, controlled the motion of the paper.

measured simultaneously) was placed next to the winch, the flexible shaft
being driven by a set of pulleys built into the winding device.
Formation resistivity being a widely variable parameter (from a frac¬
tion of an ohm-meter for soft shales or saltwater-saturated sands, to tens
of ohm-meters for oil-bearing sands or sandstones, and practically to
infinity for compact rocks such as rock salt or gypsum), the prospector had
to adjust the scale during the recording lest the curve be truncated when
the sonde passed high-resistivity formations, and a blank be left instead of
what were perhaps the most significant values. Such an adjustment was
possible whenever correlation with diagrams from neighboring wells pro¬
vided some foreknowledge of how the curve would look in the well
currently being surveyed. Otherwise, the sonde had to be lowered again
and the measurements resumed on a smaller scale. This had a negative
effect on the duration and cost of the operation; moreover, when the well
was immobilized too long, there was a risk that the sonde could not be

115
The History of a Technique

lowered again, or that the sonde and the cable would become stuck when
coming out.
One elegant way out of the difficulty seemed to be the use of a
logarithmic resistivity scale.11 Some potentiometers were adapted to this
kind of scale, but not all geologists liked it because it weakened the
character of the diagrams. It was abandoned after a few years, only to be
successfully reintroduced 30 years later.
At the end of 1931 the various improvements in electrical coring had
increased the reliability of the method to the extent that Conrad could
state in a communication to the U.S.S.R. Academy of Sciences:

With a good electrical cable, a good winch and good measuring instruments
that are both sensitive and sturdy, it is possible to operate at the bottom of a
2000 meter well and make highly delicate measurements of the phenomena
generating potential differences of only a few millivolts, and involving prac¬
tically negligible amounts of energy. Such a mating of the finest of poten¬
tiometers and a powerful and rugged rotary rig certainly makes for an odd
marriage, but so far everything seems to be going well.

The spontaneous potential curve

In many operations—at Pechelbronn, and even more clearly in


Venezuela and the U.S.S.R.—it had been observed that, even when no
current was being emitted, the potential in the MN circuit varied with the
depth. Various causes were assumed: instability in the polarization of the
electrodes; a somewhat heterogeneous mud column; currents generated
by the casing, since the latter’s oxidization may vary over its whole length
and therefore behave like one or several batteries; the action of stray or
telluric currents; etc. Any of these effects could generate electromotive
forces as a function of time and depth. Conrad and Doll thought that,
rather than compensating for these potential differences or eliminating

11 The displacement of the pencil then becomes proportional to the ratio between resistiv¬
ity values. For instance, when the measurement increases from 1 to 10 ohms, the deviation
of the pencil is 3 centimeters; when it increases from 10 to 100 ohms, the deviation is the
same; and so on. A diagram with a 1 : 1000 ratio can be drawn without changing scale; this
is usually sufficient.

116
The Origins and Development of Logging

their effects, as in the case of resistivity recording, it might be worthwhile


to measure them.
The study began in Pechelbronn in November 1930. Despite the
rather unfavorable geological conditions of this minor oil field (most of
the permeable layers were very thin), it had the great advantage of being
close to Paris; therefore Doll could direct the tests and analyze their
results personally without having to spend all his time in the field and
neglect his many other tasks. Paul Chabas, then head of the Pechelbronn
mission, was in charge of carrying out the measurements.
Rather crude devices rendered the electrodes nonpolarizing.12 The
potendal differences, measured every 20 inches between two electrodes
that distance apart, constituted “gradients.” To facilitate interpretation,
Doll patiently reconstituted the potential curve as it would have been
recorded between a mobile electrode in the borehole and a reference
electrode at the surface. The problem was to make the sum of the
potential differences or, more specifically, to integrate the gradient curve.
The first tests showed that repeated measurements in each borehole
gave a true reproduction of the values, that disturbances (stray currents,
etc.) were intermittent and infrequent, and that, except in the vicinity of
its shoe, the presence of the casing did not manifest itself. On the other
hand, the fact that in these first boreholes the layers subjected to mea¬
surements were mainly shales and hard beds resulted in curves without
relief. Finally, on February 5, 1931, a bed of conglomerate produced a
“kick,” stable with time, which, after gradient integration reached a
minimum of some 10 millivolts, and stood in sharp contrast to the rather
flat line recorded through the surrounding formation (Fig. 35). A con¬
glomerate being a rock made of more or less cemented gravels and sands
and generally permeable, the invasion of the bed by the mud could have
generated an electromotive filtration force; the lines of the current pro¬
duced circulated in the conglomerate and through the adjacent shales and
closed through the drilling mud. The value measured was the ohmic
potential drop due to the passage of the return current along the hole.
Furthermore, according to the mechanism of electrofiltration, the poten-

12 Each electrode was made of a lead wire winding and enclosed in a sleeve of heavy,
fire-hose-like cloth, tied on both sides of the electrode and filled with a saturated solution
of lead acetate with an excess of crystals. When it eventually appeared that polarization was
weak and stable, such precautions proved superfluous.

117
The History of a Technique

Figure 35. Facsimile of a page in Proselec, showing the first diagram where S.P. indicated
distinctly a bed of porous conglomerate (Pechelbronn, February 1931).

118
The Origins and Development of Logging

dal inside the permeable layer increases in the direction of the flow;
therefore the potential in the mud facing the layer was expected to be
negative (Fig. 36). Observations confirmed this theory precisely. Since the
phenomenon had been induced by the contact of a hole full of mud with
the formations cut by the drill, it was, strictly speaking, not a natural one.
Nevertheless, it took place without any artificial source of current and so
was called “spontaneous potential’’ (S.P.).13 Renewed tests in various
boreholes during ensuing weeks confirmed the consistency of the mea¬
surements and showed, on other diagrams, the existence of similar kicks
at the level of permeable layers.
These results were immediately communicated to the crews operat¬
ing abroad; in fact, the entire May 1931 issue of Proselec dealt with them.
The engineers were asked to measure this new parameter whenever
feasible by using one electrode in the borehole and one at the surface for
the direct production of potential diagrams. At that time electrical coring
was in operation in basins generally made up of shale, sand, and sandstone
series, where the many reservoirs reached thicknesses of 15 to 30 feet or
more. Such conditions were much more favorable than those at Pechel-
bronn and produced spectacular results: whereas on the S.P. diagrams the
shales appeared as almost straight lines, the top and bottom of the sands
were marked by sharp deflections, and the departures14 were no longer of
a few units but of several tens of millivolts.
Here was a new and far-reaching discovery. The definition by the
S.P. curve of every permeable layer provided an invaluable complement
to the resistivity curves, which were much less reliable and accurate in the
location of oil-bearing strata than in the definition of correlations among
boreholes. Indeed, although oil and gas contained in a horizon generally
produce peaks in the resistivity curve, compact beds intercalated in softer
formations likewise give peaks. In most cases the ambiguity could be
removed by the S.P.; a resistivity peak without an S.P. anomaly was most
likely to mean a nonpermeable and hence dry horizon. Moreover water¬
bearing horizons and their limits, a subject of concern to the drillers, were

13 It has become customary to use the abbreviation ‘‘S.P." for “spontaneous potential
curve.”
14 The departures are also called “S.P. anomalies,” an extension of geophysical terminol¬
ogy-

119
The History of a Technique

4 -q - k-20-H + mV

I'l 1 O'
1
1 L11
1

ill
j!| —

- - —

Shale-

—- -- -—- — —

• • *» \ » •

1
. ; • * v* • » '.
4—W*-*
Permeablé• •* i?
»» « *»*»»*«*»
VI * » :v

v V:
-,T

d
M* •» \ *

-
a.
(V
Û
—--
Shale-
--
--

Figure 36. Left: Schematic cross section showing mud filtration in a permeable layer and
circulation of currents due to spontaneous potentials. Right: Recording of ohmic drops created by
the current along the borehole.

now marked, an achievement hardly possible with resistivity alone. In


short, the efficiency of electrical coring was substantially enhanced.10
Once the usefulness and potentialities of the S.P. measurements

15 When the S.P. curve was introduced into the oil basins (the U.S.S.R. excluded), it was
presented under the name “porosity curve,” a physical term more meaningful for industry
than “spontaneous potentials.” Indeed, for most professionals, porosity connoted perme¬
ability; it referred to the ability of a rock both to contain fluids and to allow their circulation,
thus excluding the shales, with pores so small that any flow is practically impossible. The
S.P. curve marked precisely the limit between the shales and the porous layers as
understood by professionals.

120
The Origins and Development of Logging

were confirmed, the technique was soon perfected. Adjustments were


made in the circuits so that S.P. and resistivity could be measured simul¬
taneously. To this end, advantage was taken of a characteristic of the
AMNB quadripole expressed in the reciprocity theorem, according to
which apparent resistivity remains the same when the roles of AB and MN
are inverted.16 The electrical coring configuration AMN, with B at the
surface, was thus replaced by an equivalent MAB configuration with N at
the surface: the bottom electrode M transmitted through the same con¬
ductor both S.P. and resistivity signals (the voltage corresponding to the
current emitted between A and B). For a few months the plotting took
place point by point; the signals passed through two potentiometers in
series, and at each station the S.P. and resistivity measurements were
made, ofie after another. This very primitive form of telemetering involv¬
ing a chronological system was soon replaced by the continuous and
simultaneous recording of both parameters through an appropriate con¬
nection in the circuit of the recorders and the pulsator. The principle of
this wiring remained unchanged for several decades, as long as the pul¬
sated system was used.

In accordance with the terms of the contract between Schlumberger


and the U.S.S.R. petroleum organizations, Doll arrived in Grozny in
February 1932, to conduct tests. The S.P. measurements had become
routine procedure by that time, and hundreds of diagrams had been
plotted. They were considered only as accurate qualitative indicators of
the limits of permeable layers, and interpretation of the magnitude of the
low values produced by these layers had not been the subject of any
study. Nevertheless, according to the theory of electrofiltration, the
electromotive force generating the S.P. had to be proportional to the
resistivity of the mud and to the pressure differential between the mud
and the fluids inside the layer, whether water, oil, or gas. A simple
calculation showed that, if the mud pressure varied by a known quantity,
everything else being equal, the formation pressure should, in principle,
be obtained by comparing the S.P. amplitudes.

16 This theorem, long known in the case of linear circuits, had been set forth by Wenner in
1912. Its demonstration is immediate for a homogeneous indefinite medium. In the 1930’s
it was mathematically demonstrated by Lienard for any heterogeneous medium, and
experimentally verified in surface work and in the Pechelbronn oil wells. The reciprocity of
the quadripole has been so frequently applied that today it is taken for granted.

121
The History of a Technique

One or two rather promising experiments to this effect had been


made at Pechelbronn, and part of Doll’s program was to pursue them in
the U.S.S.R. In the Grozny area the muds had low salinities and hence
rather high resistivities, and thick sandstones frequently gave S.P. values
80 to 100 millivolts in magnitude. Two series of measurements were
made within a span straddling such a sandstone. In the first series the mud
level was lowered to some 120 feet below the wellhead: in the second the
well was filled to the top, making for a pressure differential of several
atmospheres between the two recordings. The S.P. value of the sandstone
showed substantial variation, and the value computed from the mea¬
surements for the pressure appeared reasonable to local geologists.
The test was repeated a few weeks later in Baku. As the muds were
considerably more saline, the S.P., although providing a good definition
of the boundaries of sands and sandstones, was of a lower average value.
Contrary to expectations, when the mud level changed, the curves
showed no discernible difference, opposite formations with S.P. anom¬
alies of 20 to 30 millivolts, this with a pressure differential about the same
as in Grozny. Calculation immediately showed that to explain S.P. values
of such magnitude by the electrofiltration hypothesis amounted to assum¬
ing that the formation pressure was essentially nonexistent—a practical
impossibility. The question arose: if the pressure change effect on the S.P.
was perhaps not instantaneous, might it not have had time to occur within
the experiment? Indeed, plans were made to resume the test at a more
leisurely pace as soon as a borehole became available, but experience
already suggested that S.P. might have other causes than electrofiltration.
Thereupon Louis Bordât, who had a crew in Surakhany (one of the
best producing fields of Baku), reported to Doll the existence of a sand
horizon, saturated with highly saline water, where the S.P. value reached
80 millivolts and more, that is, above what had been observed anywhere
else. Doll assumed from Bordat’s observation that the salinity of the
formation waters could play a part in the S.P., that an electromotive force
of the electrochemical type could develop at the contact of these waters
with the less saline mud,17 and that the currents generated flowed and
closed through the borehole in the same manner and direction as the
electrofiltration currents. A few hasty experiments by Conrad in Paris
showed that such an assumption was not altogether unfounded. Thus in

17 The two waters constituted what physicists call a concentration cell.

122
The Origins and Development of Logging

April 1932 came the discovery that at the depth of formations containing
salt water the S.P. has an electrochemical component. It was established
somewhat later that in formations containing hydrocarbons water in capil¬
lary form occurs everywhere, even if production is practically water free,
and that here, too, the electrochemical S.P. can be observed.18

Current output sonde and normal and lateral sondes

The problem of the span of the electrode configuration (or sonde)


had arisen with the first Pechelbronn experiments. After a rather
haphazard beginning with AM = 3 meters and MN = 1 meter, the
spacings Underwent various adjustments in order to obtain diagrams as
sharp and differentiated as possible. It was known that the quantities
being measured were apparent resistivities19 whose values were affected
by the mud-filled borehole. In 1927 no calculation had yet been made to
define this point more accurately. However, on the assumption that the
medium in which the sonde is placed is homogeneous (as is the case when
the mud and formation resistivities are the same), the distribution of the
equipotentials had become standard knowledge. On this basis it was
assumed as a first approximation that the volume of ground comprised in
the measurement could be represented by a horizontal disk, with a
thickness equal to MN and a diameter equal to AM. However rough, this
interpretation was convenient, not unlike the AB/4 value long adopted in
surface prospecting to represent the depth of investigation.
The fact that the volume of ground was of the order of several cubic
meters offered an advantage over the situation prevailing when the infor¬
mation provided was confined to the diameter of the mechanical coring

18 For the record, Conrad had already envisaged the possibility of electrochemical reac¬
tions occurring between the various formations encountered in drilling. Reference to this
had been made in his communication of December 10, 1931, to the U.S.S.R. Academy of
Sciences, but he considered such reactions to be very weak, and negligible compared to
electrofiltration potentials.
19 When an electrode configuration AMN (with B remote) or its equivalent MAB (with N
remote) is placed in an indefinite homogeneous medium, the value measured is, because of
the calibration of the system, equal to the resistivity of that medium. However, what is
practically measured by electrical coring (as well as by surface prospecting) is an apparent
resistivity (see p. 38), in other words, an average value that is a function of the geometry
and the resistivities of the media within range of investigation of the system.

123
The History of a Technique

tool. However, to the extent that cores are recovered, geologists can
examine them almost inch by inch (in reality, they do this only in special
cases and usually limit themselves to one or two observations per foot).
To reach such a fine degree of resolution seemed very difficult with the
sondes in use; their size could not be unduly reduced lest the mud effect
become overwhelming and blur the peaks of the diagrams. To distinguish
between even the thinnest layers, Conrad turned to quite a different
measuring device.
In a very large medium the resistance of an electrode is defined as
that between the surface of the electrode and infinity. In fact, beyond a
certain distance (which is itself a function of the size and shape of the
electrode), the equipotential surfaces become very large, and the resis¬
tance practically zero. For an electrode, say 2 inches in length and diame¬
ter, 90 percent of the resistance will be confined within a 10 inch radius,
and 98 percent within a radius of 4.5 feet.
The left-hand side of Figure 37 is a cross section of a short electrode
(say 2 inches long) suspended at the lower end of an insulated conductor
at the level of a very thin horizontal layer (e.g., 8 inches), which is more
resistant than the adjacent formations, for example, a hard sandstone bed
between salty clays. The current lines flowing out of the electrode en¬
deavor to avoid the resistant formation. Therefore the resistance encoun¬
tered by the current is hardly affected by the hard bed; and if the sonde
moves in the borehole, the hard bed will show on the diagram, if at all,
only as a weak, flattened-out peak.
Let us now move to the right-hand side of the figure and assume that
electrode A is flanked by two elongated electrodes, A' and A", and that all
three are maintained at the same potential. All the current flowing out of
A is now forced to pass through the hard bed, and on the diagram the
resistance of A will reach a much higher value when passing in front of this
bed than it will for the adjacent formations; the result will be a well-
contrasted peak. “Guard electrodes” was the name Conrad gave to the
two added electrodes which emitted current for the purposes of prevent¬
ing that flowing out of A from dispersing, thus forcing it to squeeze within
a nearly horizontal sheet.
The same design, resumed years later, was realized through electronic
circuits. The wiring of Conrad was utterly simple. The two guard elec¬
trodes, A' and A", were short-circuited and connected by one of the
conductors of the tricable to a source of current at the surface. The casing

124
The Origins and Development of Logging

Figure 37. Distribution of the current emitted by an electrode. Left: Without guard electrodes.
Right: With guard electrodes.

constituted the return electrode. A and A' were connected by a shunt


whose resistance was sufficiently low for the guard and A electrodes to be at
about the same potential, yet high enough for the potential difference at the
terminals of the shunt (connected with the two other wires) to be measura¬
ble. At each station the potential difference was measured between A and a
surface electrode, as well as the difference between shunt terminals, from
which measurement the value of the current emitted by A was computed
(Fig. 38). The ratio between the two values was the resistance of electrode
A. Calibration of the system in a homogeneous medium of known resistiv¬
ity (a miniature sonde in a tank full of water) led to the conversion of the
resistance diagram into a diagram of apparent resistivities.

125
The History of a Technique

-►C

Figure 38. Facsimile of a sketch by Conrad


Schlumberger of the current output sonde (October
13, 1927).

The device was called a “current output sonde” because Conrad


wanted to measure the current emitted by the central electrode main¬
tained at a constant potential. The first test took place at Pechelbronn in
December 1927; although others followed, the system never found prac¬
tical application. The measurements took more time, and it was impossi¬
ble to reconcile the concern for detail with the depth of investigation, as
can be done with the AMN configuration. After a series of experiments
from which the Pechelbronn mission adopted electrode configuration
values of AM =4.5 feet and MN = 20 inches, the current output sonde
was put aside.
However, the idea of using guard electrodes was not altogether
abandoned. In a long handwritten note dated May 1, 1932, Conrad no
longer envisaged elongated guards but visualized electrodes which, as in
standard sondes, would be small enough in respect to their spacing to be
assimilated to points. In 1950 Doll reverted to the concept of a sonde
with guard electrodes restraining the central flow. He designed and per¬
fected point electrode systems that offered much greater flexibility than
the current output sonde and that made possible the simultaneous record¬
ing of the S.P. as well as other logging methods. Under the name

126
The Origins and Development of Logging

“Laterolog,” they have today become basic tools of electrical logging (see
p. 254).

By mid-1930 it had been amply demonstrated that electrical coring


was able to delineate bed boundaries and define correlations. On the
other hand, in Grozny (New Fields) and in Venezuela (Concepcion and
Cabimas) the sole resistivity curve had already made it possible to distin¬
guish between certain water- and oil-bearing formations, and even to
forecast productivity. The results were empiric, obtained as they were by
extrapolating the data resulting from the comparison between the produc¬
tion and the amplitude of the peaks on the diagrams. It was recognized
that only very favorable geological circumstances had provided these
results and that, in general, the correspondence between production and
apparent resistivity is not so simple. In 1930, however, not all the factors
involved in this relationship had been detected or analyzed. For the sake
of clarity, it seems appropriate at this point to anticipate the sequence of
events and to summarize the present degree of knowledge on this ques¬
tion.
Production is contingent on various factors, particularly the amount
of hydrocarbons contained in a bed. This amount can be calculated from
the thickness, the porosity, and the saturation of the horizon. Saturation
(note 10, p. 10) in turn is reflected in the value of the true resistivity of
the bed; true resistivity is also a function of several other formation
variables: its porosity, the resistivity of the water occupying part of the
pores, the amount of shale contained in the rock,20 and, to a lesser extent,
such other characteristics as grain size and shape.
Figure 39, which shows a cross section of a borehole drilled through a
permeable formation—sandstone, for example—explains the relationship
between the true and the apparent resistivities of a reservoir. Such a
formation is always more or less deeply invaded by the water contained in
the mud, which repels the fluids from their original place. On the drawing
the slightly undulating line marks the limit of the invasion. It is known
today that there is no such sharp limit and that the transition to the virgin

20 The shale may occur in the form of very fine beds (laminations) or of minute grains
disseminated in the pores of the reservoir. The shale content of the formation is often low
enough to have a negligible effect on the resistivity of the formation.

127
The History of a Technique

Borehole

128
The Origins and Development of Logging

zone is gradual, but both calculations and experience show that the
intermediate zone may be disregarded with little error. It is further known
that, when the mud deposits its clay particles along the wall of the hole, a
kind of crust (mud cake) is formed and that, beyond this crust of a fraction
of one inch, only clear water—the mud filtrate—penetrates the rock. The
apparent resistivity (R a) measured by the sonde depends not only on the
true resistivity (R t), but also on the borehole diameter (d), the resistivity
of the mud (.Rm), the diameter (D,), and the resistivity (RXo) of the
invaded zone. Furthermore, unless the formation is very thick compared
to the spacing of the sonde’s electrodes, Ra will be influenced by the
shoulders of the formation, and this will involve their resistivity (Rs) and
the thickness {h) of the formation.21
The above description provides an idea of the number of factors
standing between a formation’s production and its apparent resistivity.
The outstanding results obtained in the U.S.S.R. and Venezuela were due
to the fact that, saturation excepted, the multiple factors influencing
productions and true resistivities remained rather constant in each hori¬
zon and there was little difference between apparent and true resistivities
(rather thick beds and shallow invasions).
In most cases the logging tools used today give the values of porosity
and true resistivity. On the other hand, it is generally possible to know the
resistivity of the formation water and to assess the effect of the interstitial
shale. From these parameters experimental relations lead to the almost
exact value of the saturation, and from there to the volume of hy¬
drocarbons in place. In 1930, however, when electrical coring could give
only the formation’s thickness and apparent resistivity, hardly anything
was known of the relation between the latter and the true resistivity.
Moreover, although the porosity could be obtained from core analysis
(with serious limitations), the above-mentioned experimental relations
were still unknown. The history of logging is marked by the stages of
progress made toward gathering, under the most diverse geological condi¬
tions, the fullest data on the fluid contents of the formations and their
ability to produce.

21 Usually D i varies from an inch or so to as much as 6 feet. All the notation here is taken
from recent English terminology: m for mud, s for shale, t for true, i for invaded.
Laboratory and field experience shows that the apparent resistivities are practically the
same for a dipping layer as for a horizontal one.

129
The History of a Technique

In the effort to reach what is called the petroleum diagnosis of a


horizon, an initial idea in 1930 was to lengthen AM to about 30 feet, thus
increasing the range of investigation to the point of “seeing” beyond the
invaded zone. One or two tests of this kind were conducted in Indonesia,
but it quickly became apparent that mere extension of the sonde was no
answer to the problem. Invasion was not the sole factor producing the
discrepancy between true and apparent resistivities: the thickness of the
formations also played an important role. When the thickness of a resis¬
tant layer was near or below AM, nothing much showed on the diagrams,
whereas very thick layers could exhibit very high apparent resistivities.
There was also a practical problem requiring urgent solution. Anx¬
ious as they were to have information on the potentially productive hori¬
zons as soon as the bit penetrated them, the oil companies demanded that
measurements start as close as possible to the well bottom. Unfortunately,
because of the length of even the shortest devices {AM between 7.5 and 9
feet) and of the weight extension below the sonde, the lowest possible
measurement described mainly the formations some 15 feet above bot¬
tom, which meant that often the most significant part of the operation
might be missing on the diagrams. A solution to the problem was seen in
inverting the configuration, that is, placing MN below A (the result was
called the “MN-below” sonde). The proposal appeared to be sound, since
everything happens as if the surface electrode were infinitely remote; and
consequently, in an infinite homogeneous medium, the value measured
will be the same whether MN is above A or below. There was, however, a
suspicion that in a real medium composed of layers with various resis¬
tivities the diagrams obtained with the two positions would not be identi¬
cal because the AMN configuration was asymmetrical. This was confirmed
by an operation in the northern coal basin of France.
In February 1931 a coal mining company had asked for the electrical
coring of a reconnaissance borehole near Courrières. Drilling had re¬
vealed a rather thick coal seam, about 2000 feet deep, but extremely
fragmented cores had not permitted an accurate location of its limits. A
first diagram, recorded with a 7.5 foot AMN sonde and MN above A,
showed very sharply a peak approximately 4.5 feet thick. A second
record, obtained after inverting the device, revealed a peak of the same
width and height, but shifted by 7.5 feet toward the bottom in relation to
the first diagram. The prospector was intrigued by this difference; he

130
The Origins and Development of Logging

Figure 40. Two resistivity diagrams through a coal seam, proving that the AMN configuration
was asymmetrical.

suspected some mistake and repeated the recordings, taking the utmost
care with the depth measurements. The offset was confirmed (Fig. 40).
A thorough and systematic study followed in the wake of this ex¬
perimental illustration of the effect produced by the asymmetry of the
AMN system. Little light had been shed on the problem thus far by the
few calculations made on the basis of highly simplified assumptions, the
principal one being to disregard the borehole altogether. Doll improved
these calculations by an approximate mathematical treatment relating to
the case of drilling through infinitely resistant layers22—a more represen¬
tative model of the real configurations. Tank measurements, where lime¬
stone slabs simulated formations, expanded and confirmed the calculated
results. These studies were recorded in March 1932 in a house publica¬
tion; they marked a crucial stage in the development of the Schlumberger
techniques. They were the first to give a more accurate definition of the
curves recorded with the AMN sonde. When MN is below, a resistant

22 This simplification was based on the working assumption that oil- or gas-bearing
horizons had an infinite resistivity, whereas, according to the theories on petroleum
migration, a fraction of the pore volume was always occupied by more or less mineralized
water, retained by capillarity. Many diagrams from different regions rapidly proved that
this resistivity is always finite, that is, that the capillary water forms a continuous, electri¬
cally conductive system.

131
The History of a Technique

layer thicker than AM shows as a peak near the bottom, and a low-
resistivity zone underneath its top, approximately equal to AM. The top
of the layer is then said to be “eaten.” With MN above, the curve is
inverted. This explained the offset observed at Courrières; furthermore,
the true thickness of the coal seam had to be 12 feet (7.5 + 4.5), a value
that indeed corresponded with the information gathered from the cut¬
tings. Another conclusion was that it was hopeless to use the AIN-below
arrangement to study the formations opened up by the last 12 or 15 feet
of the well.
The design of a new electrode system, described by Conrad on May
1, 1932, was yet another result of these studies. In this system N was
placed at a considerable distance from AM. What was measured was, in
fact, the potential of M with respect to infinity (this was called the
“AM 00 sonde”).
In September of the same year the “1932 sonde” marked new prog¬
ress over the AMN device (Fig. 41). On the left-hand side of the figure the
current is emitted between A and B (24 feet apart), while the potential is
measured between M (1.5 feet below A) and the surface electrode. The
system is equivalent to the AM °o sonde, since the distance from B to A is
16 times greater than the distance AA4, and hence the effect of B on the
measurement is negligible. This system, where the measurement was
recorded at the center 0 of AM, was called the “normal sonde.” Its
advantage was that it required only a very weak current. The diagrams
recorded were symmetrical about the center of a layer, whose top and
bottom were accurately marked. The AA4 spacing being very short, mea¬
surements could be made very close to the bottom of the hole. On the
other hand, because the system had only a small radius of investigation,
the measurements were generally affected by the invasion. Finally, it
responded very weakly to the thin layers.
On the right-hand side of Figure 41, the current is emitted between
the closely spaced A and B electrodes, while the potential is measured
between the remote electrode, M, and the surface: this is now the MAB
configuration, which, according to the reciprocity theorem, is the equiva¬
lent of AMN. Called the “inverse sonde,” it was not unlike the one used
since the beginning of electrical coring. With AM 24 feet long, the radius
of investigation became substantial—hence the other name, “lateral
sonde.” This sonde was well suited for detecting thin layers, even though
it could not measure their true resistivity. Its shortcomings were that the

132
The Origins and Development of Logging

^/\s cjU>'
V I / I * '
>J3$ •..O «7 /
x. /.
V / f'r

N A>. Equivalent Equivalent f'.p


iff k >^/ /
0 iT
l~ setups
J MV,\
■s
fy J-J'-'d
O'H!
bvf
■>n setups ï
p5■ „-3
IS:
<i/N
lr ' \\
pi*
>1 N
&VM

m 4 &\v
P5
'V -/<:
"TO Si
§vi
\ aa M éb
y%.
A\f âp
fcXd <i'/i fc$l l\\s
k^l
y&</'
V/

Ï\U m 4. pi '>$ >k


if/-p •Ÿ-\ .-N
4
'r/$
\y,
f\J
<l "v
Vdf
b'
P'/
\'
.yV,
J'/'r
?\X
Î
c>| i* :\ w
i
• / /- ^4a
4'4
'PAS
b
fi
'"--a
'/
Pi
fe:,
Jk V'
r>b\
CO?
yvil
/' Lr / i-
y Hî *'\ N
i,\'^
;'4 15);
Ü'
£vy / /:
y è Measurement \_ ^n_t_

*M \ A £\ia' M p-
C'4l
>,'/$ A g
\"/*
■,v|M bp
point v|
' \V4 /' wVf* feb
Normal sonde Reverse or
lateral sonde

Figure 4L Normal sonde and inverse or lateral sonde.

top of a thick, resistant layer was deeply “eaten,” and that a very strong
current was required. With either type of sonde, normal or lateral, the M
electrode simultaneously recorded the S.P. curve. Moreover, when both
lower electrodes were eventually placed on a rigid cylindrical body dou¬
bling as a weight, the measurements could practically reach the bottom of
the hole.
The question was raised as to whether, with the normal sonde, one
conductor could not be made available by bringing B to the surface: this
would have been strictly an AM sonde. However, experience soon
showed the proposal to be impracticable with pulsated-alternating current
recording. As a current wire and a measure wire wound on the winch
amounted to an induction coil, the pulsing on and off of the current circuit
generated a dangerous overvoltage, and the electromotive force induced

133
The History of a Technique

in the measure circuit considerably altered the reading. The use of the
third wire for the current returning to electrode B near the bottom of the
hole reduced these effects to acceptable limits, although it was impossible
to prevent them altogether because of the asymmetry between the two
wires. However, the width of the dead segments on the measure collector
of the pulsator was such that at 6 revolutions per second the induced
millivolts had disappeared by the time the measure circuit was closed
(see p. 53).
From 1932 on, the normal sonde generally became the basic tool (the
U.S.S.R. excepted), whereas for several years the lateral sonde was used
only in specific cases. The dimensions of both sondes underwent various
adjustments to suit local geological conditions, and they remained in use
until they were replaced, at least in oil wells, by induction logging and the
Laterolog. They continue to be a part of the simplified and portable
equipment employed in hydrological surveys.

Other measurements in boreholes

Together with the developments described above, technical studies


were oriented as early as 1930 toward other projects, the principal ones
being a device to measure the dip of the strata, an electrical resistance
thermometer, and a technique for locating water inflows in boreholes.
Inspired by Conrad, these various projects were essentially realized by
Doll. Simultaneously, Marcel was tackling the problems of sidewall for¬
mation sampling (lateral coring) and casing perforation (see p. 182 ff). He
had already considered pressure measurements at the bottom of
boreholes, a problem that was to be taken up again 25 years later.
When a first borehole is located on the basis of geological and
geophysical surveys, it is usually very difficult to foresee what, at the
presumed depth of the pay zones, its precise position will be with respect
to the overall configuration of the field under exploration. In the case of
an anticline, for example, little is generally known as to whether the
borehole will reach the interesting formations close to, or far away from,
the axis, or even on which flank this will occur. There are several reasons
for this uncertainty: geophysical interpretations allow a certain leeway;
various features—faults, transgressions, unconformities—may result in
a deep structure being quite different from its representation on the basis

134
The Origins and Development of Logging

of surface geological observations; since no borehole is ever exactly


vertical, the difference between bottom and wellhead coordinates may
amount to as much as 50 feet or more.
The azimuth and the angle of dip of the layers encountered in
drilling, and the drift of the hole itself, contribute substantially to defining
the configuration of the deep objectives and guiding their exploration. To
be specific, when a first well reaches an oil-bearing horizon, it is important
to know whether a second well will intersect the same horizon at a point
nearer the axis of the structure or more remote from it; in other words,
whether, as geologists say, the second well will be located updip or
downdip in relation to the first one. This is the reason why geologists have
always tried to determine the magnitude of the dip from examination of
the cores, especially when the cores come from finely layered formations
like alternating shales and sandstones. Yet all the method yields is an
angle with a plane perpendicular to the axis of the hole for the depth of
each core. The values observed are merely apparent dips, and the differ¬
ence between them and the true dip can be important if the hole is
sufficiently slanted. Moreover, little is known of the orientation of the
core, and hence the azimuth of the dip.
It was therefore important for the petroleum industry to find ways of
determining the dips. A proposal to use electrical measurements for this
purpose was made by Conrad at the time of the first Pechelbronn experi¬
ments in October 1927. The principle was nothing new: as in surface
prospecting, it consisted of utilizing the electrical anisotropy of the
sedimentary formations. All the rest, then, was mere technology—but
how complex! Let us recall that, if a current is emitted by an electrode
inside a borehole at the depth of an anisotropic, inclined, sufficiently
homogeneous and thick layer, the equipotential surfaces are not spheres,
but are flattened revolution ellipsoids whose axes are perpendicular to the
plane of the strata (Fig. 42).23 For the sake of simplification the figure has
been drawn in the vertical plane passing through the steepest gradient of
the strata. Placed symmetrically about the axis of the drillhole and at a
certain distance (say 3 feet) above the power electrode, the two electrodes
R and Q are on two equipotentials, and the potential difference between^
and Q is a function of the slope of the ellipses at the point where they are

23 These ellipsoids are actually slightly deformed at their intersection with the borehole, but
this is only a secondary and generally negligible effect.

135
Battery

Figure 42. Equipotential curves surrounding an electrode in a homogeneous and anisotropic


dipping formation.

136
The Origins and Development of Logging

intersected by the borehole. From the measurement of this difference, it


is possible to determine the angle of slope of the ellipses, this angle being
less than the true angle of dip. In practice the cable, when lowered into
the borehole, has a tendency to untwist under the increasing stress of its
weight; the sonde revolves several times around its axis, so that the orienta¬
tion of electrodes R and Q varies with the depth. For this reason the
system comprised two identical and perpendicular pairs of electrodes; by
composing the values measured on each pair, a vector was obtained giving
the azimuth of the dip with respect to these electrodes, that is, to the
orientation of the sonde around its axis. This orientation then had to be
determined with respect to the north. Furthermore, the determination of
the true dip angle required both the value of the apparent dip as deter¬
mined from the cores, and the angle of slope and the azimuth of the
sonde’s axis.24 From there on, a rather simple spherical trigonometrical
computation gave the true dip. This led to the design of a sonde in two
parts. The first one, initially called the “inclinometer” and later the “tele-
clinometer,” indicated at each measurement level the inclination of the
sonde and its orientation with regard to its axis; the second one, termed
the “dipmeter extension,” gave the azimuth and angle of the dip with
regard to the orientation of the sonde.
To design the circuitry, Doll had to manage with the tricable as the
only electrical and mechanical connection between surface and drillhole.
It was out of the question to manufacture a cable with a greater number of
conductors since this would have required major and exceedingly costly
changes in equipment design, especially for the winches. As one conduc¬
tor of the tricable was mobilized for the emission of the current actuating
the system, only two remained available for the return of the four signals
of the inclinometer and the two of the dipmeter, each of the order of a
few millivolts. The borehole apparatus required a distribution relay con¬
trolled from the surface. For purely technical reasons the project was
achieved in two stages: the inclinometer first, and the dipmeter extension
some years later.
The inclinometer was the first Schlumberger instrument lowered into
a borehole that was no longer limited to three sections of lead wire
serving as electrodes, but involved a delicate assembly of electromechani-

24 Centering tools on the sonde ensured the actual coincidence of its axis with that of the
borehole.

137
The History of a Technique

cal components enclosed in a cylindrical housing only a few inches in


diameter that protected the instruments from being contaminated and
crushed by the mud. In spite of their small size and of temperatures
reaching 120 to 180 degrees F, the components were expected to work
without failure. The watertight bronze housing contained an induction
compass25 made of a billiard ball driven by an electric motor around a
diameter coinciding with the axis of the sonde. A coil of insulated wire
was wound around a groove machined into the plane of this diameter and
was connected with a collector contacted by two cross-shaped pairs of
brushes. The same current fed the motor and excited an electromagnet
maintained by a universal joint in a vertical position above the compass
(Fig. 43). The two electromotive forces simultaneously induced in the
compass by the terrestrial magnetic field and the electromagnet were
rectified through the two diametrically opposed dead segments in the
collector ring. To measure them separately, the current was reversed;
since at the instant of reversal the direction of the motor’s rotation did not
change, the electromotive force induced by the terrestrial field did not
vary, whereas that induced by the electromagnet changed sign. The
former defined the orientation of the sonde with respect to the magnetic
north; the latter served to compute the angle and the azimuth of its
inclination.
The whole was dipped into an insulating liquid—a mixture of oil and
kerosene—filling the housing. A metallic bellows at the bottom of the
sonde transmitted the mud pressure to the liquid and prevented the thin
bronze housing from collapsing. The electrical connection between the
inside wiring of the sonde and the three conductors of the cable was
achieved by elements designed to ensure, despite the pull of the cable, a
positive seal against mud intrusion and good electrical insulation. Such
devices, reminiscent of those used in motor engineering, were called
spark plugs. The distribution relay included a ratchet cleverly driven by
the rotation of the billiard ball, which, drilled through its center and
working like a centrifugal pump, created in the oil an overpressure that
pushed a piston driving, in turn, the ratchet, with the result that the latter
moved by one notch at each switching on or off of the current. In this

25 An instrument designed to define the direction of the magnetic field of the earth by
utilizing the electromotive force induced by this field in a coil revolving around a vertical
axis.

138
The Origins and Development of Logging

Figure 43■ Part of the mechanism inside the inclinometer. Top: Electromagnet suspended on its
dial. Center: Hollow billiard ball used as an induction compass. Bottom: Piston controlling the
ratchet of the relay.

manner, at each station—every 30 or 50 feet—the four average values


of the voltages between the brushes of the compass were obtained, from
which the inclinometry data could be computed. Two additional positions
had been provided for the relay to serve for measurements between the
electrodes of the dipmeter extension. The results of the computation
were interpreted through charts, especially drawn for each region (the

139
The History of a Technique

local value of the terrestrial field intervened in the calibration), and it was
possible to make the measurements and draw the profile of the drillhole
for the entire open-hole section (Fig. 44).
The inclinometer (or teleclinometer), made available to industry in
1932, answered an important need. Because all oil wells drift from the
vertical, it is important to know accurately their underground positions at
reservoir and key geological marker levels. On the other hand, slanted
holes gave rise to many drilling mishaps (e.g., stuck drill pipe). Methods
had already been devised to measure the drift of drillholes,26 but none
had industrial application. The various devices utilized in Baku were
inaccurate and inconvenient, and none (including the Shakhnazaroff de¬
vice) gave the azimuth of the inclination. This explains the enthusiasm
with which the teleclinometer was welcomed by the oil industry; its
success and widespread utilization were almost immediate.
It took longer to achieve the complete instrument: the teleclinome¬
ter and dipmeter extension. After many tests at Pechelbronn, it actually
reached the market in 1935.

Another important problem for oil operators was the location of


water intrusions in boreholes. This had been studied from the very
earliest days of electrical coring. It often happened that petroleum was
accompanied not merely by traces of water driven out of the pay zones,
but also by unacceptable discharges originating from active aquifers lo¬
cated either within the producing zone or above it behind the casing (see
Fig. 28) and flowing through faulty cement. Remedial action required
accurate knowledge of the depth of the water intrusion. The method in
use began by injecting into the borehole a mud dense enough to prevent
the intrusion of any fluid—in oil field terms, to “kill’ the well. The mud
was then circulated to ensure homogeneity throughout the zone of study,
after which the mud level was gradually lowered to the point where the

26 Particular mention should be made of a U.S. patent granted in 1915 to H. M. Smitt, in


which an instrument is described whereby a photograph is taken of a circular bubble gauge,
maintained by a gyroscope in a fixed and known azimuth. The principle of the design was to
be developed and marketed much later by Sperry-Sun in the United States and O.
Martiensen in Germany. These instruments were less convenient than the Schlumberger
teleclinometer, but since they were not based on the terrestrial magnetic field, they worked
inside the casing.

140
The Origins and Development of Logging

Figure 44■ Doll calibrating a teleclinometer in Baku j(1932).

water could begin to flow. The intrusion level was located through an
appropriate device lowered at the end of a cable.
The Schlumberger process was based on measuring the resistivity of
the liquid filling the hole from top to bottom. A special sonde was
designed, including three small lead electrodes about one inch apart, and a
cylindrical shell that kept them sufficiently remote from the borehole wall
(whether casing or open hole) so that the measurement would not be
affected. The system was calibrated above ground in a tank filled with
water of a known resistivity. In general, this resistivimeter was rather easy
to handle in the borehole; all that was needed was to circulate ordinary
mud, which normally is much less salty than the water to be detected.
Over the years the determination of water flows with the resistivime¬
ter had become routine, but the method was used less and less frequently

141
The History of a Technique

as the techniques of deep exploration and production were perfected. In


particular, the completion of wells by perforation (see p. 188 ff) provided
a much more effective way of sealing off the pay zones against water
intrusion.

The increase of temperature with depth—a phenomenon charac¬


terized by the geothermal gradient—was also utilized to locate water
intrusions. When the mud had been circulated in a borehole and the
temperature then measured from top to bottom, the measured gradient is
much less than the true gradient. The reason is that below 50 or 100 feet
the mud is colder than the underground. Since formation water flowing
into the borehole is warmer than the mud at the same depth, a new
temperature diagram recorded at a later time will show an anomaly
locating the water intrusions.
The idea of utilizing temperature measurements led to the study of a
borehole thermometer that could be adapted to the electrical coring
equipment. The only instruments employed in the industry around 1930
were maximum-minimum thermometers that required delicate handling
and, moreover, produced only one measurement for each run at the cost
of an hour to reach temperature equilibrium. Around 1928 an electrical
thermometer had been used in Virginia, but in addition to being very
expensive it did not allow for continuous recording. The principle of the
Schlumberger thermometer was based on a Wheatstone bridge, one leg of
which was a ferronickel resistor very sensitive to temperature; the other
three were of invar alloy. The instrument not only had to be both sturdy
and highly sensitive (to a fraction of a degree), but also had to have a low
time constant and good electrical insulation. Several models were tested
in succession with dubious results. Finally, in 1933 the required perfor¬
mance was achieved by a model in which the ferronickel wire, insulated
by a thin layer of varnish, was wound in a helicoid groove machined in a
steel core; the whole was encased in a copper sleeve forced under pres¬
sure against the wire. There was a close contact between copper and wire,
without any damage to the insulation, and hence a very rapid temperature
response.
The thermometer had various applications. It was used not only for
the location of water intrusions but also for the detection of gas or oil
inflows, the study of fluids circulating between two permeable layers with
different pressures behind a defectively cemented casing, and the control

142
The Origins and Development of Logging

of cementing (see p. 179). Thought had also been given at some time to
utilizing the thermometric diagram for the differentiation of formations
according to their thermal conductivity, but this parameter proved much
less sensitive than electrical conductivity and hence without practical
interest. Today the drillhole thermometer, greatly improved over the
model of the 1930’s, is a standard tool of production logging.

Field operations

In a letter to Léonardon dated May 31, 1927, Conrad and Marcel had
suggested that a drilling company, preferably an American one, might be
associated with the industrial development of electrical coring. The pro¬
posal remained vague and was promptly abandoned. The initial operations
at Pechelbronn seemed too promising to warrant association with outsid¬
ers. By establishing correlations among boreholes over several miles,
electrical coring reduced and often replaced mechanical coring, thus
effecting economies in oil exploration. These results, together with im¬
proved equipment, led the Pechelbronn company to enter into a service
contract with “La Pros.” The contract, signed on July 12, 1928, marked
the date that the Schlumberger process became commercial. The contract
provided that in return for a monthly fee of 12,000 francs a crew was to be
made permanently available to Pechelbronn; it further provided Schlum¬
berger with the right of access to boreholes for the purpose of continued
testing and development of new techniques. The Pechelbronn oil wells
thus became a kind of extension of the Paris engineering office. This
cooperation lasted until World War II.
However encouraging this first commercial success, nobody could
foresee in 1927-1928 that what was involved was an invention with a
commercial future far beyond that of surface electrical prospecting. At
that time electrical coring was regarded as just another process, applicable
to the location of key geological horizons, ore bodies, coal seams, and
water-bearing formations, as well as to petroleum exploration; in other
words, as an additional volume of business. Indeed, the operating condi¬
tions at Pechelbronn—percussion drilling, very short open-hole sections,
and mainly thin, low-pressure, oil-bearing strata shown on the diagrams as
blunted peaks—were too special to allow any foresight of what the
method could yield in large oil fields outside France.

143
The History of a Technique

Electrical coring was sometimes used to determine the location and


thickness of coal seams more accurately than was possible with cuttings or
cores. The first operation of this kind was conducted by Sauvage in
August 1928 in a borehole of the Lorraine coal basin at Saint-Avold.
More followed; Courrières in February 1931 was particularly significant
because that diagram became the source of very important studies. Other
surveys to assist in the interpretation of surface measurements took place
in exploration boreholes for iron ore in Normandy. Still, such operations
were few, and it soon became evident that efforts had to be concentrated
on the oil industry, where the number of boreholes exceeds by far that of
all other drilling activities combined. Proposals were submitted to Royal
Dutch-Shell; the use this company had already made of surface electrical
prospecting had given it an appreciation of Schlumberger standards. The
merits of electrical coring were stressed by Marcel in a letter of October
8, 1928, emphasizing in particular the low cost of the process as compared
to mechanical coring. Royal Dutch was interested and sent Dr. Mekel,
director of its geophysical department, to observe an operation in the
field. The utmost care was taken in preparing for his visit. A well was
selected where the top of the Hydrobiae marls was known to stand out
sharply on the diagram. Deschâtre and Sauvage tidied everything up; the
access track to the well was sanded, and the wheels of the old pickup truck
received a new coat of paint. Mekel arrived in the morning with Conrad.
He was shown the equipment, and the operation was explained to him.
Sauvage made point-by-point measurements, immediately followed by
the plotting of the diagram on a strip of paper. Mekel wished to repeat the
measurements by himself. The sonde was lowered again; he found exactly
the same values, confirming the authenticity of Sauvage’s diagram. This
apparently conclusive visit resulted in the signing of a contract on
November 8, according to which Royal Dutch was to test electrical coring
in its fields in Venezuela and the Dutch Indies.
A few months later the merits of electrical coring were obvious
enough to justify its public introduction to the petroleum industry. The
occasion was the Second International Drilling Congress in Paris in Sep¬
tember 1929; the paper, written in English, was captioned The Electrical
Coring (Fig. 45). It set forth the principles of resistivity measurements
with several examples of correlations provided by Pechelbronn and one
by Venezuela. This was the first of a long series of papers to be written on
logging.

144
THE ELECTRICAL CORING

PAPER SUBMITTED BY

C. et M. SCHLUMBERGER

AT THE SECOND INTERNATIONAL DRILLING CONGRESS

Paris (September 1929)

SOCIETE DE PROSPECTION ELECTRIQUE

( PROCEDES SCHLUMBERGER )

30, RUE FABERT - PARIS - Vile

AND

SCHLUMBERGER ELECTRICAL PROSPECTING METHODS

25, BROADWAY - NEW-YORK CITY

Figure 45. Facsimile of the cover page of the first article on electrical coring (1929).

145
The History of a Technique

The Royal Dutch contract for Venezuela provided for an electrical


coring crew of two engineers to be made available for a monthly sum of
$2500. It was agreed that the crew, when available, could be assigned to
surface prospecting. However, between 1929 and 1931 there were only a
few short surface surveys of minor local problems.
Responding to a telegram from The Hague, Gilbert Deschâtre and
Pierre Bayle presented themselves in February 1929 at the Shell office in
Maracaibo, unaware of the nature of the problems they would be asked to
handle. Here was a new situation: the problem was no longer, as in surface
operations, to conduct measurements in an exploration area where at least
some preliminary plan had been drawn up from the outset, and where the
prospectors were granted some leeway in their work by the client. The
local geologists, somewhat uninformed about electrical coring, told Des-
châtre and Bayle that their primary concern would be not with explora¬
tion drilling but with day-to-day production, and especially with the
accurate definition of the depth to which the casing was to be lowered for
cementing and water shutoff (see p. 99).
At that time Venezuelan production, almost in its entirety, was
provided by the western part of the country: Lake Maracaibo and its
surroundings, Ambrosia, La Rosa (Cabimas), Lagunillas, Mene Grande,
Concepcion, La Paz. The production of these fields seemed adequate for a
shrinking world market, and the three main operators—affiliates of
Shell, Gulf, and Standard of Indiana—were doing little exploration. In
eastern Venezuela, 20 years of exploration had been unsuccessful except
for the newly discovered Quiriquire field.
Oil and gas occurred in sand or sandstone horizons separated by
generally rather soft shales. Tectonics were often complicated, for exam¬
ple, anticlines fragmented by many faults. In spite of the many cores
required by the geologists, there were frequent errors in forecasting the
depths of the pay zones and of the water shutoff. The prospectors had no
doubt that electrical coring would be of substantial help in solving these
problems. One asset was the great number of wells being drilled—
especially in La Rosa, where electrical coring was to begin—and hence
the possibility of rapidly establishing correlations from well to well to
demonstrate the validity of the method. The local geologists were rather
reticent. However receptive to new techniques, these seasoned profes¬
sionals who thought they knew everything about their fields could only
regard skeptically the youngsters who, under an agreement concluded

146
The Origins and Development of Logging

between remote head offices, were supposed to help them with paper
strips covered with diagrams. After all, what kind of experience did they
have outside of some negligible French oil field?
The prospectors were required not only to conduct the mea¬
surements but also to retrace the diagrams by hand and interpret them,
that is, to identify the electrical markers, to establish correlations from
which to deduce the deep configurations (especially the existence and
throw of faults), and to forecast the depth of the pay zones. It was not that
the geologists were uncooperative: their experience and data were avail¬
able; but, at least in the beginning, they left to the Schlumberger staff the
full task and responsibility of practical diagram interpretation.
The first operation took place on March 6, 1929, in the La Rosa field
(Fig. 46)‘. Dozens of diagrams followed, and after 10 months the number
of Crews had grown from one to three. For the Shell field staff, Schlum¬
berger had become a regular auxiliary with a recognized contribution. In
1930 operations were extended to the oil wells of Gulf, Standard of
Indiana, and Richmond.
Addressing a meeting of geologists of the Maracaibo Basin, Bayle
reviewed the results of 18 months of electrical coring. In certain areas,
multiple markers on the diagrams provided clear correlations over tens of
miles, eliminating almost completely the need for mechanical coring; in
other, tectonically more complex areas, the data gathered from the dia¬
grams complemented, refined, and corrected the geological data. The
results were less outstanding with respect to oil-bearing horizons. They
showed beautiful peaks whose heights, in certain zones, provided a clue
for production; but in general it was difficult to distinguish between
oil-bearing and water-bearing strata, or simply between permeable and
impermeable layers. It will be recalled that at the time electrical coring
involved only a single curve recorded with a 9 foot lateral sonde.
After their initial reticence toward electrical coring, many geologists
and operators became inclined to expect more from it than it could give,
such as the differentiation of aquifers, the determination of the water
contents in oil-bearing horizons, and the detection of gas occurrences.
These problems would not be solved until much later in most oil fields.
In any case the results were spectacular: 800 electrical coring opera¬
tions in 300 wells. These totals had not been achieved without the
strenuous efforts of the crews. Monthly performance reached 50 opera¬
tions under conditions of extreme physical hardship: difficult transport;

147
RESISTIVITES OHMS M2/m
O 5 10 15 20 25 30 35 40 45 50 55 60 65 70

Figure 46. First electrical coring diagram outside


France (La Rosa, Venezuela, March 1929).

148
The Origins and Development of Logging

debilitating climate; round-the-clock, short-notice calls; long, uninter¬


rupted night operations on the rigs—all of which, by the way, continue to
be the lot of well-logging operators in many parts of the world. What has
changed a great deal is the ease of communications and the quality of
equipment. In 1930 there was no transatlantic airline to South America.
In Venezuela the oil companies already flew their own aircraft, but for
prospectors travel generally meant almost impassable trails or slow, un¬
comfortable, and crowded lake boats. Often a trek of several days, inter¬
rupted by all sorts of obstacles—mud holes, sand dunes, fords, rickety
bridges—was required to reach a few wells located in the heart of the
jungle. Also, communications with Paris were very slow. For reasons of
economy, cables were used in emergencies only. In 1929 mail took 1 full
month to reach France; this time was reduced to 2 to 3 weeks with the
advent of airmail between Venezuela and the United States. It took
almost 2 months after customs clearance for equipment, shipped mostly
from Paris, to reach the field. Of course, this gave the prospectors a great
measure of autonomy.
In addition to everything else, the equipment was a constant source
of headache. The original set had been adapted from the Pechelbronn
experience, whereas the Venezuelan oil wells were much deeper—3000
to 4500 feet—and drilled with thicker and more viscous mud. Heavier
and heavier weights were needed before the drillers could be convinced
to circulate a lighter mud, thus making the walls of the hole as smooth and
“clean” as possible. In cases where an obstacle blocked the descent of the
tool, an expedient, often used elsewhere, was to lower the drill pipe
without the bit to a point just below the obstruction, and then run the
sonde through the pipe to the bottom of the hole. Cables wore out
quickly; since replacements from Paris were slow to arrive, it happened
more than once that sections of old cables had to be pieced together. The
power of the electric motor was so inadequate that the prospectors
preferred to drive the winch with a wheel of the truck. In addition, several
months passed before Shell complied with the contract to provide road¬
worthy trucks and helpers who were of some use. As to the prospectors’
board and lodging, these necessities were provided by the oil companies
in their more or less comfortable field camps. It was a life without privacy,
and not always in the choicest environment.
In 1931 all this activity collapsed. World economic depression led
the companies first to curtail and then to almost completely cease drilling

149
The History of a Technic/ue

operations. By the end of the year, electrical coring operations were


reduced to one or two a month. Only one Schlumberger engineer, Jean
Mathieu, remained on the spot. By mid-1932, however, things took a turn
for the better.

Operations in Trinidad began in July 1932, after 3 years of negotia¬


tions with British Controlled Oilfields, one of the English companies
operating on the island. The company had experimented with electrical
coring in the state of Falcon, Venezuela.
Trinidad’s oil reservoirs were lenticular: very thick sand formations
often enclosed in impervious shales. The lenses were many but small, and
the fields were so fragmented that production required a considerable
number of wells and intensive mechanical coring. Furthermore, excep¬
tionally high formation pressures necessitated very heavy muds (density
near 2). Consequently, production costs were very high, and in the middle
of the world depression in 1931 the Trinidadian companies questioned
the usefulness of pursuing their operations. It was under these circum¬
stances that they turned to electrical coring.
The first diagrams, already including resistivity and S.P. curves, were
a success; they outlined accurately the permeable zones and differentiated
clearly between oil- and water-bearing horizons. Within a few weeks the
use of electrical coring led to a substantial reduction in production costs,
particularly through the elimination of mechanical coring. This, together
with the general economic recovery, gave the Trinidad petroleum indus¬
try the opportunity not only to survive but also to expand.

In the United States the first tests were conducted on August 27,
1929, in Shell's B well between Hanford and Kettleman Hills, California.
With a hand-driven winch, a 1 ton cable, and no recorder (in short, the
equipment of the earliest Pechelbronn tests), Deschâtre and Roche suc¬
cessfully made (with the loss of a weight) a series of measurements from
2700 feet to the casing shoe at 54 feet (Fig. 47). The ensuing operations
were slow and painful, and on September 19, one of them, at the 6000
foot deep Ruthenfort wildcat, proved a disaster: pressure and tempera¬
ture made the cable so leaky that no dependable measurement was
possible. All these operations pointed to the inadequacy of the equip¬
ment; and with wells spaced so far apart that correlations were impossible,

150
The Origins and Development of Logging

Figure 47. First electrical coring diagram in the


United States.

the local geologists, regarding the method as useless, terminated the tests
after a few weeks (Fig. 48).
A contract with Gipsy, an affiliate of Gulf Oil, took the operations to
Oklahoma. The wells were less deep (2500 to 3600 feet); the formations
were harder, and lowering the sonde into the hole was easier. Operations
ran rather smoothly, though at the cost of endless troubleshooting and
repairs. The 23 jobs carried out at Seminole between October 15 and
November 15 (13,600 point-by-point measurements) kept the two
engineers at a daily work average of 15 hours; moreover, 8000 feet of

151
The History of a Technique

Figure 48. One among the earliest electrical logging jobs in the United States (California,
1929).

ruptured cable had to be abandoned in the wells. At this point the crew
received a hand recorder and a winch with electric drive.
Operations spread to Kansas, where Gipsy had undertaken the drill¬
ing of a great number of 750 foot deep structural coreholes. In both areas
electrical coring revealed several markers from which accurate correla¬
tions could be drawn. This achievement afforded the geologists little
satisfaction, however, for either the correlations affected series of no
special interest or electrical markers could have been just as easily located
by examining the cuttings,27 at less cost and without any interruption in
drilling. Finally, electrical coring failed to differentiate adequately the oil-
and water-bearing strata and the hard beds. For these reasons and because
of the fact that Schlumberger could not accept the 50 percent reduction in

2' Even in deep exploration the routine practice at Gipsy was to rely on examination of the
cuttings rather than on mechanical coring.

152
The Origins and Development of Logging

its already modest rates demanded by the client, the contract was canceled
at the end of the year.
Nevertheless, interest in the Schlumberger methods continued. In
January 1930 the Hugush Group (Humble, Gulf, and Shell) signed a 6
month, lump-sum contract which specified that the crew was at the
complete disposal of the client for work extending from the Mississippi
River to the Mexican border. In over 5 months the prospectors and their
overloaded truck covered some 8000 miles of sandy or muddy country
roads and trails (at that time paved roads ended some 20 miles beyond the
towns) to carry out 53 jobs on 45 wells scattered over 22 fields. The
drilling sites were often surrounded by a quagmire that required incredi¬
ble effort to cross. Again, the wells were so widely spaced that correla¬
tions we're practically impossible. Furthermore, the dearth of adequate
production data made it difficult to establish empirical criteria, and the
petroleum diagnosis was uncertain. Also, the many breakdowns and the
frequent difficulty in lowering the sonde to the bottom of the hole would
not have created enthusiasm in any client. With the depression as an
additional consideration, Hugush decided not to renew the contract.
One crew remained in the area for a few more weeks to carry out
scattered operations in small exploration drillholes. The United States
offered a market of immense potential, but with surface prospecting at a
standstill, and after the disappointments of the first logging tests, Schlum¬
berger could hardly support a money-losing crew with its inadequate
equipment. Temporarily, at least, it seemed wiser to concentrate on
technical improvements, while taking advantage of the contract with the
U.S.S.R., where electrical coring was in full swing. Business in the United
States was postponed to better times.
However, even the failures were not altogether fruitless. At least a
perception remained of the intrinsic merits of electrical coring. After the
first shock waves of the depression had subsided, Schlumberger returned,
offering greatly expanded and improved logging services (better cables,
simultaneous recording of S.P. and resistivity, power-driven winches,
etc.) in California (July 1932) and on the Gulf Coast (early 1933). This
time it took only a few months of demonstrated know-how and intensive
activity for the prospectors to establish logging firmly in the United States
petroleum industry. Contributing to this result was an exhaustive com¬
munication presented to the February 1932 meeting of the American

153
The History of a Technique

Institute of Mining Engineers,28 as well as the outstanding results ob¬


tained in the U.S.S.R., Venezuela, and other countries.

As had been the case in Venezuela and the United States, electrical
coring at Grozny was made difficult at first by equipment ill adapted to
depths exceeding 4000 feet. Moreover, the drillers, concerned over pos¬
sible cave-ins or uncontrolled blowouts, hesitated to add clear water to
the mud in order to facilitate the lowering of the sonde. Adequate cables
and winches had been shipped from Paris, but slow customs clearance and
rail transport delayed their arrival. Sauvage, the leader of the first
Schlumberger crew, skirted the problem, as Bayle had done in Venezuela,
by lowering the cable inside the drill pipe.
If a slow and not very efficient administration in the U.S.S.R. could
create difficulties, it had, on the other hand, a great advantage over
countries where operations are conducted by various companies, in that
here there was a single owner and the diagrams could be compared freely
across any geological structure. This circumstance, coupled with propi¬
tious sedimentary features, resulted in the production of many accurate
correlations by electrical coring within a period of several months. Also,
the world depression had left the U.S.S.R. untouched, and the First
Five-Year Plan provided for a rapid growth of petroleum production; to
that end, anything was welcomed that could increase the efficiency of
drilling and production. In no time the Grozny geologists made a basic
tool of electrical coring. While relying greatly on the guidance of their
French colleagues, they took interpretation into their own hands and
made the most of it. The results were especially good in the New Fields,
where large producing horizons showed peaks of up to several hundred
ohm-meters, while the surrounding shales did not exceed 4 or 5 ohm-
meters. Such sharpness of contrast was explained by the fact that in these
areas the formations had a very high oil saturation, whereas the remaining
interstitial water was of low salinity. It was recognized early by the
geologists that for some reservoirs the resistivity value reflected the order
of magnitude of the production rate; it could be anticipated from the
height of a peak whether the horizon would be gushing, and even whether
the daily production would be, say, 500 tons of oil without water, 10 tons

28 C. and M. Schlumberger and E. G. Léonardon, Electrical Coring, a Method of Determining


Bottom-Hole Data by Electrical Measurements, A.I.M.E., New York meeting, February 1932.

154
The Origins and Development of Logging

of oil with water, or water only. Thereafter it became easier to convince


the drillers that electrical coring was no waste of time, not only because it
avoided the complications of mechanical coring, but also because it added
safety and speed to accuracy in exploration and production. By October 1,
1930, 1 year after its arrival in Grozny, the Schlumberger crew had a
record of 240 operations in 101 wells, of which 6 were wildcats, with
measurements covering 105,000 feet of drillhole.

For the first operation on behalf of the Azneft Trust (Baku, October
1930), the same primitive equipment was at the site: hand-driven winch
and nonrecording potentiometer. One wonders whether the Azneft
people, though well aware of the value of electrical coring, wanted to test
it under conditions that looked somewhat like a trap. The well designated
for the operation belonged to the Surakhany field, one of Baku’s best
producers; but no one had told the prospectors that it was located on the
flank of the structure, where the strata were known to yield nothing but
salt water. Being accustomed to the Grozny diagrams with their huge
peaks, the prospectors were rather upset when, foot by foot, they read
very low resistivities varying between 0.5 and 1.5 ohms. This seemed to
be a bad omen for the future of electrical coring in Baku, but for the
Azneft people, on the contrary, it was definite proof that Schlumberger
meant business. The well selected for the next test was located near the
top of the structure, and, as expected, peaks appeared at the depth of the
oil-bearing strata.
In fact, Baku provided conditions even more suitable for electrical
coring than Grozny. The more than 500 wells drilled annually were less
deep, and lowering the sonde proved less laborious. Nearly the entire
production of some 20 sands interbedded with shales came from four
main fields: Surakhany, Lenin district, Bibi-Eibat, and Kara-Chukur.
These pay zones had variable thicknesses and were disturbed by multiple
faults. Clear geological markers were few (all the beds from top to bottom
were more or less the same, and there were essentially no fossils). These
circumstances explain why, especially after the introduction of the S.P.
curve, electrical coring could very rapidly make an invaluable contribu¬
tion. In the spring of 1931 four crews were at work with a daily average of
three to five operations, sometimes more; the strain of overwork was all
the greater in that, for lack of recorders, the plotting was still done point
by point, and only a few power-driven winches were available. In 1932

155
The History of a Technique

there were six crews with three recorders. The number of jobs exceeded
1200 in 1931 and 1800 in 1932.
During the same years the Surakhany and Bibi-Eibat fields were
comprehensively resurveyed using electrical coring. Accurate and detailed
structural maps defined the locations and throws of the faults and de¬
lineated the oil-bearing horizons, including some that until then had
remained unknown. A close relationship was observed between the resis¬
tivity and the contents of a horizon: as the formation waters were highly
mineralized, a fraction of an ohm-meter indicated an altogether aquifer¬
ous sand, whereas a few ohm-meters were a sure sign of an oil sand.
Bolder than their Grozny colleagues, the Baku geologists reached the
point where they could forecast from the diagram not only the order of
magnitude, but also exactly what the initial production of a reservoir
would be; they called this “translating ohms into tons.’’ Their enthusiasm
was such that, in spite of the warnings of the French engineers, they
abandoned mechanical coring altogether and conducted development and
even exploration wells under the sole guidance of electrical coring.29
Yet there were zones where the translation of ohms into tons could
lead to errors, and where even a qualitative diagnosis was uncertain. It
happened that in a field where electrical coring had just been introduced
resistivity peaks marked all the sandstones, whether oil- or water-bearing.
It took some time to realize that in such cases the mud had been prepared
with fresh water, and the invaded zone amounted to a resistant mantle
around the hole.30 After this experience and a few other experiments had
shown that diagrams alone would not solve all their problems, the Soviet
geologists gradually returned to more rational operating rules wherein
electrical and mechanical coring complemented each other.
On the other hand, electrical coring allowed for the use, in Baku, of
the “bottom to top” operating method. Standard procedure in most oil
fields was to exploit each horizon by a series of wells, systematically

29 In the summer of 1933 Maurice Martin, then assigned to Baku, was told that in lieu of
cuttings, cores, or geological cross sections the Azneft people showed long paper strips on
which undulating lines were supposed to represent electrical parameters. He was greatly
amused to find that the source of this “confidential” information was none other than
Messrs. Herold and Uren, the reputed American oil specialists who had been invited by
the Soviet industry for consultation.
30 The same difficulty was later encountered in many other regions of the world. The
solution was the use of a sonde with a sufficient investigation diameter.

156
The Origins and Development of Logging

spaced 50 to 100 feet apart. The separate exploitation31 of each of the


multiple superimposed horizons required the drilling of a huge number of
wells (hence the dense forests of derricks on old photographs). To obviate
such a proliferation, the Soviet engineers thought of drilling down to the
deepest reservoir, lowering and cementing a string of casing to the bottom
of the hole, putting the reservoir into production, plugging it once it was
depleted, then putting the next upper reservoir into production, and so
on to the top reservoir. Such a method required the accurate location and
delineation of each reservoir, and for this electrical coring was eminently
suitable. As holes had to be made through the casing and the cement at
the depth of each productive horizon, the Azneft engineers designed a
rather strange tool, consisting of a drill, which was lowered to the bottom
and drivén from the surface by cables and pulleys. It took several hours to
pierce a single hole at depths never exceeding 2000 to 2500 feet. In most
cases the pressure and permeability were such that a substantial flow of oil
followed the first perforation; the tool was then abandoned and the
wellhead shut in fairly quickly. This primitive method served in Baku for
several years without any serious accident, until it was replaced by gun
perforators of the Schlumberger type.
In 1931 and 1932 electrical coring, conducted by the Russians with
or without French advisers, was expanded to the Donetz coal basin, the
Maikop (North Caucasus) region, Georgia, the Emba Basin, and the
Ferghana Valley in central Asia.

A contract signed on November 8, 1928, with Shell provided for the


assignment of a crew to the Dutch Indies, but operations in Sumatra did
not start until early 1930. The main producing sands were very friable. All
cores recovered by the drillers were so fragmented and washed out by
mud that it was difficult to detect the hydrocarbons with the methods then
available.
This, indeed, was the main concern of the local geologists, but
electrical coring was unable to give a dependable answer, although the
results were excellent in regard to correlations and the delineation of
sands (even without S.P.). The mission chief, who attributed this failure to
the effect of deep mud invasion, saw a remedy in extending the lateral

31 It is important to prevent reservoirs from flowing into one another, to control water
intrusions, and to keep a precise production record for each pay zone.

157
The History of a Technique

sonde to lengths of 30 to 50 feet. It was not known at the time that, with a
lateral sonde of such length, insulation faults could entail substantial
errors in measurements. The curtailment of drilling operations resulted in
cancellation of the contract (end of 1930) before the question had been
clarified. Not until 5 years later was work resumed in the area.

A 4 month contract with Steaua Romana, a company for which


Schlumberger had already conducted several years of surface work, intro¬
duced electrical coring into Rumania (March 1931). The team of Hubert
Guyod and Roger Henquet had a power winch driven by the wheel of a
truck. The formations (a sequence of shales and sands) were not unlike
those at Baku. Experienced drillers, using advanced rotary rigs, drilled
through these soft layers to a depth of 6000 to 7500 feet with a single
surface casing, an operation requiring especially heavy, viscous mud (den¬
sity above 1.5). The 60 pound weight at the end of the 1 ton cable was
often too light to reach the bottom of the hole even after endless at¬
tempts. If additional weights propelled the sonde to bottom, pulling it out
was likely to cause unfortunate cable breaks with ensuing unpleasant and
costly fishing jobs. The prospectors fumed, demanding by letter and
telegram a stronger cable on the order of 2 or even 4 tons. Paris recom¬
mended more care in splicing and elaborate lubrication to reduce friction,
and even designed a detonator tool that could scuttle the weights if the
cable became stuck. Finally, everything was straightened out. The pros¬
pectors received their 2 ton cable and recorders, and as they became more
familiar with the borehole conditions operational efficiency increased.
The company appreciated the results of the electrical measurements to
the point of considering them indispensable, and so became more tolerant
of mishaps; it even accepted the fact that the wells needed to be
thoroughly prepared for logging operations by prolonged circulation of
mud of lesser density.
By the end of 1931 Schlumberger services were being provided to all
Rumanian and foreign companies: operations were in progress in the
fields of Gura Ocnitei, Aricesti, Boldesti, Moreni, and so on. Cores were
no longer taken in development wells; as in the U.S.S.R., diagrams were
used in support of the “bottom to top” production method. The technical
balance sheet was highly positive. Yet in Rumania, too, the depression
curtailed drilling, and by the end of 1932 the number of operations did
not exceed 200.

158
Phase Two: 1933-1940

An overall view

In January 1933, 8 not excessively busy crews were operating outside


the U.S.S.R.: 1 in Pechelbron, 1 in Morocco, 1 in Rumania, 2 in the
United States, 2 in Venezuela, and 1 in Trinidad.1 By the middle of 1934
there were 20 crews, and at the end of 1935, 40—many of which were
overworked. When World War II broke out, the number had reached
140, and more than half of these were in the United States. The others
were spread over some 20 countries, with a special concentration in
Venezuela and the East Dutch Indies. In comparison with the slow and
unsteady progress of the 1920’s, the Company’s growth had become
stable and continuous.
Paris was still the only research and development center, as well as
the main source of equipment. By 1936 one fully equipped truck was
produced by the workshop each month. Twelve-hour shifts and night
work were often required for such an output. The Schlumberger Well
Surveying Corporation (S.W.S.C.), a United States corporation estab¬
lished in Houston, Texas, in late 1934, was contributing to the procure¬
ment of the cables; it was also purchasing and assembling trucks for the
crews in the United States.
The offices, workshops, laboratories, and testing stations, all at Nos.
40 and 42 rue Saint-Dominique, underwent a substantial and necessary
expansion and reorganization after 1936. The testing station, located in
the basement, included, among other things, wells designed by Marcel,

1 According to the workload, a crew included one or two engineers and their helpers.

159
The History of a Technique

where instruments could be exposed to pressures and temperatures simi¬


lar to those encountered at the bottom of boreholes. Three of them, one
45 feet deep, had been built in 1933 for tests up to 1500 PSI. Three
others, for 9000 PSI tests at 400 degrees F,2 were installed in 1936. Only
the barrels and breeches of heavy guns responded to such specifications,
and the manufacturing facilities of the French army and navy were called
upon. Highly elaborated thermal treatments were required: it took 6
months to temper the steel in wood ashes.
Paul Charrin played a leading role in the management of operations:
outside of his sojourns in the U.S.S.R., he divided his time between Paris
and the United States, where Eugène Léonardon was resident manager.
Jules Miller was in charge of accounting; René Clairin headed the en¬
gineering office. René Seydoux had joined the management of “La Pros”
in 1931. New contributions were made in 1938 by Jean de Ménil, an
experienced financier, and the young scientist Eric Boissonnas.
The number of engineers on missions, or in Paris ready to depart, had
grown from 28 in early 1933 (the U.S.S.R. included) to 70 in July 1935
and was to exceed 100 in 1938. The engineering and manufacturing staffs
had increased commensurately. In 1934 Schlumberger recruited some of
the engineers who were to have long and distinguished careers with the
Company: William J. Gillingham, Louis Magne, and Alain Morazzani, all
of whom reached the top echelons, and Maurice Tixier, an outstanding
logging expert. Also in 1934, S.W.S.C. began to recruit an increasing
number of American engineers for its United States operations. The first
of them was Harold B. Markam.

Measurement techniques and interpretation methods

From 1933 on, S.P.E. ceased its surface work activities except for
incidental technical interventions. Electrical logging became its principal
activity, complemented by other initially modest but rapidly developing
processes: sidewell sampling and perforation (see p. 182 ff), teleclinometry,
temperature measurements, and dipmeter surveys. Additional services
technically related to logging involved water intrusions, well depth de-

2 Such figures correspond to 15,000 foot boreholes, a depth that was far beyond the
drilling record of the time.

160
The Origins and Development of Logging

terminations, water—oil contacts, and so on. Cables and winches also


served occasionally for lowering and igniting explosive charges (tor¬
pedoes) for the purpose of fracturing very compact rocks and facilitating
the drainage of oil or gas.
The practice of drilling oil wells with a degasified oil-base mud loaded
with clays originated in the mid-1930’s. Its purpose was to minimize the
contamination of potentially oil-bearing horizons during drilling. Whereas
an oil-base mud invades a reservoir by only a fraction of an inch, the depth
of invasion may reach several feet or more with water-base muds, and
affect the reservoir to the point where the production rate drops substan¬
tially.3 Also, when drilling with oil, the center of core remains practically
unaltered; this is not always the case with water.
These reasons explain why oil-base muds became popular, but their
use remained limited by high cost and the generally disagreeable condi¬
tions encountered in their handling. As oil is an insulator,4 there was no
electrical connection between the standard lead electrodes and the forma¬
tions. An induction logging system was the answer, but was not realizable
with the electronic techniques then known. Therefore Schlumberger
devised “scratcher” electrodes that rubbed against the borehole wall. One
model, a kind of bronze wire cylinder brush, was tested with some success
in 1937 in Venezuela and California. With sufficiently soft formations the
resistivity diagrams resembled those obtained with standard electrodes in
mud- or water-filled holes, and even some acceptable S.P. curves could
be recorded. However, such logs lent themselves to correlations only. In
hard formations, the contact of the electrodes was unpredictable and the
readings were meaningless. Beset by other operational problems such as
difficulties in lowering the scratchers into the borehole, the method found
few applications.
Great emphasis was placed during those years on making available to
the engineers accurate and systematic documentation on techniques and
instruments. Proselec (see note 10, p. 110), an invaluable source of infor-

3 The water of the mud displaces a large part of the oil or gas. As the adhesion of the water
to the grains of the rock is usually greater than that of hydrocarbons (water is said to wet
the rock better), the invaded zone acts like a seal, inhibiting the motion of the hydrocar¬
bons toward the well once it is in production. Furthermore, the interstitial shale may swell
under the effect of the filtrate, thus reducing the permeability of the invaded zone.
4 There is always a small fraction of water suspended in the oil, but in the shape of separate
droplets, and the mixture has no electric conductivity.

161
The History of a Technique

mation, was published regularly undl World War II, along with a number
of pamphlets and memoranda (Water Intrusions, October 1934; Sondes and
Weights, December 1934; Dipmeter, September 1933; Photographic Re¬
corder, 1937; Gun Perforator, 1939; etc.). With the hiring of American
personnel, some of this literature had to be issued in English. The first
English document, Principles of Electrical Logging Operations (January
1936), was followed by an issue of Proselec (November 1936) devoted
solely to S.P. Both were written in Paris in the usual Schlumberger format
and style. The writing of technical memoranda began in Houston with the
introduction of photographic recorders. In this respect it is interesting to
compare the two memoranda on the three-galvanometer recorder as
written in French in Paris and then in English in Houston, by the same
French engineer. The first, although concerned with practical directions,
reserves a large section for the functional description of its various com¬
ponents; the second consists mainly of push-button-type, complete, and
detailed instructions for use.
The electrode spacing of the normal sonde5 was set in each region
after a few tests and varied between 10 and 20 inches. Soft sand-shale
series contain many rather thick, porous horizons, with little invasion, so
that the measurements were close to the true resistivities. When the
lithology and the water salinity of these horizons were fairly uniform,6 a
comparison with production results in a few wells led to a sort of calibra¬
tion of the resistivities, and the basis for interpretation obtained in this
empirical way could be extrapolated to the whole field. Yet in other
regions, such as northern Texas and the Mid-Continent, where more
consolidated producing zones occur, true resistivities were rarely read on
the normal sonde diagram, mainly because of deep invasion, and it was
difficult to differentiate between hydrocarbon- and water-bearing forma¬
tions. For several years the long lateral7 (12 to 24 feet) was used from time
to time in these regions. There was nothing systematic in this approach,
one reason being that the overworked staff had little time for the record¬
ing of a third curve. Around 1937 the number of crews was increased;
they were better equipped and worked under more normal conditions,

5 The Baku operators for years kept the same 7.5 to 12 foot AMN devices as prevailed
when logging was introduced into the U.S.S.R.
6 This is mainly the case with waters of very high salinity (over 100,000 PPM).
7 In Trinidad a deep lateral sonde was frequently used to study very thick and deeply
invaded sands.

162
The Origins and Development of Logging

and with competition becoming a factor, the long lateral sonde was
offered as a standard service under the name “third curve with great
investigation range." In spite of its limitations, it was favorably received
and quickly became a standard tool.
Another improvement was introduced at the end of 1938 in the form
of the second normal. Thereafter, and for years to come, standard equip¬
ment included the short normal (16 inch), the long normal (64 inch), the
lateral, and the S.P. (Fig. 49). With the simultaneous introduction of the
four-conductor cable, the operation took place in two stages: first the two
normals and the S.P., then the lateral and again the S.P., so as to accurately
adjust the depths between the two logs. Sometimes one run could be
saved by making the lateral recording while lowering the cable into the
well. This was a compromise made to gather as many data as possible in
the shortest time. In many cases, however (e.g., consolidated formations
or horizons made of alternating thin layers such as sands, shales, and hard
beds), log interpretation remained uncertain. There was much room for
improvement, but it would have to wait until after the war, when focused
sondes would be put into operation.

Between 1930 and 1932 S. Stefanesco, already mentioned in the


discussion of surface prospecting, had laid the theoretical basis needed to
calculate the relationship between true and apparent resistivity.8 The
actual environment was simulated by a model in which the formation was
assumed to be homogeneous, isotropic, and infinitely thick; the wall of the
borehole and the limit of the invaded zone were represented by two
circular coaxial cylinders, on the axis of which the electrodes were assimi¬
lated to points.9 The curves thus obtained were plotted in logarithmic
coordinates, with the sizes of the device (normal or lateral sonde) as
abscissae and the apparent resistivities as ordinates. Since the curves gave
the value of the apparent resistivity as a function of the investigation
diameter, they were called “lateral sounding curves” by analogy with the
vertical electrical sounding curves of surface prospecting. However useful
in facilitating interpretation by imparting greater accuracy to the re-

8 The same problem was dealt with at the same time in Leningrad by Professor V. A. Fok,
but with a different mathematical approach.
9 There is no strict analytical solution when the media involved in the measurement are
separated by both cylinders and planes, as is the case in a formation having a finite
thickness.

163
The History of a Technique

Figure 49■ Log recorded with standard equipment: short normal, long normal, lateral, and S.P.

sponses of the sondes, they were of little help in actual cases in obtaining
the precise values of the true resistivities. To do this would have required,
among other things, that the formations be sufficiently thick and
homogeneous, a situation almost never encountered.

What needed to be defined more precisely to improve empirical


interpretation methods was the relation of true resistivity to saturation,
which, in turn, required many measurements on samples in a specially

164
The Origins and Development of Logging

equipped laboratory. Absorbed by technical developments and the de¬


mands of an industrial activity in full expansion, Schlumberger was unable
to cope with such a task. The first studies on the subject were done in
1934—1935 at the Petroleum Research Institute of Azerbaijan.10 They
covered a great number of moderately consolidated samples of sands
originating from several horizons of the same field, where a network of
production wells was exploiting every pay zone.
The method was, in each case, to determine the porosity and then to
proceed with a series of resistivity measurements by first impregnating the
rock with water, followed by the injection into it of oil.11 The results were
recorded as a family of curves in which each corresponded to an average
porosity value. Once the resistivity of the water had been measured for
each horizon, the authors used these curves to deduct the oil saturations
from the resistivities read on the logs. Combined with the porosity values,
the thickness of the formations, and the lateral extension of each oil¬
bearing zone, these saturation values made it possible to compute the
volume of oil (reserves) in each horizon. As a verification the same com¬
putation was resumed on one of the horizons on the basis of logs taken,
on the one hand, between the end of 1932 and the beginning of 1933 and,
on the other hand, between late 1934 and early 1935. Subtraction of the
two evaluations of the reserves showed a drop of 1,230,000 tons against
1,154,000 tons for the same period according to production statistics—a
difference of only 8 percent. Without minimizing the part played by
chance—after all, the statistics were not dated with great precision—the
accuracy of the method seemed reasonably confirmed. The authors had
postulated that, in a practical sense, the logs gave the true resistivity of the
horizons, and the result proved a posteriori that the assumption was
reasonably valid. It may be added that they were unaware of the effect on
resistivity of the interstitial clays; the horizons studied probably contained
so little of them that this effect was negligible. This research marked an
epoch in log utilization.
Observations made shortly thereafter in American laboratories on

10 See I. Kogan, Ispolsovanie Karottajnykh Dannykh dla Opredelenia Nieftianykh Zapassof


( Utilization of Electrical Coring Data for the Determination of Petroleum Reserves), Azerb-
Neft. Khoziatsvo, Baku, October 1935.
11 This operating method endeavored to repeat the natural process whereby, according to
prevailing theories, petroleum penetrates the reservoirs.

165
The History of a Technique

rocks of various origins proved consistent with the Baku measurements.1"


Finally, in an article published in 1938,13 Schlumberger presented a full
discussion of the relations between reserves and true and apparent resis¬
tivities. The article was a forerunner of what would later be called quan¬
titative interpretation.” Yet it still failed to give any numerical relation of a
general nature between saturation and resistivity; this meant that applica¬
tion of the method required long and costly work for each formation. For
this reason empirical interpretation processes continued to prevail for a
long time.

In pursuance of his studies on S.P., initiated in the spring of 1932,


Conrad devised a special apparatus for measurements to confirm the law on
electrofiltration potentials. The main direction of his research, however,
was toward electrochemical potentials. Actual ground phenomena were
simulated by means of three glasses (Fig. 50), the first and third of which
were filled with the same sodium chloride solution,14 representing muds
of R mf resistivity, whereas the second glass contained a more concen¬
trated solution (not exceeding, however, a few tens of grams per liter),
representing formation water of Rw resistivity. The first and the second
glasses were connected by a sand-filled syphon, the image of the perme¬
able layer, in which the contact of the two solutions generated an elec¬
tromotive force as in the concentration cells. The shoulders of the forma¬
tion were simulated by a sausage-shaped curve of clay connecting the
second and third glasses, between which another electromotive force, not
fully explained at the time, occurred. The potential difference between
the first and the third glass corresponded to the electromotive force
which, in actual cases, generates the spontaneous currents. These spread
through the ground and squeeze again inside the borehole; therefore,
when the permeable formation is of such thickness that its resistance to
the current is negligible, the ohmic drop in the mud column measures
practically this whole electromotive force.
Many measurements made with varied concentrations and types of

12 R. D. Wyckoff and H. G. Botset, Physics, September 1936; J. J. Jakosky and R. H.


Hopper, Geophysics, Vol. 2, No. 1 (January 1937); M. C. Leverett, Transactions of the
AIME, 1938, p. 1.
13 M. Martin, G. H. Murray, and W. J. Gillingham, Geophysics, Vol. 3, No. 3 (July 1938).
14 Sodium chloride is by far the most widespread salt in drilling muds and formation water;
mf and w are current symbols for “mud filtrate” and “water,” respectively.

166
The Origins and Development of Logging

Figure 50. Schematic representation of the laboratory device used by Conrad Schlumberger to
study the electrochemical component of S.P. (1932).

clay demonstrated that the potential difference between glasses 1 and 3 is


expressed in millivolts by the formula

E = K log,, -^SL.

The various values found for coefficient K averaged 17, whereas more
recent studies have shown that its precise value is 71 at 25 degrees C. The
reason Conrad found a much lower value was that in his experiments the
clays were under less stress than those in the subsoil, where they sustain
the pressure of the surrounding media. Although these results were
known to lack accuracy, the conclusion was drawn that the electrochemi¬
cal component of the S.P. was relatively small.15 This idea prevailed for a
decade, during which time nothing was done to better define the respec¬
tive parts of the two main components of S.P. All that was known was
that, in areas where mud and formation water resistivities were close and
rather high, the electrofiltration phenomenon might be preponderant, as
shown by tests in Burma (1934), Pennsylvania (1936-1937), and Califor-

15 Yet Tanida, Kyoto, Vol. 4, p. 1333 (quoted in Proselec, November 1936), had given a K
value of 58 for potassium chloride solutions (similar to sodium chloride) separated by a
collodion diaphragm, whose electrochemical behavior is comparable to that of clay.

167
The History of a Technique

nia (1934-1939). On the other hand it seemed that, when the mud was
relatively conductive, although less so than the formation water—as in
Baku or southern Texas—the electrochemical phenomenon perhaps pre¬
vailed. In boreholes where the mud was more conductive than the forma¬
tion water the S.P. anomalies became positive, a discovery that could be
explained only by the preponderance of electrochemical phenomena.
At the same time, it was realized that the amplitude of the S.P.
anomalies, as well as of the apparent resistivity, might be a function of the
geometry and the resistivities of the media in which the current flows
(borehole, invaded zone, undisturbed zone, formation shoulders), and
that, in particular, it decreases opposite thin formations. It gradually
became apparent that the spontaneous potentials, continuous and low as
they are (usually below 100 millivolts), are exposed to perturbations of
various origins: stray currents (often caused by operations in a nearby
well), the vertical component of telluric currents, instability of the surface
electrode, voltages induced by the terrestrial field in the slack cable
between winch and well, uneven polarization of the metal of the sonde
(bimetallic effect), heterogeneous mud column, and so on. The protection
of S.P. measurements against such disturbances has been the subject of
careful and continued studies.

It will be recalled that during his early work Conrad had observed the
polarization induced in the ground by the flow of a direct current (I.P.)
and had thought of utilizing it in the search for conductive minerals (see p.
31). From the introduction of electrical coring, he had advanced the
hypothesis that rocks might behave like a system of resistors and
capacitors, in which the water would be the armature and the mineral
grains the dielectrics. Such a system would generate I.P. potentials reflect¬
ing the porosity and oil content of the rock. He hoped that these poten¬
tials might be substituted for S.P. whenever the latter lacked sharpness,
and complement the resistivities in the identification of oil-bearing hori¬
zons. A few tests in the U.S.S.R. between 1930 and 1933 confirmed that
I.P. generated measurable potentials, but no clear correlation could be
established between those and other properties characterizing the forma¬
tions.16 Schlumberger was then so engaged in urgent developments that

16 It had not occurred to Conrad that the clays might have a part in the mechanism of I.P. It
emerged much later that their role was a preponderant one; hence the values he had
measured had a meaning different from his interpretation.

168
The Origins and Development of Logging

the continuation of a study holding little promise was postponed


indefinitely.

As logging spread over the world, new problems arose daily. It was
noticed that the presence within permeable rocks of very fine particles,
mainly clays, may substantially reduce the resistivity of oil-bearing strata.
In certain regions like southwest Texas, the reduction could be so sig¬
nificant as to prevent petroleum diagnosis. Thus there was, by 1936, a first
awareness that interpretation could be complicated by the presence of
clay in the reservoirs—a prelude of the many studies made after World
War II on the “shaly sands” problem.
Other problems arose in the study of limestone fields. It was found
that S.P. 'diagrams gave only broad, approximate limits of limestone series,
and showed only vague inflections through permeable or shaly intervals.
In these limestone series it was also observed that, contrary to what occurs
in sand-shale series, the pay zones were generally more conductive than
the impermeable adjacent formations. Furthermore, since limestones usu¬
ally have very high resistivities, most of the current was channeled along
the borehole, which made the diagrams confusing; it was difficult to read
the depth of the contacts between formation boundaries, and there was
only a remote relationship between apparent and true resistivities. In
most cases nobody could tell whether the conductive intervals corre¬
sponded to permeable zones or to marls, let alone differentiate water from
hydrocarbons. Moreover, it frequently happens that limestone fields are
topped by thick layers of rock salt, which saturate the drilling mud and
result in still more featureless resistivity curves and the reduction of the
S.P. curve to an almost flat line. Not very successful attempts were made
to revive the sharpness of the resistivity logs by fitting insulating sleeves
on the sonde between electrodes. In certain cases it was possible through
thermometric measurements to locate roughly the potential gas-bearing
zones, but such information remained of only local and contingent use.
The utility of logging in limestone fields thus continued to be very
limited. In the 1950’s this situation was remedied by the introduction of
improved measuring techniques. It is quite conceivable, however, that if
early logging tests had been conducted only in limestones the failure
might have been final. Fortunately, Schlumberger’s reputation had been
so firmly established by the Company’s success in logging sand-shale
basins that, when the limestone problem had to be tackled, the partial

169
The History of a "Technique

failure of the method was ascribed by professionals to the particular


circumstances rather than to any shortcomings of the technique.

Cables, winches, and miscellaneous tools

From 1928 (see p. 109) to about 1938 the types of cables hardly
changed. The tensile strengths were 2, 4, and finally 8 tons. As these
values meant more steel, increased tensile strength involved higher den¬
sity, and the running in of the sonde became easier. In 1938 a fourth
insulated conductor was added to the cable (see p. 163).
The protective braid was still made of a textile fiber: first hemp, flax,
or jute, and somewhat later ramie for its higher resistance to abrasion and
decay. Yet in spite of the utmost care in its preservation—drying in the
sun, spraying for disinfection—the braid would be practically worn out
while the insulated conductors were still intact. The action of oil-base
muds was especially harmful. Friction tape was used to patch up thread¬
bare or torn-away sections to the extent that the braid was gradually
replaced over its full length. In many cases, especially when replacements
became scarce during World War II, shreds of braid would dangle be¬
tween winch and well like old clothes on a line, and the cables were called
“rag lines.” However unsightly, the situation in no way lessened the
strength and insulation of the conductors and did not affect the accuracy
of the logs. What the drillers did not appreciate were the bits of friction
tape carried by the mud through hoses, pumps, and valves.
Certain formations yielded gas, or oil charged with gas, which under
high pressure and temperature became embedded in the natural rubber.
On emergence from the hole, the expansion and discharge of the gas
produced open blisters in the rubber. This happened mainly when perfo¬
rations brought an immediate flow of hydrocarbons, rather than in the
course of logging operations. Shortly before the war the introduction of
synthetic rubbers (Duprene and neoprene) as insulating materials began
to obviate this shortcoming.
For years sondes were made of a rubber-coated section of tricable,
and electrodes of lead wire coils. The weights were of cast lead or brass.
To ensure that the measurements reached as closely as possible to the
bottom of the hole, weights were placed above the AM electrodes of the
normal tricable sonde; the lower end of the sonde carrying the electrodes

170
The Origins and Development of Logging

was housed in a kind of cage, whose rubber-coated bars had adequate


rigidity to maintain the electrode spacing when touching the bottom.
About 1936 weights and sonde were consolidated by placing the elec¬
trodes on a long, insulated steel sleeve, and the first connectors (heads)
were built; instead of lengthy and inconvenient splicings and unsplicings,
the cable and sonde could now be readily connected or disconnected. For
depth measurements the braid was marked at regular intervals while the
cable was under tension when being pulled out of the hole.

Three models of winches were operational by 1934: a rarely used


lightweight unit, a medium-duty type for depths around 4500 feet, and a
heavy-duty model for deeper wells. The first two were mounted on the
platform' of the truck. The lightweight unit was hand driven. The
medium-duty type had an electric motor drive with gearshift, but could
also be driven by a rear wheel of the truck. The heavy-duty winch, as in
current equipment, was an integral part of the chassis. It was driven by
both rear wheels of the truck, firmly propped up by two hinged jacks.
This winch was so sturdy and powerful that it could rupture a 4 ton cable
stuck in a borehole. The first trucks delivered had a berth behind the
driver’s seat; the space gained by its elimination became the measurement
cab with its two recorders, batteries, and other components: rheostats,
switches, and so forth. There was not much working space, but the
engineers were sheltered.
Later, in another model, the recording cab was separated from that of
the driver (Fig. 51). Because of the inelegant shape of the cab, not unlike
the outhouses behind many farm dwellings (Fig. 52), the truck was given a
rather crude nickname by the field crews. Although satisfactory in hot
climates, the arrangement was completely unsuited for the harsh winters
of the central United States. Whereas the engineer was reasonably pro¬
tected in his cab, the winch operator’s only shelter was a collection of
tarpaulins flapping in the wind and wide open in the back. In Kansas,
Colorado, and Wyoming, heating was provided by a blowtorch under the
seat. On the other hand, the roadability of the truck had been substan¬
tially improved by moving the winch drum (i.e., the heaviest part of the
equipment) to the fore of the rear axle.
In 1937 the California Division acquired a heavy-duty winch with a
12,000 foot 8 ton cable capacity. It could be driven either by the wheels
of the truck or by a power takeoff. The latter device, both more flexible

171
The History of a Technique

Figure 51. 1958-type truck: inside of the cab, displaying the two hand recorders, the cylindrical
housing of the pulsator, and the slanted stem transmitting the cable motion to the paper drums.

and more convenient, became standard on all Schlumberger trucks, first


in the United States and then all over the world.
In California, which was asking for installations polished to the brass
tacks, efforts were being made to correct the industrial untidiness charac¬
teristic of many oil fields and production centers. The Schlumberger
people had apparently felt the same urge, for it was there in 1936 that the
first truck appeared with its winch under a shiny paneled body. The
resulting enclosure was used for both winch and recorder operation. In
1935 blue, the color of the French automobile racing team, was chosen as
the standard Schlumberger color. The trucks, newly painted, soon be¬
came the “blue fleet”—to be seen and recognized in every corner of the
world where Schlumberger operated.
Among other accessories, the early wooden Roman balance carrying the
sheave was replaced by a more compact, all-metal A-frame with a dynamome¬
ter indicating the tension of the cable. The A-frame, like the Roman balance,
was attached to, and positioned on, the rotary table. When, after completing

172
The Origins and Development of Logging

Figure 52. The truck fully equipped. The winch has been located in front of the rear axle for
better load distribution.

measurements, the cable was pulled out at full speed, it sometimes happened
that the winch operator forgot to slow down in time and the weight or sonde
was pulled into the sheave; this usually meant that the cable broke and the
sonde and weight dropped to the bottom of the well. To enable the operator to
sight the emerging sonde immediately and stop it in time, an improvement was
devised in 1934; the cable was passed over a sheave suspended by the traveling
block some 30 feet above the rotary table. In the same year a first “logging
stuffing box” mounted on the wellhead made it possible to lower the cable
in wells drilled under pressure.

The Advent of photographic recorders

In 1933, when logging operations were resumed in the United States,


Venezuela, and other countries, the crews were equipped with hand-
operated recorders. Before the two instruments (one for the resistivity, the

173
The History of a Technique

other for the S. P. curve) found shelter in the cab of the truck, they were, as
in the early days of surface prospecting, placed side by side on their tripods
next to the vehicle.
Hand recording was simple enough: as the cable was pulled out of the
hole, the engineers watched the curves being drawn and took note of all
the perturbing incidents (equipment deficiencies, stray currents, and
other parasitic effects). However, when there were several thousand feet
of open hole to be surveyed and the recording took several hours,17 the
operators were placed under heavy strain; on top of this they later had to
make clean ink copies of the diagrams. On the other hand, certain com¬
panies required that all logging data remain confidential, especially for
“tight” exploration wells.18 Even though secrecy was a tenet of the
Schlumberger staffs professional ethics, these companies were reluctant
to see the paper strips, with their wealth of data which could often, so to
speak, be read like an open book, remain in the hands of the prospectors.
Hence there were several reasons warranting automatic photographic
recording: aside from its operational advantages, such a system would
guarantee secrecy since the client could receive the recorded log on
unprocessed film. Moreover, a number of automatic measuring devices
for all sorts of parameters were already on the United States (especially
the California) market, and clients who considered neatness important
tended to look down on the somewhat makeshift array of two hand
recorders with their plates and cranks.
Studies on a photographic recorder began in Paris in late 1933. The
Russians were pursuing the same project. The principle was the same on
both sides: to replace the operator by a servomechanism capable of
rotating the knobs of the potentiometer. The French solution was based
on a photoelectric cell. Triggered by the deviations of the needle around
zero, it actuated a servomotor through a system of amplifiers and vacuum
tube rectifiers. After laboratory tests, however, this arrangement was
found to be too slow and was abandoned even before a real field pro¬
totype had been built. The Russians, for their part, persevered. About
1936 they came up with a huge engine containing a servomechanism that
was much less agile than the operator’s arm, so that the system’s inertia
limited the recording speed to less than 1500 feet per hour.

17 The highest recording speed was then 1500 feet per hour.
18 Tight in the sense of proof against any leak of information.

174
The Origins and Development of Logging

Doll, meanwhile, had begun designing an altogether different ver¬


sion, no longer aiming at the automatic simulation of manual work, or at
utilizing the potentiometer and the null method—all inherited from
surface prospecting. The resistance of the M and N logging electrodes is
negligible in relation to the total resistance of the circuit, which remains
practically constant over the whole section to be logged. Since, on the
other hand, this resistance is very high in relation to that of the ground
between the two equipotential surfaces passing through M and N, it was
possible to use a galvanometer serving as a voltmeter, or rather a milli-
voltmeter. A first variation, begun in October 1934, included a high¬
speed mirror oscillograph-like galvanometer (made by Siemens) and a
special jmlsator. It produced a film on which the light spot shifted alter¬
nately from the S.P. to the resistivity curve (Fig. 53). A prototype was
tested in California in the fall of 1935, but was used only a few months
and then discarded because the oscillographic loops were not sufficiently
sensitive.
In another variation the circuits were the same as in hand recorders,
but the potentiometers were replaced by galvanometers (Hartmann and
Braun, A.O.I.P.).19 The 1936 two-galvanometer model, designed by
Maurice Lebourg, worked with an alternating-current generator and al¬
lowed for the successive recording of the normal and the S.P., followed by
the lateral (Fig. 54). As the recording of the lateral required as much as
0.5 ampere, the alternating current generated parasitic electromotive
forces (mutual induction and capacitance coupling between conductors) in
the circuit, whose compensation required a rather delicate adjustment.
The instrument was brought to California, where, in spite of its shortcom¬
ings, it quickly replaced the hand recorders. In 1937 there was a return to
pulsated current, with dead segments timed in such a way as to have the
measuring circuit closed after the inductive and capacitive effects had
practically disappeared (see p. 53). This was the year when the quadri-
cable and the long normal were introduced; to record the latter simultane¬
ously with the short normal, a third galvanometer was added, and a
three-collector pulsator was designed (one for the current and one lor
each normal). The three galvanometers constituted the elements of as
many cells; the cells were identical and could be interchanged by an

19 Association des Ouvriers en Instruments de Précision (Association of Precision Instru¬


ment Makers).

175
The History of a Technique

Figure 53■ Facsimile of a log recorded with the high-speed mirror galvanometer. On the
right-hand side, the two-scale resistivity; on the left-hand side, the S.P. (1935).

17 6
The Origins and Development of Logging

Figure 54. T ivo-galvanometer photographic recorder, as introduced in California in 1936-the


first of its kind ever put into service.

arrangement of resistors allowing for the desired sensitivities and damp¬


ings. Each cell further included a filter (resistor, coil, capacitor) that
absorbed most of the pulsated components. The three parameters were
recorded side by side with the same depth scale for both halves of the
film: the S.P. to the left, the two resistivities to the right. At the usual
depth scales (1/500 or 1/600), the maximum recording speed, as deter¬
mined by the inertia of the galvanometric cells, the characteristics of the
optical circuitry, and the film emulsion, was three or even four times that
of hand recording.
The requirements of the optical circuitry illustrate the difficulties
encountered in the design of these recorders. With a single light source
(an incandescent lamp) and through a system of prisms, screens, and
lenses, the recorder had to perform the following optical functions: the
light spots tracing the curves were duplicated for each track in a manner
such that, when the first spot left the film, the second spot entered on the
opposite side and ensured continuous recording; bold, lengthwise (verti¬
cal) lines separated the tracks, while thin, equidistant lines facilitated the

177
The History of a Technique

reading of the values measured; crosswise, equally spaced lines marked


the depths, whose 100 unit values appeared in figures; on a frosted glass
viewer the operator could monitor the recording through six different
symbols reproducing the motions of the galvanometers; finally, the long
normal curve was recorded as a dotted line so as to be distinct from the
short normal.
The truck engine made the galvanometers vibrate, producing on the
films a bold, smeary line that promoted neither neatness nor accuracy.
Putting the recorder on a football bladder proved to be an inadequate
remedy, so a better solution was devised: the tripod carrying the table of
the recorder was set on the ground under the truck, with three openings
for retrievable legs cut into the floor of the cab. With this arrangement the
recorder was physically isolated from the truck’s vibrations, but other
incidents occurred. A visitor in the cab might lean on the table and
reconnect it mechanically with the truck; this resulted in distorted and
vibrated logs at the corresponding depth. Sometimes, too, the tripod
would be overlooked when the truck, after completion of the operation,
began to move, and the consequences need no description. Another
omission, more frequent and equally disastrous, was failure to lock the
galvanometer for transport; the suspension wire would break, and the
instrument would have to be sent to a center equipped for repair and
calibration. In 1938 the study of a vibration- and shockproof galvanome¬
ter was undertaken. Initiated in Paris by Maxime Picard and continued at
the cost of long effort during the war years, it was to be counted among
the most significant and original achievements of Schlumberger (see p.
196).
A word must be said about the studies undertaken the same year in
Paris, at Doll’s initiative and with the major participation of Joseph
Bricaud, for the purpose of achieving a system that could measure two
resistivities and the S.P. with a single-conductor cable. The main purpose
was to lighten equipment, but the problems were many. The best that
might be obtained by using the electronic components then available,
both at the surface and in the sonde, were resistivity variations instead of
absolute values, and curves with baselines drifting with depth. Since such
a degradation of logging standards was unacceptable, other solutions were
sought, involving a sonde with a current source of double frequency, the
signals being separated above ground by filters or resonance galvanome¬
ters. It took great ingenuity to produce components whose geometrical,

178
The Origins and Development of Logging

mechanical, and electrical specifications were far above standard. Several


prototypes were tested in Pechelbronn. After more than a year of effort,
it appeared that no satisfactory result was in sight, and the project was
shelved.

Temperature, well drift, and formation dip

Only minor improvements had been made on the well thermometer,


which was used mainly to control cementing. The purposes of cementing
are, first, to prevent oil in production from mixing with oil or water
flowing from behind the casing, and, second, to seal off any communica¬
tions in the annular space lest a high-pressure aquifer gradually invade a
nearby oil-bearing horizon reserved for subsequent exploitation. It is
therefore important to locate the top of the cement behind the casing; and
since the setting of the cement releases a substantial amount of heat, this
top can be determined by recording the temperature of the mud in the
well a few hours after cementing.

The teleclinometer was intensively used in the U.S.S.R. (in the Baku
area in 1933, 794 operations involving 11,653 measuring stations in 426
wells) and gradually introduced wherever there were logging crews. From
1940 on, however, the teleclinometer was almost completely replaced by
the photoclinometer, a simpler instrument perfected by Marcel during
the 2 preceding years. Centering devices brought the axis of the photo¬
clinometer into practical coincidence with that of the borehole. It con¬
tained a steel ball rolling freely on a graduated spherical cup, and a magnet
kept horizontal by a universal joint suspension. The whole was photo¬
graphed at each station in a single exposure, which indicated the drift
angle of the well and its azimuth with respect to the magnetic north. A
series of consecutive pictures could be taken at intervals of 15 to 50 feet
on a 35 millimeter film electrically controlled from the surface. The
profile of the well could be easily computed from the data of each frame.

A first anisotropic—also called electromagnetic—dipmeter was sent


to the U.S.S.R. in early 1934. The instrument was introduced in Ven¬
ezuela at the end of the same year, and somewhat later in Rumania and the
East Dutch Indies. Perfecting it for field service was a long and difficult

179
The History of a Technique

process, requiring months of persistent effort and the know-how of the


field engineers, especially Lucien Beaufort and Louis Bordât. Field tests
were conducted by Doll in California (summer of 1935). A memorable
episode occurred when the two dipmeters available for the tests became
stuck in the wells, one after the other. This was a rather rough situation
for Schlumberger’s technical manager, who had come in person to test an
instrument he had invented and built. As he himself stated, a staff pros¬
pector, for the same mishap, would have been blasted with a tirade of
unprintable French and sustained a serious loss of face. It was Doll s good
fortune that one of the dipmeters could be fished out without too much
damage and repaired on the spot, and tests resumed in another well
without further difficulty. The dipmeter was put into operation in Califor¬
nia but could not compete with the equivalent and less expensive
Sperry-Sun process.
Aside from the lengthy and delicate operation required, the dipmeter
had other serious limitations: it could be used only where the formations
were sufficiently homogeneous and anisotropic, which practically re¬
stricted it to shales; the accuracy of the measurements could be affected
by the frequent occurrence of caves;20 to compute the true dip, it was
necessary to know the apparent dip from cores; finally, when mea¬
surements were made too close to compact formations, the latter’s influ¬
ence on the shape of the equipotentials could lead to erroneous dip
values.
Another type of dipmeter was built at Doll’s suggestion. This one had
three small electrodes, placed in a plane perpendicular to the axis of the
sonde and carried by three rubber arms at 120 degree angles, which
maintained electrode contact or near contact with the wall of the borehole
(Fig. 55). Three S.P. curves were recorded in sections straddling forma¬
tion boundaries. In a section exhibiting dip the deflection of each S.P.
curve at the level of a boundary occurred at a different depth. The amount
of this S.P. shift was a function of formation dip and of geometrical
elements (electrode orientation, well drift angle, and azimuth) as indicated
by a photoclinometer mounted above the electrode assembly. Without
any need for cores, the computation of these values gave the azimuth and

20 A dip measurement at the depth of a certain formation consisted in computing the


average of a series of readings made at intervals of a foot or so; the results of these usually
did not fully agree.

180
The Origins and Development of Logging

Figure 55. Schematic view of the S.P. dipmeter.

angle of the dip at each level where the boundary of two formations was
shown by a sufficiently sharp S.P. deflection, that is, mainly at the top and
bottom of the permeable layers.
A preliminary test, conducted in early 1936 in the Long Beach,
California, field with a locally made electrode assembly, had demonstrated
that the depth differences between the three S.P. curves could be
evaluated with adequate accuracy. The tool was further perfected in
France in 1938-1939. A. Claudet was instrumental in promoting its use in
Louisiana in 1941; thereafter it was marketed throughout the United
States.
The new dipmeter was not only faster, safer, and easier to operate
than the electromagnetic dipmeter, but also gave more complete data. It

181
The History of a Technique

would be followed by a series of instruments of ever-improved perfor¬


mance, called correlation dipmeters, based on the principle of comparison
between three or four resistivity curves recorded along the wall of the
hole.

Sidewall sampling and casing perforation

Logging had offered the petroleum industry subsurface cross sections


of any length desired by the operators, in which three types of formations
could be distinguished according to their electrical characteristics: shales,
compact rocks, permeable and porous formations, and among the latter,
in favorable cases, those containing hydrocarbons. However, the logs did
not reveal any other rock properties such as mineralogical composition,
shape and size of grains, porosity and permeability values, and types and
number of fossils. Although logging had led to a reduction of mechanical
coring, the latter remained, in many regions, indispensable in exploration
and even production. Yet it happens that mechanical coring takes place
during drilling and before logging: the decision on where to take cores is
based, therefore, only on correlations with nearby wells, oil shows in the
mud, or even remote geological extrapolations. None of these is a very
accurate guide, and the result is that coring does not always occur exactly
where it should. Hence there was a need for a tool which would take
samples after logging, that is, in the walls of the boreholes. Some instru¬
ments had already been designed for this purpose,21 but credit for the first
workable tools goes to Marcel.22
On the other hand, well completion techniques were rapidly evolving
toward running casing down to the bottom of the hole, then cementing,
and perforating at the level of the productive zones. By 1932 this tech¬
nique was already standard in Baku and in Rumania, and was spreading to
the western hemisphere. The Russian twist-drill perforator was slow and
undependable. In the United States the Lane Wells Company used for the
first time in December 1932 a more convenient and efficient process:
perforation by the firing of bullets through the casing. A similar device,

21 Among others, the U.S. patent of S. A. Williston, filed in 1925, granted in 1931, and
assigned to Sperry-Sun.
22 Sidewall sampling was originally considered as having the primary purpose of refining
and checking log interpretation.

182
The Origins and Development of Logging

though with a lower performance, was already in use in Rumania. The


perforating market was, in fact, still a nascent one, and it was not too late
for Schlumberger to enter the race.
Industry could only welcome good sidewall samplers and perforators
as auxiliary services to logging, especially when their operating cost was
reduced by the use of the same cables, winches, and trucks. Indeed, these
new activities complemented the logging from which they had emanated,
but their techniques and operations were quite distinct. For Marcel the
special mechanical and ballistic problems entailed in designing the new
tools opened a novel area of research suited to his frame of mind and
afforded his creative ability full expression. A group of engineers—first
among whom was Boris Schneersohn—and designers lent him valuable
support'during the many years when better and better models were being
devised.

The first sidewall sampler (1931) was a long mandrel inside which a
mechanism could swing a normally almost vertical lever around an axis.
The actual coring tool was a hollow cylinder of alloy steel perpendicular to
the lower end of the lever. The motion driving the coring tool into the
formation and then bringing it back with a sample was not unlike the
stroke of a pick. To swing the lever, various solutions were tried: an
explosive charge, oil pressure, motor power. No prototype went beyond
the experimental stage. In a later model (1933) the driving force was the
hydraulic pressure of the mud column. A prototype successfully tested in
Baku (1935) remained in the U.S.S.R.
The lever sampler was a heavy and complicated tool and could take
only one core at a time. Marcel also had the idea of a sidewall sampler
firing bullets. There has been little basic change in this tool since it was
first completed in 1934. The core bullet is a hollow tube, projected into
the wall by the deflagration of an electrically ignited powder charge. The
bullet is connected with the body of the tool by two steel wires, coiled
when the bullet is inside the barrel, and having a 16 inch range when
uncoiled by the shot. These wires retrieve the bullet and its contents
when the sampler is pulled out (Fig. 56).
The sampler was tested at Pechelbronn in April 1935, with a tempo¬
rary arrangement of three barrels a few inches apart. Each barrel was
connected to a conductor of the tricable and carried a 1 inch inside
diameter bullet with a core length capacity of 1 Ys inches. Satisfactory

183
The History of a Technique

Gun loaded ready for firing

Bullet shot into formation

Figure 56. Schematic cross sections of gun sidewall sampler.

184
The Origins and Development of Logging

results led to the manufacture of a commercial tool: it was a strong


mandrel, about 12 inches long, mounted at the bottom of a rigid sonde
(Fig. 57) and fitted with three adjoining barrels. One conductor of the
tricable crossed three series-connected igniters and ended on the body of
the sampler, which served as electrode A. The igniters were designed in
such a way that the measuring current was too weak to cause deflagration.
After the electrical log—simultaneous S. P. and resistivity—was re¬
corded, and while the sonde was still in the hole, the sampling depth was
selected. The partial repetition of the S.P. log ensured that the sample
would be positioned exactly where desired. To fire the three bullets
separately, the igniters were calibrated so as to reach combustion temper¬
ature with increasing current intensities.
Dispatched to the United States and Trinidad in early 1936, the
sidewall sampler was received by the drillers with mixed feelings; hollow
tubes thrust into the wall aroused misgivings lest they become stuck and

Figure 57. The first sidewall sampler, used as a current emission electrode.

185
The History of a Technique

result in cable breaks entailing long and costly fishing jobs. As the first
operations proceeded smoothly, these fears were promptly allayed.
Operating conditions tougher and more varied than Pechelbronn’s, how¬
ever, led to a gradual change in certain design features. For instance,
bullets were plugged by a small brass front plate which would snip off
under the impact, so that the mud could not fill them and oppose their
penetration into the formation. A whole set of front plates was required,
and the choice—based on the powder charge, the well depth and
diameter, and the nature of the formation—was a rather delicate one.
Most of the time the plates snipped off too soon or not at all and were
abandoned. Instead a hole was drilled at the bottom of each bullet, so that
the mud could escape when the bullet penetrated the formation. Another
improvement was to insert a disposable aluminum end plug in the bottom
of the bullet.
Three bullets in one run were too few, especially when, on the
average, one out of three was either lost or arrived at the surface empty
because of faulty ignition, a very hard formation, the sample being flushed
by the mud while coming out of the hole, or some other circumstance.
Based on the same design, a six-shot unit followed. Finally, the 1939
model comprised three six-shot units fitted in the same body and operated
by the quadricable (Fig. 58). A conductor went to the igniters of each of
the three units, the circuit being closed by grounding. There were now
two separate operations, each requiring a trip into the well: on the first
trip the log was recorded; on the second the sidewall samples were taken
at the depths indicated. The S.P. curve, recorded with one electrode
connected to the fourth conductor, positioned the sampler.
All these improvements, together with the engineers’ growing
operating skill, made the sidewall sampler an efficient tool in the soft but
cohesive formations constituting mainly the sand-shale series. The opera¬
tional yield (i.e., the ratio of the number of samples retrieved to bullets
loaded) reached an average of 70 percent. The data on formation charac¬
teristics gathered from such samples are not always as accurate as those
from conventional cores because the former are much smaller; hence
their analysis, whenever feasible, is less dependable. Such samples usually
come from formations where oil and gas, if any, have been displaced by
mud filtrate; or where the violent impact may alter the structure of the
mineral grains and consequently the porosity and permeability of the
formations. Nonetheless, because of its convenient operation and low

186
The Origins and Development of Logging

Figure 58. Sidewall samplers and perforators.

cost, the sidewall sampler found a wide field of application wherever


sand-shale formations prevailed: California, the Texas-Louisiana Gulf
Coast, Venezuela, Argentina, Rumania. The operation was often justified
by the mere fact that hydrocarbon traces showed in the samples retrieved.
Since the results obtained were less favorable in friable formations, a
sampler was designed that would prevent the loss of rock fragments when
pulling out. A valve at the front end of the core barrel opened upon
impact and closed almost immediately thereafter. Unfortunately, the tests
of this so-called sand bullet proved disappointing. Most of the time it
emerged empty, and the fragments retrieved were too few to characterize
the formation. The yield was likewise poor in hard formations: in spite of
all efforts to adapt the shape of the bullets, either they broke on impact,
or, for the most part, the percussion cracked the rock and only a few
worthless chips were retrieved.
In formations where retrieval is good, particularly in shales, sidewall
cores can often be indicative of stratification. This, it was thought, could

187
The History of a Technique

help in the computation of the dip, on the assumption that the actual
positions of the cores in situ, especially their azimuths, could be deter¬
mined. The idea was to complement the sidewall sampler by adding a
device like the photoclinometer, but the difficulties, especially the
vibrations produced by the firing, were such that the project was dropped.
In connection with the subject of sidewall sampling, Conrad’s at¬
tempt in 1935 to build the so-called “sucker,” a device to draw the liquids
out of formations, is worthy of mention. At that time, however, what
seemed to be a promising concept got no further than the drawing board.
It was successfully taken up again 20 years later.

At the depth set for perforation, a well may be cased by two or three
coaxial strings. The bullets were therefore expected to pierce one, two, or
three steel casings—each more than Y% inch thick—and two or three
sections of cement, while maintaining enough momentum to penetrate
the formations and open drains. Furthermore, the bullet had maximum
efficiency when fired perpendicularly to the casing wall, which left a bare
4 inches for the barrel and the powder chamber. To avoid excessive
charges with the risk that their deflagration might deform or even rupture
the powder chamber, Marcel’s first idea was to have a longer barrel,
parallel to the axis of the well, with a muzzle bent at a right angle. A
preliminary test is said to have been made with a hot-bent shotgun barrel.
At any rate a prototype was built in which the bend was lined with needle
bearings to facilitate the turn of the bullet. After a few tests, this “gun to
shoot around corners” (everyone, from workshop to field, made fun of it)
was also abandoned. Moreover, since perforation with bullets was claimed
in a patent granted to Lane Wells in the United States, it seemed prefera¬
ble to look for a different concept. A hydraulically controlled punching
tool, designed and worked out by Doll, was tested in Pechelbronn, the
U.S.S.R., and Rumania (1935). From punch ruptures and other unlucky
incidents it soon became apparent that its development would require a
long period of work, and so the design studies veered toward a system
with an electrically driven twist drill. Unfortunately the tool resulting
from studies initiated in 1936 proved too complex for current field
service.
Meanwhile, Marcel, backed by considerable experience already ac¬
quired with the sidewall sampling gun, had reverted to the idea of bullets.
He designed a powder chamber surrounding a transversal barrel; in this

188
The Origins and Development of Logging

way not only could the latter be given all available length, but also
deflagration gases built up a very high pressure on the bottom of the bullet
by the time it began to move (Fig. 59). The perforator based on this
principle was made of three adjoining cylindrical steel blocks (Fig. 58).
The barrels were screwed into chambers (four to eight) machined into
each block at regular intervals (8 to 12 inches) along helicoidal lines. Each
chamber received powder in bulk; the ignition system was much the same
as for the sidewall sampler. Each block could fire a burst if desired: all that
was necessary was to substitute an element with a discharge port between
contiguous chambers in place of the usual igniters; the deflagration would
then automatically spread to all the chambers after the first one had been
electrically fired.
r

Figure 59. Schematic cross section of the annular chamber perforator barrel. The dark object in
front of the bullet is a rubber plug ensuring watertightness before firing.

189
The History of a Technique

A first prototype was completed in a few months, and in 1936 two


series, 100 and 85 millimeters, were ready. Strict safety measures had to
be instituted, especially for handling on the well sites. The electrical
ignition system was foolproof. A special tool used to screw the barrels
into and out of the loaded chambers included a man-sized steel tube set
on the muzzle of the barrel to protect the operator from burns in case of
an accidental deflagration.
Thereafter, except in the United States, the Schlumberger per¬
forators became operational wherever its logging crews were at work.
Strong marketing incentives were the Company’s already existing local
facilities and the reputation which it had earned. Perforating services were
offered in the United States (1938) after settlement of a patent suit and
were successfully developed in spite of active competition.
The gun perforator underwent various improvements over the years.
One of the most remarkable was the “lined bullet” whose base was
inserted in a light metal sleeve, released immediately past the muzzle; the
muzzle velocity and penetration power were substantially increased.
As perforation became the completion method for reservoirs of
increasing thickness, the oil companies, as, for example, on Lake
Maracaibo, requested the shooting of hundreds and even thousands of
holes in a single oil well. In response to such a demand, it was not always
sufficient to bring a battery of preloaded guns to the well; most often they
had to be reloaded on the site, thus entailing many hours of continuous
work. To facilitate and accelerate large perforation programs, Marcel
undertook in 1937 the design of yet another model made of units assem¬
bled end to end, each comprising three radial barrels and three small-
diameter chambers parallel to the axis of the instrument. With the cham¬
bers set at 120 degree angles, the barrels had the greatest length compati¬
ble with the size of the instrument (Fig. 60). The powder was contained in
copper-lined cartridges; in bursting them, the gases built up the initial
pressure required. Cleaning and reloading the chambers was both more
convenient and speedier than for the other perforators. The assembly of
the required number of units increased the number of bullets fired in a
single run without any loss of penetrating power. The weak point, as
demonstrated in rather extensive use of this perforator in Sumatra (1939),
was the electrical connection between units. No correction could be made
before 1940, and further development of the unit perforator had to be
suspended during World War II.

190
The Origins and Development of Logging

Operations

Under a new contract with Shell Oil, one crew (Gilbert Deschâtre
and Jean Legrand) was assigned to California in July 1932, and another
(Jean Mathieu) to the Gulf Coast in early 1933. Shell Oil supplied the
vehicles and drivers, and the crews were free to work for other clients.
The service was paid for at a fixed rate per operation.
Initial operations got under way with a minimum of physical facilities.
The prospectors’ only office was at their place of residence, and space for
the workshops and stores was rented from garages. In spite of such
modest means, the potential market to be served included many clients
scattered over an immense territory. Once the crews were able to survey
several wells in the same field, they quickly demonstrated, particularly by
the discovery of numerous faults, that correlations furnished by logs gave
an accurate representation of the subsurface; in fact, the correlations often

191
The History of a Technique

changed the prevailing concepts of the geologic structure. Enhanced by


the successful location of several pay zones, these results made such
effective publicity that after 1 year a commercial breakthrough took place,
and the road was open. For greater operational flexibility Schlumberger
canceled the Shell vehicle supply arrangement. From this time forward
the Company became the free and sole master of its working tools,
offering its services on a performance basis according to rates fixed on the
actual depth reached by the sonde, the interval logged, the mileage to site,
the stand-by time, and similar factors.
With four crews on the Gulf Coast, three in California, and one in
Oklahoma, the volume of business had grown by 1934 to a point where an
organization under American law was in order, and the Schlumberger
Well Surveying Corporation (S.W.S.C.) was established under a contract
of technical support with the Société de Prospection Electrique. Business
was expanding rapidly (3225 operations in the United States in 1935), and
seven districts were created: one in California, two in Texas, three in
Fouisiana, and one for the Mid-Continent, each led by a chief engineer.
Fegal and financial staff began to join the Houston headquarters.
There was a language problem with the American engineers, and a
technical department was set up in 1934 for better liaison between the
United States and France. An equipment department followed in 1936,
whose main functions were to assemble the items coming from France on
American truck chassis, to centralize procurement, and to store instru¬
ments and machinery. The offices and sheds of yesteryear were out¬
moded: land was purchased and construction begun, first in Houston and
then all over the world.
From 1933 on, operations, while continuing in Trinidad, were re¬
sumed in Venezuela and included the Cumarebo (Falcon) and Quiriquire
(Monagas) fields. The discovery and exploitation of new fields in Peder-
nales (1935), Temblador (1936), Oficina (1937), and Jusepin (1938), all in
eastern Venezuela, afforded a new expansion of Schlumberger services in
the country. With a crew in Ecuador (1935), and soon thereafter one in
Colombia, 23 engineers were at work in this part of the world. There was
also a new start in Rumania, where, under the leadership of Poirault, four
or five prospectors had been assigned by the time World War II broke
out. In Germany the first logging took place for Elwerath (February
1933).
Although local circumstances varied, the introduction of Schlum-

192
The Origins and Development of Logging

berger services into numerous other countries had one common aspect: in
spite of agreements concluded at “the highest level,” the crews were
usually received by uninformed and reserved people. Two examples out
of many may be cited. Under a contract with the Burmah Oil Company of
London, Sauvage conducted, at the end of 1934, some 10 logging opera¬
tions at Yenanguang, Burma. Yet the local geologists, confined in their
routine, showed such little interest in his cross sections and structural map
that it took the intervention of a young engineer just arrived from
England, who fully understood the meaning of the data, to talk his
colleagues into using the logs as working tools. Similarly, Gabriel
Guichardot and Elie Paulin went to Argentina under contract to Shell Oil,
but 2 years of effort was required to enlist the cooperation of the heavy-
handed Y.P.F.,23 in spite of satisfactory results in the Patagonian oil basin.
Nevertheless, logging was spreading from year to year: Assam
(1934); Morocco (1935); Austria, Colombia, and Japan (1936); Sumatra,
which in 1938 became a rather large operation with 10 engineers; Poland
and Hungary (1938); Iraq and Kuwait (1939). Such expansion on an
international scale could not avoid competition, and in its footsteps fol¬
lowed patent suits and infringement actions. Even though, in 1942,
Schlumberger lost the benefit of its basic patents, such conflicts were
salutary inasmuch as they demanded increased efforts to compensate for a
measure of relaxation into which, at least in the United States, the
Company had been lulled by a period of exceptional security and prosper¬
ity.

Thus, before World War II, Schlumberger teams were already all
over the world. Each engineer, whether alone or serving as the chief of a
mission (the word “mission” had not yet become a clear-cut concept), had
physical, technical, legal, and commercial responsibility. The crew, the
basic unit described in a brochure captioned Operations sur le Terrain
(Field Operations), was in fact a kind of spontaneous outgrowth, a logical
result of similar experiences in different parts of the world. Consisting of
one engineer and two driver-helpers,24 the crew was on duty around the

23 Yacimentos Petroliferos Fiscales. The first operation in Argentina (and in the southern
hemisphere) took place on November 30, 1934.
24 According to the brochure, they were “conscientious, sober, serious-minded, punctual,
dedicated, healthy, and honest workers, who could be depended upon at any time. For

193
The History of a Technique

clock, 7 days a week, ready to answer the client’s call and drive the truck
with its required equipment to the drilling site. Whatever the duration of
the operation (with hardly a chance for a meal or a rest), it could well be
one in a continuous sequence of equally demanding tasks. In many
regions permanent readiness was expected for the servicing of far-away
wells requiring expeditions of several days and employing various con¬
veyances. Most of the time the Schlumberger engineer would thoroughly
examine the diagram with a geologist or engineer present on the site, and
his professional pride was great when the log pointed to a probable
productive zone otherwise undetected during drilling. Back at his base he
had to draw fair copies of the logs, clean and, if required, repair the
equipment, and only then think of sleep. Sometimes an engineer-trainee
was assigned to the crew, and the helpers were only too happy to relegate
to him the most tiring and dirtiest part of the work. It was his initiation.
Only much later would he deal with more refined tasks, such as cables,
film, and control panels.
It was a tough but exciting job, completely unpredictable as to work
schedule, yet, outside of work, unhampered by any constraint. Once in a
while, of course, fatigue, isolation, and harsh climate produced clashes,
but as a general rule teamwork and comradeship prevailed in the small
group without impairing discipline and mutual respect. In countries
where language was originally a problem, contact with assistants and
drillers soon made the prospector fluent in profanity—which he would
then use elsewhere in total innocence. From Venezuela to Indonesia,
from Assam to Oklahoma, these hardworking crews were proud of a life
so different from that of others. They could depend only on themselves,
but they went along, solving their problems, and on the whole giving their
best at the cost of great effort.

this reason, their salaries were slightly above average. “With all that is expected of
servants,’’ asked Figaro, “does your Excellency know many masters worthy of being one?”

194
Phase Three: 1940-1945

France
*

Depleted by mobilization at the beginning of World War II, the Paris


center endeavored mainly to supply the missions, all of which were
continuing their operations because most of the field engineers had earlier
been classified as on ‘special duty.” In June 1940 the headquarter’s staff
fell back on Clairac in southwestern France, but returned to Paris during
the summer, leaving behind a small nucleus at Saint-Gaudens with the
staff of the Compagnie Générale de Géophysique, which was undertaking
electrical prospecting in southern France. As this was in the unoccupied
zone, this echelon managed to maintain a few contacts with overseas
centers, especially in the United States, and to assist demobilized prospec¬
tors, or those who had escaped from prisoner of war camps, in rejoining
the missions.
The Paris technical staff was reduced to some 10 engineers and 30
technicians, draftsmen, and workers (compared to about 80 before the
war), all disheartened by events and none too eager to go back to work. It
took all the imagination and energy of Marcel to get things going again by
setting up research objectives for the staff and to maintain morale as well
as he could. Food and transportation shortages, air raids, police raids,
poor heating, gas and electricity failures, and, later, bombing raids—
none of this was conducive to creative thinking, particularly since
stimulating contacts with the field had essentially vanished. Yet a few
projects were advanced, especially the vibrationproof galvanometer,
which, with many improvements, remains a part of today’s recorders.
The principle of this instrument was to fill its housing with a liquid of

195
The History of a Technique

such density that the Archimedes forces were equal to the weight of the
coil. Since the latter then behaves like an integral part of the liquid, it does
not tend to move with respect to the liquid or to the housing when the
system sustains an acceleration.1 It is, in particular, unaffected by truck
vibrations; moreover, the torsion wire (or rather ribbon) is no longer
under tensile stress (thus cannot break), and it becomes unnecessary to
lock the coil. Since the centers of gravity of the coil and the liquid
coincide, the angle of the coil with respect to the vertical is immaterial, a
valuable asset whenever it is impossible to station the truck on a strictly
horizontal platform.2 Additional desirable features distinguish the instru¬
ment. It is small enough for nine units to be housed in the recorder. Its
sensitivity can be accurately adjusted by a magnetic shunt as demanded by
circumstances. It can sustain 100 times its design voltage and still return
perfectly to zero.3 Even informed observers marvel at the sturdiness of
this galvanometer under the roughest handling.
The inventor, Maxime Picard, had to overcome tough theoretical and
technical problems before completing such an outstanding instrument.
While meeting the specifications set for volume, accuracy, sensitivity, and
damping, he had to find a liquid that had the desired density, viscosity, and
refraction index and was, furthermore, neutral with respect to the metal of
the housing.4 The equilibrium of the coil inside the liquid necessitated
highly delicate adjustments. The reduction to zero of optical dispersion in
a system working with white light required study of the placement of
lenses between the mirror of the coil and the window of the housing,
which led to endless and inconclusive calculations. The problem was
solved by closing the window with a spherical lens centered on the axis of
the coil. Much work was still required to bring the galvanometer to final
perfection. This was not achieved before the war ended. Felix Barreteau, a
young engineer recruited in Paris, made a decisive contribution.
Another long-term project was the improvement of the remote
transmission system. The monocable solution, studied before the war, was

1 Only a rotation around the axis modifies the relative coil-housing position.
2 This advantage is even more valuable in marine operations aboard rolling ships.
3 This is equivalent to an ordinary 110 volt light bulb being unaffected by a voltage of
11,000 volts.
4 What proved impossible was to prevent the etching of the metal by the liquid in the long
run and the ensuing formation of microscopic gas bubbles sticking to the connection of the
wire and modifying the torque and the response of the instrument.

196
The Origins and Development of Logging

again put on the drawing board with the objective no longer merely
lighter equipment, but also shorter operations by simultaneous recording
of the lateral and other parameters. Marcel then reformulated the prob¬
lem in broader and more flexible but ambitious terms, namely, to record
four parameters with three conductors, five parameters with four conduc¬
tors, and so on. Two years were spent on this effort, made more difficult
because all testing had to be done in the laboratory since Pechelbronn was
occupied by the Germans. The work proceeded by trial and error with
great imagination and skill, but doubtful success, until the idea for a new
device, the sequential (later called the chronological) system, originated in
the mind of Bricaud. The idea of multiple frequencies was abandoned; the
measurements were to be made according to the standard technique, but,
at each depth, one after another and in a given order. The result was
obtained by appropriate timing of the pulsator, allowing signals to be
switched into the proper circuits. With the Houston technical services
participating,5 the project was completed in 1946. An important question
to decide was the number of conductors required in the armored cable: on
the basis of tests conducted in France, Bricaud again proved conclusively
that a correct simultaneous recording of S.P., two normals, and one lateral
required six conductors. This was a far cry from the lighter equipment
advocated earlier by Marcel, but the system was operational and served
for many years.
Other studies sought to improve minor technical features of per¬
forators, winches, the photoclinometer, and the S.P. dipmeter. Another
problem tackled during the first postwar months was the rational layout of
the various control instruments, such as rheostats, switches, and amme¬
ters, used in all operations (logging, perforating, sidewall sampling, etc.).
Except in the United States all these instruments had, until now, been
lodged in the cab according to the operators’ preferences and without any
fixed rule. Le combine (a multipurpose wiring system) was designed to
obviate such diversity: it was a kind of big trunk housing all surface
electrical controls, but if the look was neat and orderly, there was a
drawback to its multiplicity; the slightest malfunction entailed a total,
time-consuming and laborious disassembly to locate the trouble. “Le
combiné” was unable to prevail against the system soon to be adopted in

5 Even though the state of electronic technology in the United States at that time might
have supported a multiple-frequency solution as the natural one, the fact remains that
modern electronics makes extensive use of chronological systems.

197
The History of a Technique

the United States, whereby each operation was controlled by a separate


and interchangeable panel.
A project initiated before the war and subjected to rather thorough
tests involved a device operating at the lower end of the cable for the
purpose of injecting cement rings behind the casing and preventing
selectively the circulation of fluids between formations. There were to be
guns similar to those of the perforator, containing a cement slurry or any
other suitable sealing liquid; once the holes were punched, the liquid
would be projected by a piston driven by a gas enclosed in a high-pressure
chamber. The testing station saw the end of this project.
Migaux, who, until late 1942, was the manager for administration and
operations, devoted his forced leisure to a thorough study of the mecha¬
nism of electrofiltration. Although he had no laboratory at his disposal
and had to rely solely on a large body of documentation, theoretical
reasoning on the basis of his high competence in physics led him to a
remarkable analysis of the phenomenon. It is regrettable that this work
was never published.
Finally, it will be mentioned for the record that at the beginning of
the winter of 1940 Marcel, to ensure a modicum of comfort to the staff,
designed a heating vest with electrical resistors.6

The United States

After the Munich Agreement of September 1938, S.W.S.C. began


planning for the worst by expanding its local supply sources and manufac¬
turing facilities. At the same time, Paris shipped to Houston the drawings
and specifications of all the equipment items being built or planned. Not¬
withstanding these precautions, after the 1940 armistice, with Paris com¬
pletely cut off, Houston had to face the huge task of providing for the
needs of 60 American crews, not to mention those in other parts of the
world.7 Difficulties increased after Pearl Harbor, when the whole of
United States industry was mobilized for the war effort and submitted to

6 A patent was even granted, captioned “Improvements to Individual Heating Systems.”


7 These had become solely dependent on Houston. The war made it necessary to consoli¬
date them into legally independent companies, which still exist: Schlumberger Overseas
for the Far East, and Surenco for Latin America except Argentina, where the Compagnia
de Investigaciones Geophysicas Schlumberger was established.

198
The Origins and Development of Logging

Draconian regulations. The War Production Board ordered rationing and


priorities for strategic raw materials: steel, nonferrous metals, and rubber,
as well as for a substantial number of manufactured items such as au¬
tomobiles, trucks, and tires. Although the massive slowdown in drilling,
for lack of tubular goods, entailed a drop in Schlumberger’s activities for a
few months, petroleum production was soon given high priority, and the
Administration granted the necessary steel allocations. An increased de¬
mand for logging followed; but, deprived of part of its staff and hampered
by shortages, the Corporation had great difficulty in responding in spite of
its reserves in cables, tires, truck chassis, and other items. None of these
difficulties, however, prevented Houston—with special thanks to Robert
Leger and Maurice Lebourg—from answering the call of duty. Certain
key items like pulsators, photoclinometers, and optical recorders hap¬
pened to be available in numbers about sufficient to supply the crews, but
galvanometers had to be copied. The servicing of all this specialized
equipment was done by a few technicians trained before the war; they
were the nucleus of a group that, when the war ended, was able to work in
close liaison with Paris.
The problem of cables was especially critical. Their composition
(alloy steel and copper wire, high-grade rubber) required hard-to-get
priority allocations. Repairs with friction tape became standard practice.
This was no longer a question of patching a few bruises, but rather of
substituting thousands of feet of textile braid. An electrical wrapping
machine using wide friction tape was built to speed up and improve the
repairs.
From 1943 on, Schlumberger services were granted top priority, and
more liberal allocations became available. But with over 100 worn-out
trucks in the United States alone, it remained extremely difficult to supply
the crews until the war ended, and the amount of maintenance and repair
work reached unprecedented dimensions, both in the workshops and in
the field.

When Doll arrived in Houston at the end of 1940, he initiated a


research and development program. His memorandum of March 21,
1941, records his study on S.P. in boreholes. Based on reasoning showing
his acute understanding of the current’s tridimensional distribution, the
study leads to an appropriate analysis of the effect brought to bear on the
shape and amplitude of the anomalies by the geometry and the resis-

199
The History of a Technique

tivities of the media where S.P. currents flow. This became the nucleus of
a comprehensive and rigorous treatise which Doll was to present in a long
1948 communication (see p. 229), a basic reference document for the
interpretation of S.P. logs.
Another research project on S.P. proposed by Doll was pursued for a
few years by André Blanchard. The principle was to submit the mud to
quick and continuous impulses, thus producing periodic variations of the
pressure and hence of the electrofiltration potentials. The hope was to
record by this process, at the level of permeable formations, a “vibrated”
S.P. characterizing them distinctly, whereas ordinary S.P. gives only vague
indications, as in limestone fields. The project never got out of the
laboratory.
Finally, in 1943, a device to tag formations by radioactive bullets
represented the first application of electronics by Schlumberger to either
surface or downhole equipment. It did not reach the field until 1946.

From 1942 on, Doll and several members of the Engineering De¬
partment devoted most of their time to work for the War Department. A
nonprofit organization, Electro-Mechanical Research (E.M.R.), was es¬
tablished at the initiative of André Istel to support the war effort; its stock
was shared equally by Schlumberger, Doll, and two other French
partners. The principal achievement of the new organization was a mine
detector carried in front of a jeep. Up to May 1940, Doll had already
done considerable work on this project on behalf of the French Ministry
of Armaments after being released from the army on special duty for this
purpose. At the end of 1942, he assembled a team8 to resume the project,
and within a few months an operational prototype had been completed.
The detector proper, located far enough forward not to be affected by the
metal parts of the vehicle, rested on wheels with special soft rubber-lined
rims, so that the ground could be scanned smoothly. To prevent the
whole thing from being blown up when driving over a mine, the wheel
pressure was reduced by a lever and countersprings. The electronic circuit
was designed in such a way that the presence of a mine ahead would
automatically and instantaneously set the brakes of the vehicle; the same
happened in the case of circuit failure, an essential safeguard for equip-

8 The team was composed of M. Lebourg for the mechanical part and Ch. Aiken and G. K.
Miller for electronics.

200
The Origins and Development of Logging

ment whose very vocation was to court clanger. This system also compen¬
sated for the slow drifts of the signal caused by the distortion of the
winding under high temperatures. Another electronic component, based
on the phase discrimination between emitter and receiver, eliminated the
stray signals produced by mechanical distortions of the detector or by
certain features of the terrain, like magnetic soil (e.g., a granite-paved or
slag-surfaced road) or seawater on a beach. About 200 detectors of this
type, some tank mounted, were made by the U.S. Army.
Toward the end of the war, Doll improved his detector by the
introduction of an original feedback circuit. An essential feature was a
variometer tube, which separates the in-phase component of a signal
almost instantaneously and completely from the out-of-phase component
—up to 100 times larger—and conversely. This phase selection system
became standard in the manufacture of hand detectors for antipersonnel
mines. In recognition for its financial contribution to the project, Schlum-
berger retained full commercial rights to patents granted relevant to
logging, particularly to those applying to the above detector system,
which was later to become a part of induction logging.
Finally, research work was conducted by E.M.R. on the automatic
guidance of missies.9

In January 1942 G. E. Archie of the Houston office of Shell Oil


Company published an article10 that became the basis of quantitative log
interpretation, which Schlumberger had announced in 1938 (see p. 166).
The article first presented the results of many laboratory measurements
on sandstone samples of very diverse porosity and permeability from the
Gulf Coast. From his observations the author had deduced a simple
relation:
Ro = FRW

where Rq is the resistivity of the sample fully impregnated with


mineralized water of resistivity Rw, and F is the “formation factor.”
Archie had further observed that F is a function of the type and character¬
istics of the rock, particularly its porosity, and that this function is ex¬
pressed by the formula

9 See New Weapons for Air Warfare, Little, Brown & Company, Boston, Chapter XXVI.
10 "The Electrical Resistivity Log as an Aid in Determining Some Reservoir Characteris¬
tics,” Petroleum Technology (T.P. 1422), 1942.

201
The History of a Technique

4>m
in which the exponent m usually varies between 1.8 and 2.2.
In addition, the author had analyzed the results of measurements
published by other researchers (see note 12, p. 166) in which the oil
present in the pores intervened, and had worked out the average formula:

r> _
Kt-cTi
J w

where Rt stands for the resistivity of the sample, Sw for its water satura¬
tion, and n for another exponent which, in most sands and sandstones
without interstitial shales (“clean sandstones”), seems to be close to 2. The
saturation was then given by the equation

or, again,

Archie underlined the approximate nature of these formulae and


cautioned against using them in cases more complex than those upon
which his study was based. He also called attention to the fact that mea¬
sured resistivity values frequently required corrections (effects of the
borehole, the invaded zone, etc.). The article contained examples of
applications to oil-bearing formations, where these corrections had been
relatively unimportant.
Supported by a vast experience and the prestige of Shell, the Archie
study was above suspicion of any commercial bias and aroused much
greater interest than would have been accorded a similar Schlumberger
publication. The proposed formulae offered a convenient way of evaluat¬
ing a reservoir; aside from confirming the value of logging and its large
commercial potential, they had a definite technical interest for Schlum¬
berger. However approximate, the Archie formulae helped in the under¬
standing of various instances of success and failure; by clarifying the
problems, they suggested the developments necessary to obtain good
results under conditions heretofore considered adverse.

202
The Origins and Development of Logging

One of the most remarkable studies made in the wake of Archie’s


publication was that of Tixier, then the head of Schlumberger’s Rocky
Mountain Division. To compute the saturation, F and Rw had to be
measured on samples of rock and formation water, respectively. Not only
were these measurements sometimes of doubtful significance, but also
they usually required rather long and costly, and sometimes unfeasible,
operations. Tixier’s main idea was to substitute logs alone for these
measurements, by using the resistivity R( of the invaded zone and the
deflection of the S.P. curve.
The application to the invaded zone of the Archie formula gave

FRm
Ri =
~sr
where Rm stands for mud resistivity, and N for water saturation (there
always remains a substantial amount of oil or gas that the mud filtrate
cannot displace).
When this equation was combined with that for the undisturbed
zone:

FRw
Rt
~sV
the result was

Rj _ Rm S2w

Rt Rw ~ST
where F has been eliminated.
As the horizons studied by Tixier in the Rockies were usually very
thick, had average porosity, and hence were rather deeply invaded, he
assumed that after certain simplified corrections the readings of the short
normal and of the long normal (or of the lateral) gave the values of Ri and
Rt, respectively. On the other hand, the many comparisons he made
between the amplitude of the S.P. deflection and the values of Rm and Rw
(as measured on samples) led him to conclude that the electrofiltration
component was practically negligible, and that the electrochemical com¬
ponent of the S.P. could be expressed by

E = -110 log 4^-

203
The History of a Technique

Finally, Tixier established empirically between St and Sw an average


relation which led to the formula for the calculation of Sw:

R-i
V
i
1
i o

R>n ç
^ W

This was far from rigorous: in particular, as the water content of the
pores diminishes gradually from the borehole to the undisturbed zone,
parameters Rf and St are rather vaguely defined averages. However, these
approximations did not prevent the Tixier method from being used in
many cases to take the greatest possible advantage of the logs, and this far
beyond the Rocky Mountains.
Tixier also deserves credit for having shown experimentally that a
reasonably approximate value of Rw can be obtained from the S.P. An
article by Humble’s W. D. Mounce and W. M. Rust11 brought a contribu¬
tion to this problem based on laboratory measurements concerning the
electrofiltration and electrochemical components. In the postwar years
the interest engendered by all these works motivated extensive research
that confirmed the predominance of the electrochemical component and
gave more accurate values for the coefficient in the formula defining it.

11 “Natural Potentials in Well Logging,’’ Petroleum Technology (T.P. 1626), September


1943.

204
PART
THREE

After 1945
Phase One: 1945-1957

The situation and the prospects at the end of World War II

Only in the western hemisphere, and primarily in the United States,


was Schlumberger able to maintain a continuous presence during the war
years, in spite of serious difficulties.
The services offered were few, and electrical logging (resistivity
logging and S.P.) was the primary activity. Aside from the Shell Oil
Company, few companies used the quantitative interpretation methods
advocated by G. E. Archie in 1942. It must be said, however, that most of
the problems relating to sand or sandstone reservoirs having little shale
content and drilled with comparatively freshwater mud could be solved by
standard four-curve logging: one S.P., one short normal, one long normal,
and one deep lateral. Only the difficulty in procuring equipment slowed
development in the Gulf Coast, northern and eastern Texas, northern
Louisiana, California, part of Oklahoma, the Rockies, and Illinois. On the
other hand, the results in limestone and dolomites, or in wells drilled with
salt-saturated mud, were practically limited to well correlations.
In 1937 the settlement of a lawsuit on logging patents resulted in
Schlumberger receiving a license from Lane Wells to perforate in the
United States, while Lane Wells received a license to perform electrical
logging. Each company was confident that it could make serious inroads
into the business of the other. However, Schlumberger had a decided
advantage in that it had been perforating for several years outside the
United States and had already developed the equipment to do the job.
Lane Wells, on the other hand, had to start from scratch developing its
own logging equipment. Consequently, Schlumberger began commercial

207
The History of a Technique

perforating operations in 1938, 2 years before Lane Wells could start its
logging services.
Taking advantage of its lead in logging, Schlumberger claimed that,
although depth measurements with a cable were not always perfect in
absolute value, they were still relatively consistent; and, therefore, it is
preferable to lower the perforator with the same cable which was used for
logging. The argument was a somewhat weak one because the conditions
in a cased well, filled for perforation purposes with a generally low-density
mud, were quite different from those at the time of logging (i.e., open
hole and a heavier mud). However, it had considerable appeal where the
productive sands were thin, 5 to 10 feet, and it led to the development of
substantial activity in the Louisiana and Texas Gulf Coast. With greater
well depths, these questions of depth correlation became increasingly
thorny and required a new approach.
On the other hand, Lane Wells had good equipment, their armored
cable was a particular advantage over Schlumberger’s rag line, and many
companies preferred to split the business by giving Schlumberger the
logging and Lane Wells the perforating; hence perforating never de¬
veloped into as large a business as had been expected.
Schlumberger’s auxiliary operations included sidewall sampling,
which was very successful in alternating sands and shales; temperature
recording, specifically used to locate the cement top and the rarely
occurring actual gas intrusions in the drilling mud—a technique im¬
properly applied for locating oil zones in Kansas; and, finally, dip mea¬
surements, much appreciated by geologists, especially in areas where the
generally quiet oil field tectonics had been disturbed by the upsurge of
salt domes. Although the anisotropy dipmeter was still being used in
certain parts of South America, only the correlation dipmeter was em¬
ployed in the United States, for it gave not only the azimuth of the dip (as
did the anisotropy dipmeter) but also its magnitude.
The truck had been somewhat modernized since its initial
standardization in 1936. It was an International or Mack weighing about 12
tons gross. The winch accommodated 12,000 to 15,000 feet of four-
conductor, rubber-insulated, and textile-braided cable.
The electrical equipment, designed for resistivity logging and
sidewall sampling, was luckily, because of the nature of the mea¬
surements, well suited for temperature logging; but the dipmeter opera-

208
After 1945

tion involved a few tricks and additional apparatus in the cab. Perforating
required only the cable and the winch, with the firing devices connected
directly to the winch collector.
In addition to the standard or heavy truck, there was a model for
fields where the depth of the sedimentary rocks did not exceed 4500 feet,
as in Illinois, Kansas, and eastern Oklahoma. Generally similar to the
Mack or International type, though lighter and less powerful, the equip¬
ment was mounted on a Ford chassis.1 Several of these trucks endured
some of the roughest winter operations in Illinois. The terrain difficulties
encountered in the muddy Louden and Salem fields were attested to by
the fact that, for certain months, the trucks remained in the area and
were pulled by caterpillar from well to well; as a result, the gas consump¬
tion of fhese vehicles with little recorded mileage was 1 gallon for each
mile charged to the client. Perforating, which included on-site transporta¬
tion of heavy, loaded guns, utilized a pickup truck to assist the logging
truck.
Needless to say, by the time the war ended, all of the equipment was
in rather sad shape. Nothing had come from France since 1939. In spite of
the shortages, considerable equipment had been made in the United
States because of the crucial need for logging in top-priority petroleum
production. Basic items like photographic recorders, galvanometers, pul-
sators, and photoclinometers had to be drawn from prewar stores. The
American-made truck chassis, engines, and transmissions were also giving
out, because replacements were allocated only in cases of dire need.
In South America, where communication with Paris had been cut off
and contact with the United States was difficult, things were even worse.
Some local procurement was attempted, particularly in Argentina, and
great ingenuity was shown by engineers and assistants in stretching the life
of the equipment, but it was reaching the end.
While the apparatus required a thorough renewal, it was also time to
take a fresh look at a number of pending problems and to sort out the
wartime technological developments that could be applied to logging.
Both the need for renewal and the opening of new avenues warranted fast
and proper action: fast, because the clients who had shown forbearance
and understanding when Schlumberger failures stemmed from wartime

1 The chassis was too short for the winch to be installed in front of the rear axle.

209
The History of a Technique

shortages in men and material would no longer tolerate them; and proper,
because the new course to be charted would determine the direction of
the future for years to come.
At the end of the war the economic climate in the United States was
highly favorable. Notwithstanding temporary or local vicissitudes, the
petroleum industry was clearly off to a new start and new record perfor¬
mances, first in the United States and then all over the world. The time of
shortages was over; all available energy resources had to be expanded to
produce rapidly the means to operate the cars, refrigerators, and toasters
that everybody wanted. For Schlumberger the market was open and the
challenge tremendous.
New available technology and techniques invited bold approaches. In
1939 electronics was limited to radio; now it was all-embracing, from
aeronautics to nuclear industry. Whereas it would have made little sense
to try to accommodate the equivalent of a 1938-model superheterodyne
in a battered sonde, it had become quite conceivable to incorporate in the
sonde circuitry elements like those of a proximity fuse strong enough to
withstand the muzzle acceleration of a shell. In 1939, when Doll con¬
ceived the induction sonde that was to revolutionize resistivity mea¬
surements, it probably was technically unfeasible; but in 1946 he was able
to construct it with the components of the mine detector he had de¬
veloped during the war.
Psychologically, too, the climate was stimulating. Although unpre¬
pared initially, the United States was convinced that the war had been
won by its industrial might, its organization, and its techniques, but, above
all, by the moral and material superiority of the American way of life.
Never had there been more self-confidence, and it was already taken for
granted that landing on the moon would be a mere question of money and
time. At every level the same confidence permeated companies that
wanted to forge ahead and, for that purpose, kept abreast of new technical
and other developments.
For Schlumberger this was nothing new: technical development was
its very life. But close attention was now given to the latest trends in
management and to the favorable effect of the general climate on client
relations. To these new directions the company owed not only its survival,
but also the impetus that was to increase its activity tenfold and deeply
affect petroleum exploration, drilling, and production.

210
After 1943

The administrative reorganization

One of Schlumberger’s remarkable achievements of that time was to


simultaneously and successfully conduct an administrative and a technical
reorganization. Since the former affects the latter, it is worth brief men¬
tion. Here again it was in the United States that the move was launched
before it was carried out in the rest of the world.
The administrative reorganization of 1945 began at the top. In 1943
Pierre Schlumberger, Marcel’s son, became controller of the Schlum-
berger Well Surveying Corporation; in 1944, executive vice-president;
and, in 1946, president. One year earlier, Roger Henquet, whose attrac¬
tive personality made a deep imprint on this period, had been appointed
vice-president and general manager. In 1945 he returned to the United
States after having received commando training in England and been
parachuted into France, where he had worked before D-day with the
French underground. Influenced by his prior life in California, he main¬
tained a true Hollywood style, which, coupled with his love for snap
decisions, made him a flamboyant leader of the rather heterogeneous
team, which was ill at ease in peacetime reconversion.
In the field the reorganization consisted of a regrouping and
standardization of assignments. Aside from the geographic divisions al¬
ready mentioned, some centers or districts had developed that, although
lacking divisional status, would nevertheless report directly to Houston.
All of these independent units were promptly incorporated into the
existing operational system. For the first time the responsibility and
authority of division heads and district chiefs were clearly defined and
standardized. In 1948 increased growth mandated a more efficient man¬
agement structure, and a new reorganization took place that consolidated
the existing divisions into regional units called areas.
In 1945 what had previously been the Houston exploitation unit was
enlarged into the Field Operations Department with more specific
functions. Its first leader, Henquet, was replaced in 1949 by a vice-
president who, supported by a technical staff, was fully responsible for
operations. The area heads reported to him, and only he issued instruc-
dons to the field crews. This streamlining was much appreciated by
everyone because it greatly facilitated relations with the Houston head¬
quarters.

211
The History of a Technique

The Engineering and Research Department was diverisfied in 1947,


with each function operating separately. The Research Center in
Ridgefield, Connecticut, was established in the same year under Doll,
while engineering headquarters remained in Houston. Ridgefield was
now in charge of long-range research on instrumentation as well as in¬
terpretation; its laboratories became operational in 1948. The staff en¬
gaged in engineering and research grew between 1944 and 1949 from 22
to 83; it was to reach 224 in 1955. During the same period the Equipment
Department experienced similar growth in personnel: 82, 260, and 354.
With a growth rate of such magnitude, a Personnel Department was
required, and the new unit was created in 1948. Under the leadership of
Ame Vennema, it performed outstandingly. One of its major contribu¬
tions was the publishing of a management manual in which everybody’s
terms of reference, responsibility, and authority (from the president to
the sweeper) were properly defined. The department maintained an em¬
ployee benefits policy that had existed since the inception of the Com¬
pany under the direction of Conrad and Marcel. Employees’ profit-sharing
had begun in 1937, and in 1945 a profit-sharing trust was created in which
the shares of every employee were invested. Its insurance, holiday, vaca¬
tion, and pension system, as well as its profit-sharing trust, have put the
Company far ahead of many American firms in terms of total employee
benefits. In 1948 the Personnel Department organized a training program
for engineers and, later, a series of refresher courses.
To round out the picture, mention should be made of the newly
created Sales Department, a rather odd name since, in addition to its
responsibilities concerning high-level client relations, price lists, and sales
promotion, the department was also in charge of the technique of log
interpretation. In close cooperation with the Ridgefield laboratory, it
disseminated all available information on the subject as well as on the
many and diverse new tools that came on the market.
Thus, between 1945 and 1950, the Schlumberger Well Surveying
Corporation established in the United States a structure well suited to its
volume of activities and fully equipped to handle new developments.
Testimonies to its importance are the facilities in Ridgefield (1948, Fig.
61) and Houston (1953, Fig. 62), which have served as architectural and
functional models for offices, laboratories, and manufacturing plants. Yet
the real strength of S.W.S.C. was less obvious: it rested on the many small
centers scattered over the United States (54 in 1948, 119 by 1956), which

212
After 1945

Figure 61. Ridgefield, Connecticut, Research Center.

conducted 5000 monthly operations in 1948, 6000 in 1950, 8000 in


1951, and 10,000 in 1953.
The reorganization had multiple effects. Originally, engineers were
assigned to various departments according to need rather than back¬
ground and experience. As the years passed, however, the demand for
greater specialization resulted in an engineer or a physicist being recruited
on the basis of his suitability to a particular post. The romantic character
of the Schlumberger engineer was fast becoming obsolete; no longer was
he a jack-of-all-trades who talked to the client, interpreted the mea¬
surements, kept the payroll, salved the customs inspectors, and did the
cooking. Efficiency gained, and eventually this “Americanization” spread,
first to South America and then to the rest of the world. “Prospectors” of
an earlier day were to be found only in Paris, at 42 rue Saint-Dominique;
lend an ear, and these veterans of “La Pros” were quick to tell you how
things were “in the good old days.”
The new organization also had psychological advantages. As the field
engineers became more aware of their responsibilities and prerogatives

213
The History of a Technique

Figure 62. Headquarters of Schlumberger Wells Surveying Corporation in Houston.

and were informed monthly of the financial results of their divisions or


districts, they became increasingly aggressive promoters, though never
“peddling their wares” at the cost of their technical performance. In fact,
their frequent contact with clients gave them greater insight into the
latters’ problems and needs, and this, in turn, enabled them to clarify
these viewpoints to their colleagues in research or engineering. Addition¬
ally, the engineers were given new opportunities to voice their opinions,
either privately or at frequently scheduled meetings. As always, the
observations coming from the field were given the greatest attention and
consideration by Doll and Marcel.
The reorganization of Schlumberger Companies gradually expanded
to South America and then throughout the world. Under the limitations
imposed by geographical separation, this task was undertaken by Jean de
Menil, a man of unbending willpower and tenacity concealed beneath an
exterior of tact, human respect, and personal charm. In the process of
effecting necessary changes, he had to break with tradition and put an end
to privileges, even if it sometimes meant resorting to sheer strength.

214
After 1945

Thanks to his endeavors, by 1952 South America—scattered as the


various units were, it nevertheless boasted a strong esprit de corps—had
been reorganized in a manner similar to that of the United States.
The geopolitical realities of the eastern hemisphere hardly lent them¬
selves to a divisional regrouping. The road from Nigeria to Gabon went
through London and Paris; and although some early consolidations could
be achieved in the Middle East, reorganization often had to wait for better
communications or a concentration of drilling operations. Only after
1956-1957 did Guy Baboin succeed in establishing an administrative
structure modeled after the one in the western hemisphere.
Another equally important development, also initiated by de Menil,
gave a worldwide uniform status to field engineers: direct compensation,
fringe benefits, and standardized living conditions wherever possible.
There was now greater flexibility in worldwide staff transfer, which pro¬
moted a stimulating mixture of individuals and a fruitful exchange of
technical skills.

The Series 500 and 700 Trucks

In this discussion of technical developments, special mention must be


made of a new truck2 that first appeared in the field in 1946. Not only was
it a superb working tool fitted with many new instruments, but also the
future of the Company was almost irrevocably committed by the original¬
ity of its design features.
It had become clear that standard logging could lead to many other
activities. The “auxiliary” operations (perforation, sidewall sampling,
temperature and dipmeter surveys) were selling well, and the divisions
and districts could see from bookkeeping records that these operations
were financially very attractive since they used existing equipment, per¬
sonnel, and organization. Electrical logging was still considered the
mainstay of the Company, but there was a growing awareness that petro¬
leum exploration had opened new doors in measurement techniques. As
early as 1945, it was apparent that the drilling of every oil well in the
United States and soon elsewhere would involve conventional logging,

2 “Truck” refers not only to the vehicle, but also all the operational equipment it carried:
cable, winch, recorder, power supply, and so on.

215
The History of a Technique

but whatever the ensuing expansion of the market, overall drilling activity
would remain the determining factor. Therefore the only prospect of
expansion lay in offering a series of measurements in each well that would
provide a coherent data system for all the variables of the problem. To
think in terms of electrical logging only, and not consider these additional
measuring systems, would deter progress. This was the basis on which the
new truck was designed: it had to be suitable for a whole new range of
operations other than those of the past. Although the concept focused
mainly on electrical outfitting, it affected the entire project.
Thus, before the Series 500 truck had even reached the drawing
board, it had a diverse group of progenitors: all the field engineers
received a questionnaire seeking their opinions on a broad range of
choices.3 The final choice was the Mack 18 ton chassis, with a forward cab
and tandem dual-wheel rear axles. It provided ample room for the record¬
ing cab, the winch, and its accessories. The eight driving wheels distrib¬
uted the load over an area wide enough to minimize the risk of getting
stuck in the mud, a constant nightmare of oil people. The profile of the
body was reminiscent of the experimental California trucks in its em¬
phasis on elegance, although its real beauty lay in the way it was perfectly
designed for the work expected (Fig. 63). “Strictly functional” would be
the expression used today to describe it.
The winch was an integral part of the truck. It was operated from the
recording cab through advanced control systems (air brake, air throttle,
air clutch), and a dashboard that, in addition to giving the usual informa¬
tion on the engine performance, also indicated the velocity and tension of
the cable.
Film processing had earlier been limited to the equipment used by
the photographers of yesteryear: a box and black cotton sleeves. Now it
took place in a closed nook with a black curtain, a red lamp, a timer, and a
stainless steel sink. Instead of operating blindly, there was a film of
commercial standards, meeting the growing demand for a field record of
high quality. The photographic recorder was new. Designed in Houston,
it was the mechanical and optical heir to its predecessors. It had retained
the same film drive and the same optical circuits for the production of

3 To the question “Should the recording cab have a seat for the client?” one field engineer,
probably harassed by trifling queries of his own client, replied, tongue in cheek, “Yes,
ejectable.”

216
After 1945

Figure 63■ Series 500 truck.

light spots, but there were now nine galvanometers instead of three, and
two films instead of one; also, the film width had been increased.
Why two films? The explanation lies in the basic functions of the log:
well correlation and formation analysis, for which different depth scales
are desirable. For correlations a smaller scale, usually 1: 1000 or 1: 500
(in English-speaking countries 1 or 2 inches per 100 feet, i.e., 1 : 1200 or
1 : 600), showing a rather long section in a single look is preferable. De¬
tailed formation study requires a larger scale, 1 : 200 (or 5 inches per
100 feet, i.e., 1:240).4 Therefore the recording of two logs, one over the
entire length of the open hole for correlations, and the other at a larger
scale, limited to sections deemed to be of interest, was in order. With the
hand recorder two simultaneous logs could be recorded: only two pencils,
actuated by the potentiometer’s crank, that traced the curves on two rolls
of grid paper revolving at different speeds were needed. To obtain two

4 A larger scale, 1 : 50, is sometimes used for certain operations (dipmeter, Microlog,
Microlaterolog), as well as for certain detailed correlation studies, provided that the
vertical resolution of the device allows it.

217
The History of a Technique

logs at different scales with the single-film optical recorder, two runs were
required; this entailed loss of time and greater operational hazards, since
the sonde had to be lowered again to the bottom of the hole. The difficult
problem of impressing two films, unwinding at different speeds, was
solved by splitting the light beam reflected by the mirror of the gal¬
vanometer, in order to obtain a spot on each film. The time saved justified
this complication when the operating costs of drilling rigs rose to the
point of making standby time prohibitively expensive.
Another substantial improvement was the vibrationproof Picard gal¬
vanometer (see p. 195-196). This galvanometer was compact enough
for nine of them—more than necessary for the techniques of the
time—to be housed side by side in the new R9G (“9 Galvanometer
Recorder”). As early as 1947 this equipment gave Schlumberger a system
which proved so satisfactory that, with a few modifications over the years,
it has remained basically unchanged to this day. This is a remarkable
achievement, considering that it was impossible at that time to foresee the
range of measurements that would eventually supplement electrical log¬
ging.
The final form of the technique was determined by development
work in Paris and the United States. Although the simultaneous recording
of two depth scales was time saving, production of the three resistivity
logs in a single run would have been a still greater asset. Two choices were
open: multiple frequencies, or the sequential or chronological system, on
which Bricaud was putting the final touches (see p. 197). The advantage of
the former alternative was the possible use of a single-conductor cable,
but it required much electronics in the sonde, a nearly damning condition.
In light of today’s sondes, say a dipmeter and its transistorized cartridge,
such reluctance toward bottom-hole electronics may seem strange. In
1945 it probably would have been considered only as a last resort, and
certainly not as a substitute for a mandrel, ringed in places by electrodes
directly connected to the cable conductors. Therefore, despite the prob¬
lems of a six-conductor cable, the chronological system prevailed because
it rested on proven techniques. The various sequential connections were
made by a new revolving switch, more complicated and sophisticated than
the old pulsator, yet of the same basic, rotating segment-collector type.
Designed and built in Paris in record time, its near-capacity operation was
marginal, and hence it was the least dependable unit in the chronological

218
After 1945

system. It was replaced in 1957 by a cam pulsator with opening and


closing circuit breakers (points), which gave a much better performance.5 6
To those who used it, the memory of the chronological system is
almost as nostalgic as that of the great sailing ships or steam locomotives.
Indeed, this was the final and most efficient form of electrical resistivity
logging, until it disappeared1' and was replaced by “focused” mea¬
surements—a great step forward for the theoretical and practical in¬
terpretation of results. Its decline began in 1955 and gained momentum
in I960, and by 1965 the chronological system had been abandoned,
except for a few specific correlation problems mainly in shallow, geologi¬
cal reconnaissance coreholes. The merits of conventional resistivity log¬
ging went far beyond the success of the Schlumberger system: it opened
the way to the whole body of physical parameter measurements in
boreholes, that is, to all kinds of logging. Partially based on such mea¬
surements, a new science was born: petrophysics, an essential tool for the
efficient production and effective conservation of oil deposits. Although
accurate statistics are not available, it is safe to say that conventional
resistivity measurements have resulted in the discovery of more oil and
gas than any other method of borehole geophysics.
With regard to the previous discussion of the Series 500 truck, it can
be said that, once the main choices had been made, design and technology
offered many solutions. One of them, a modular arrangement of the
controls, had been proposed in the United States; in France, another one
was based on their unitization. The differences may have been sympto¬
matic of national idiosyncrasies; but French and American engineers had
worked hand in hand on the American solution, and it was only postwar
circumstances that had prevented American engineers from sharing in the
French proposal. Technically, both were valid; however, a fact not fully
realized at the time was that the improvements of the French solution
followed a more traditional line. They presented a system conceived for
conventional resistivity measurements and possibly adaptable to other
applications. On the other hand, the American version led to a system
broad enough to accommodate a whole family of measurements, conven-

5 It was machined with a higher accuracy than the camshaft of a racing-car engine.
6 At least in the western world. Conventional resistivity measurements with lateral sondes
of various radii of investigation were still in use in the U.S.S.R. in 1970.

219
The History of a Technique

tional resistivity being only one of them. There is hardly any doubt that
the American solution was the better and the more progressive, since it
added the flexibility needed for auxiliary operations. The modular con¬
cept of instrumentation probably resulted from a closer contact between
engineering and field, and it certainly gave Schlumberger the opportunity
to offer many new services between 1947 and 1955. The idea was simple
enough: for each parameter (resistivity, radioactivity, temperature, dip,
etc.) a sonde and a control panel were designed to suit the case. Data
gathering and transmission utilized the procedure best adapted to each
kind of measurement, without the concern for standardization which,
until then, had led to the use of pulsated current for all operations. Only
the cable, the recorder, and the generating unit7 were used in all opera¬
tions. With such a system, enabling a crew to perform a new service
without any other adaptation was merely a matter of providing it with the
appropriate sonde and control panel.8
All this may seem .obvious in retrospect, yet it was less so at the time
when a decision was required as to which direction to follow. Marcel was a
lover of fine mechanics, and many of his engineers had fond memories of
equipment performing at capacity limits. For this reason they did not
readily accept the idea of a truck offering a surplus capacity in every respect.
But times for individual feats were gone, and modern trends were calling.
The spokesman for the American system was Doll, who was able to lay
aside his own preference for electronic solutions in favor of dependability.
He retained the Paris proposals with their indispensable items like the
chronological system, the vibrationproof galvanometer, and the pulsator,
but rejected any outdated instrumentation biased toward resistivity mea¬
surements. The same discrimination prevailed for all the American propos¬
als; with reflection, explanations, persuasions, and, when needed, technical
and moral authority, he succeeded in synthetizing his choices into an
outstanding solution.
As previously mentioned, the chronological system required six con-

A small 2 kilowatt, 110 volt, 60 cycle alternating current unit providing power to actuate
the various individual measuring devices (sensors) and to mechanically control the
borehole apparatus.
8 This is still being done, but interest in the method has lessened since tool combinations
became necessary to perform in a single run an increasing number of borehole operations.
Several control panels must then be connected in parallel, a requirement that entails
technical and human problems.

220
After 1945

ductors and hence an armored cable. Anyhow, the textile braid had
outlived its usefulness. Granted, the cable had changed over the years,
from three to four conductors, from rubber insulation to hy-
drocarbonproof neoprene, and from cotton to rayon braid; but its large
diameter required huge winches, and its coating wore off quickly. With a
4 ton tensile strength, this cable could tolerate abuse without permanent
damage. The steel wire conductors took all the stress and practically never
broke, so that the only electrical faults came from insulation cuts or
bruises. Such leaks were detected by an ohmmeter, located by the guard
ring method, and then cold-repaired with adhesive rubber tape. As the
diameter of the cable was not critical, bulging patches mattered little.
Utilized in various ways by competitors and for perforating by
Schlumberger, there was a cable with a load-carrying armor and a central
copper conductor, but it was only a monoconductor. There are complex
problems in the engineering and manufacturing of a hexacable. The
strands must be calculated in such a way that the various components of
the cable—conductors and internal and external armors—have compat¬
ible elongations, to prevent possible disasters: depending on the case, the
conductors may break, or, if they become permanently elongated, the
release of the stress of the armor may produce a loop in the conductor,
piercing the insulation. But the most spectacular effect comes from an
elasticity difference between armors. Under certain conditions and after
excessive stress, the internal armor reverts to its original length, whereas
the external one remains permanently elongated. All the armor wires then
take the same shape and remain about equidistant, so that the whole thing
bulges symmetrically into what in the field is called a “bird cage.” This is a
major accident: the wires of the external armor must be cut and tied lest
they untwist, until a final workshop repair puts the cable back into shape,
because, as discovered promptly by the operators in the field, a bulge
interferes with the perfect winding on the winch, an absolute requirement
for the proper handling of armored cable.
Much time and experience are required to fix conductor breaks and
electrical leaks. Concerning leaks, the advantage of the armored over the
textile-braided cable is that leaks in the former are so obvious and enor¬
mous that no measurement is possible. This is, then, a clear “all or
nothing” situation, greatly preferred when flawless measurements are
sought.
The cable is such a basic component that continuous engineering

221
The History of a Technique

studies are needed to improve its specifications, quality, and performance.


In accordance with a suggestion of the Paris group (F. Barreteau) in I960,
a seventh conductor was added at the center. Simple as the idea seemed,
for a long time it had been deemed impossible. The seventh conductor
was straight; it became the center core around which the other six conduc¬
tors were helically wound, and unless adequate mechanical precautions
had been taken, would have broken when overstrained. Another substan¬
tial contribution at that time was the reduction of the cable diameter from
13.2 to 11.6 millimeters without affecting its tensile strength. This made
it possible to wind a longer cable, designed for the general trend to deeper
drilling, on the same winch with no weight increase. Additionally, it was
easier to lower the cable through a blowout preventer into a well
under pressure: since the force opposing the motion of the cable is
proportional to the square of its diameter, the slightest decrease in the
latter makes an appreciable difference.
The Series 500 truck had been designed and equipped to handle an
armored cable. Large-diameter sheaves prevented sharp bends between
winch and hole. The standardized location of these sheaves on the derrick
made it possible to measure9 the cable tension accurately. In addition, by
knowing the tension, the elastic elongation of the cable and hence the
exact depth of the sonde could be calculated.10
Another accessory, a “quick-mounting" head for the connection of
the sonde with the cable, played a primary role in outfitting the Series 500
truck. As long as electrical logging and an occasional auxiliary operation
were performed, a semipermanent connection was adequate. Using a

9 A conventional strain gauge, requiring an electronic measuring device, was used.


10 This important depth problem is a delicate one when the accuracy sought is 1:10,000.
Metering the number of turns of a sheave, however well calibrated, is not an adequate
solution. Neither was the tangential measuring wheel, conceived and made by Marcel to
give the scale of the logs by unwinding a length of film proportionate to the travel of the
sonde in the hole. Therefore measurement of the depth is made with a cable that has been
calibrated with a surveyor's chain every 50 meters (or 100 feet) under a known and
constant tension. The exact depth is obtained by counting the number of marks passed by
from the origin (usually the rotary table), and making the correction for elastic elongation.
Between two marks the measurements are interpolated by using the tangential wheel. On
the textile-braided cable the marks were rings of friction tape. This was not possible with
the steel cable, but it was discovered that the steel of the armor could be marked
magnetically either by using a coil or by mere contact with a good horseshoe magnet; the
mark is then detected when passing before or through a coil.

222
After 1945

combination of cable connectors, rubber tubes, and string bindings, a


trained team required hardly an hour to substitute a thermometer for the
electrical logging sonde. A perforating operation had to be speedy, how¬
ever, because three or four runs were needed if there were many bullets
to shoot. For example, if there were 80 holes to be perforated, four guns
with 24 bullets each had to be run in and pulled out in succession. This
had led, as early as 1936, to the design of a rapid coupling for the
assembly of the cable with the gun or any other instrument. Once the
electrical connections were made, tightness was achieved by a rubber or
neoprene gasket compressed by a double-threaded collar, the “differen¬
tial sleeve. The six-conductor cable not only involved more electrical
contacts,^ but also required insulation between the metal head and the
armor of the cable. The new “cable-head,” designed and manufactured in
Houston,11 was a key item for versatile equipment.
The problems just described may seem trifling and unglamorous.
Yet, with a technique like electrical logging, founded on basic ideas and
discoveries, the development of the most remarkable inventions could be
delayed by inadequate attention to practical know-how. Without doubt
the technical lead enjoyed by Schlumberger owed much to the patient
craftsmanship which produced instruments that endured higher tempera¬
tures, withstood higher pressures, and insured better insulation. Nobody
was more aware of this than Marcel. All one had to do was to see him in
the field, carefully observing every detail, every strategem, every failure.
Not only did he put his experience into practice, but also he brought his
convictions home to his collaborators (Fig. 64).
The outstanding performance of the Series 500 truck, as designed
and realized mainly by Maurice Lebourg, Roger Legeron, Pierre Dubost,
and their team, has never been questioned. When commissioned in 1947,
it gave engineers the field unit they needed to match their competitors,
who, untii then, had had the advantage of more modern-looking, though
less dependable, equipment than Schlumberger’s. The truck was sturdy,
well designed, and easy to operate, with a well-laid-out cab (Fig. 65). From
the start its chronological system gave technically flawless logs, whereas it
took several years for the competition to achieve a reasonably dependable
electronic system.

11 The first model, the seven-pin head, had seven contacts and was designed in 1947 by
Maurice Lebourg. A second model, designed by Charles Senouillet in 1953, had ten pins.

223
The History of a Technique

Figure 64■ Marcel Schlumberger in a cab on a truck.

All of the mechanical solutions introduced with the Series 500 truck
have remained in use with only slight modification. A lighter truck with a
shorter cable, mounted on an International chassis, was commissioned in
1948 as a Series 700 and was produced in large numbers. Today most
trucks are International models with a forward-tilt cab, weighing 20 tons
and carrying 25,000 feet of cable. For very heavy duty service, there is a
special 25 ton model with three driving axles, one front and two rear.12
The same concern for dependability had inspired the design of the
“Offshore Skid Unit, Type C” (O.S.U.-C), used on offshore drilling plat-

12 Instead of dual wheels on the rear axles, this truck has single wheels with tires of very
large cross section (1600 X 20), which provide much better traction on sand.

224
After 1945

Figure 65. Cab with photographic recorder.

forms or on Glomar-type ships. The winch, in accordance with maritime


fire regulations, is diesel driven through a hydraulic transmission. Its
flexibility and operational precision are excellent. The recording cab is
roomy and well fitted. The unit is suited for offshore conditions, where
the operating cost of the platform is so high that no breakdown, no delay,
can be tolerated (Fig. 66).
When chronological logging, which required a six-conductor cable
was abandoned, Schlumberger, with the help of modern electronics,
could have turned to a more conventional, more easily made, and sturdier
single-conductor cable. There were good reasons, however, for retaining
the hexa- or heptacable.
First, the chronological system was phased out slowly and gradually,

225
The History of a Technique

Figure 66. O.S.U.-C unit for offshore wells.

rather than simultaneously in all Schlumberger field operations. Second,


after a somewhat difficult start, the cable had become fairly dependable,
and there was no pressing reason to change. And although the six conduc¬
tors (even for measurements other than conventional resistivity) were not
indispensable, they had the great advantage of requiring a minimum of
downhole electronic equipment in the sondes or cartridges, the bulk of
complex equipment remaining above ground. Today the argument may
appear to be slightly antiquated, but the question had to be considered in
the light of tube electronics, where even miniaturized tubes raised serious
technical problems of space, temperature, energy dissipation, and shock-
and vibrationproofing. Dependability was also essential, since trou¬
ble shooting on surface equipment was much easier and speedier than
pulling the sonde out and lowering it again after repairs. Another argu¬
ment against placing costly electronic devices in the hole was that, if
failure of a gasket provided the slightest hydraulic leak, the inside of the
cartridge would be exposed to very high pressures (thousands of pounds
per square inch) and totally destroyed.

226
After 1943

It is interesting to note that recent technical developments have


reawakened an interest in a cable with six or seven conductors. The trend
toward recording as many parameters as possible in a single run requires
increasingly higher frequencies. Yet, beyond 20 kilocycles, any frequency
that is transmitted through a cable, several kilometers long, armored with
magnetic steel, is severely and abruptly limited by losses. However, it is
possible to successfully exceed this limit by using the appropriate combi¬
nation of conductors. Thus, in the frequency modulation technique of the
high-resolution dipmeter, signals of 100 to 175 kilohertz can be transmit¬
ted by using hexacable or heptacable.

The postwar progress in interpretation

The impact of Schlumberger’s technical development on moderniz¬


ing over a few years the measuring methods in boreholes was a logical
sequel to the progress made in interpreting the results.13
Giant strides had been made by using electrical measurements to
assess the characteristics and contents of rocks since the first, strictly
qualitative results were obtained on the Gulf Coast, and more systematic
studies had been completed in the U.S.S.R. Mention was made earlier of
G. E. Archie’s remarkable research, well documented by his long experi¬
ence and numerous laboratory measurements (see p. 201). His study of
1942 established a firm basis for quantitative interpretation. Such work
was in line with the thinking that prevailed at Schlumberger; its merit lay
not only in its clear presentation, wherein each parameter was assigned to
its proper part, but also in its adherence to the standards required for
setting such petrophysical “laws.” Whatever the a posteriori attempts to
justify them theoretically, they remain inherently statistical in their na¬
ture.
A closer analysis of the interpretation process covers the background
of the various contributions in an effort to facilitate an understanding of
how the concepts have evolved. Two main phases must be considered.

13 Only much later (1958), pressed by clients and seismologists, did Schlumberger develop
a technique that appeared to contribute nothing to interpretation: the “sonic” logs, which
measured the velocity and attenuation of acoustic waves in geological formations. In time,
however, these measurements, too, became part of the interpretation methods and played
an important, though initially unforeseen, role.

227
The History of a Technique

First, the exact values of the physical parameters must be determined, for
example, the true resistivity of a formation, based on its apparent resistiv¬
ity (see p. 127), and corrected wherever necessary. The second phase,
based on the corrected physical parameters, aims at determining the
characteristics of potentially productive formations, with emphasis on
economic factors such as porosity, oil percentage, and oil mobility.
Since the first phase relies on the behavior of the tools utilized, it
would normally be handled by the service company taking the mea¬
surements. The second phase, however, could be dealt with by the pro¬
ducer (client) on the basis of his own experience. This did not mean that
the data gathered by Schlumberger’s clients were ignored in the Com¬
pany’s research and engineering programs, but Schlumberger believed
that its competence and responsibility should be limited to the correct
determination of the physical parameters. At least at the top level, this
concept prevailed for many years. Only the features of each instrument
and the corrections to be made on each log were published. The separa¬
tion between the two phases made things difficult for the field engineers,
who worked closely with the oil people and could hardly stop halfway
through the interpretation. A quick review of some important technical
publications shows how thinking in this regard changed during this
period.
In September 1945 H. Guyod, a former Schlumberger engineer,
wrote a series of articles on electrical log interpretation in Oil Weekly
(which later became World 0//).14 His was the first in-depth coverage of
this subject, wherein he described the various resistivity sondes:
monoelectrode, normal, and lateral, giving the particulars of each instru¬
ment, and indicating when and how closely the true resistivity was repre¬
sented by the readings.
These articles were widely read in the petroleum industry and came
as a surprise to Schlumberger’s field engineers. They firmly believed that
this information, which was nothing new to them, should be available to
the industry. Their increasingly close collaboration with their clients
created an obligation to provide explanations so as not to seem evasive. It
was no longer deemed fair to supply logs without commenting on their
peculiarities, anomalies, or asymmetries. Originally, there may have been

14 Oil Weekly, December 3, 10, 17, and 24, 1945.

228
After 1945

some reason for keeping a good part of the technology and even the
technique of resistivity measurements confidential; but since the competi¬
tion was now fully cognizant of those alleged secrets, there was no longer
any point to such reticence. What really stung the field engineers was that
the disclosure had come from outside the Schlumberger group. Some¬
thing had to be done, and in 1946 “departure curves”15 were published to
compute the true resistivity subject to certain conditions: the assumption
that a formation was infinitely thick16 and that the electrode spacings, the
borehole diameter, the invaded-zone diameter and resistivity, and the
mud resistivity were known.
Even before this publication, departure curves were widely known
among the interpretation experts who were beginning to staff the oil
companies, although their application, at least in the United States, was
limited. They were difficult to use and often unreliable when major
corrections were required. Therefore, instead of serving to compute true
resistivity, the curves were used to indicate cases in which the apparent
resistivity came close enough to be applied as a substitute. Obviously, it
was necessary to develop systems and tools that either would measure the
true resistivity directly or would provide smaller and more convenient
corrections than did the departure curves.
In February 1948 Doll presented to the A.I.M.E. Congress a paper
entitled “The S.P. Log: Theoretical Analysis and Principles of Interpreta¬
tion.”17 This was a basic study, not so much about the nature of the S.P.
phenomenon, as on its current and voltage distribution. Generated by the
electromotive forces arising from the contacts between different media,
these currents propagate in both the permeable formation and its shoul¬
ders, and close through the mud column. Since the S.P. curve is a
continuous record of the potential of an electrode moving in the
borehole, the deflection between sand and shale can measure, not all of
the electromotive forces involved, but only those reflected by the ohmic
drops in the mud. The flow of current is schematically represented by the

15 Originally called “lateral sounding curves”; see p. 163.


16 Departure curves for formations of finite thickness, based on approximate calculations,
were published in 1949. This problem was more complex, since other parameters had to
be considered: the resistivity of the shoulders (themselves assumed infinitely thick) and
the thickness of the formation studied.
17 See Journal of Petroleum Technology, September 1948.

229
The History of a Technique

lines drawn on Figure 67, which clearly indicates how the relative resis¬
tance value of each circuit component—permeable bed, shale, mud
column—influences not only the S.P. value but the shape of the log as
well. The S.P. value approximates the total electromotive forces when the
resistance of the circuit section inside the mud is proportionally high
compared to that in the bed and its shoulders, as is always the case in thick

Shale

Sand

Shale

Shale

Sand

Shale

Figure 67. S.P. diagram current flow. <From H. G. Doll, “The S.P. Log: Theoretical Analysis
and Principles of Interpretation,” Journal of Petroleum Technology, September 1948 (Cour¬
tesy SPE of AIME).

230
After 1945

formations.18 As to shape, study of the tridimensional current flow shows


that the sharpness of the deflections decreases as shoulder or formation
resistivity increases.
To measure, under any circumstances, the total electromotive forces
with electrodes confined to the mud column, an open-circuit operation
would be required. As this is impossible, Doll defined a “static S.P.” as the
difference of potential that would prevail between two points, one of
them either above or below the permeable bed, and the other facing this
bed, if insulating plugs could be placed at its top and bottom (Fig. 67).
This “static S.P.’’ remained an abstraction, but made it possible to
conceive S.P. readings that were free of any interference from the current
distribution inside the borehole and its immediate vicinity. As a result the
readings could be standardized and mutually comparable. Going further
into this distribution study, Doll presented a satisfactory explanation for
the heretofore totally disconcerting behavior of the S.P. in high-resistivity
limestone systems: it takes only one or two thin, permeable layers in a
large and compact limestone series for the S.P. deflection to be stretched
out and slowly peak over its whole thickness (Fig. 68). This analysis
provided little hope that some accuracy could be achieved in the delinea¬
tion of permeable intervals, which were revealed, at best, by slight curva¬
ture variations of the log. What the analysis did uncover was the valuable
fact that, since the conventional S.P. could not be relied upon to locate
these intervals, some other way had to be found. In the meantime exces¬
sive errors in interpretation would be avoided. As a matter of fact, there
was a nearly overnight change in the perspective which engineers operat¬
ing in predominantly limestone fields had of S.P.
Great attention had also been given to the study of shales, especially
thin layers, laminations, and, at the limit, shaly sands. Although the
behavior of the S.P. in such formations, long acknowledged as important
potential oil sources, was less puzzling than in limestones, log interpreta¬
tion was still difficult.
In July 1950 Doll published another paper, actually an offshoot of
the earlier one, entitled “The S.P. Log in Shaly Sands,”19 in which he
discussed the relationship between shale content, resistivity, and the S.P.

18 With a good approximation it can be stated that the ohmic drop in the mud represents
the total electromotive forces in a sand 10 feet thick with a resistivity not exceeding 10
times that of the mud.
19 In Journal of Petroleum Technology, Vol. 2, No. 7 (July 1950).

231
The History of a Technique

tilt 7/
1 1 1 1 \ 1 1 1
1 1 1 X 1 1
1 1 1 1
1 Ï 1 1 VI 1
i i n l 1 \ 1 1
i i i i i i ti r~
1 t mr i

7
7
'//////s

\
\
\

\
\
\
, 1 , 1 ,\ ,
1 1 ! 7 1 1 \ 1
1 1 1 I I IV

1 1 III /
1 1 1 1 /
I T T1 1 Vj
l l l ~1 1 \S 1
r 1 i r i »
i r i ii »
i i i >i i
i i i l >i ii
i i 1 1 1
777777 7f/777//7z

Theoretical S.P. Circulation of


S.P. currents

Figure 68. S.P. in limestones. (From H. G. Doll. “The S.P. Log: Theoretical Analysis and
Principles of Interpretation, Journal of Petroleum Technology, September 1948) (Courtesy
SPE of AIME).

reduction in shaly sands as compared to the “clean sand”20 value. The


importance of this study lay in its applicability to the shaly sands of many
large productive basins (Texas, Louisiana, Venezuela, Nigeria). Both pa¬
pers were outstanding in their presentation, clarity, and discussion of the
subject matter.
At the time there was great interest in the importance of the static
S.P. within the general framework of quantitative interpretation. Al¬
though Doll had precise ideas of his own on the matter, he preferred to
keep them to himself, thus remaining faithful to the position that Schlum-
berger should concentrate its efforts on providing only the exact value of
the physical parameter. As described earlier, Tixier demonstrated in 1944
that, at least in the Rockies, the electrochemical component of S.P. was
dominant. From then on, it became possible to evaluate the resistivity

20 The name given to sands without any clay content.

232
After 1945

(.Rw) of the formation water, a fundamental datum that in turn led to the
calculation of R0, the resistivity at 100 percent water saturation of the
porous formation, provided its porosity is known. Any true resistivity
value above R0 indicated hydrocarbon saturation, in a proportion that
could be deduced from the ratio between true resistivity and R0, at least in
a clean formation.
It was a research physicist for Gulf Oil, M. R. J. Wyllie, who gave a
general formula for the electrochemical S.P. component in October 1948
in “A Quantitative Analysis of the Electrochemical Component of the S.P.
Curve. '21 This study, based on theoretical considerations and experimen¬
tal data, established, as a function of temperature, the relation between
the static S.P. value and the logarithm of the ratio between the resistivity
of the mud filtrate and the resistivity of the formation water (Rjnf/Rw). This
formula became even more valuable when experience showed that, as in
the Rockies, the electrofiltration component was nearly always negligible.
Here was a general method giving Rw and R{).
Between 1949 and 1953 Doll made several major contributions to
technical literature wherein he explained the principles of his newly
designed instruments, which were already on the market. These studies
included discussions on the quality of measurements, their expected
degrees of approximation according to operating conditions, and the main
corrections to be made. The tools themselves will be described further
on; the purpose here is to give a chronology of the papers published to
show how fast the logging industry was progressing. The titles of four of
Doll’s studies and the dates on which they appeared are as follows:22

June 1949 “Introduction to Induction Logging and Application to


Logging of Wells Drilled with Oil-Base Muds”

February 1950 “The Microlog: A New Electrical Logging Method for


Detailed Determination of Permeable Beds”

November 1951 “The Laterolog: A New Resistivity Logging Method


with Electrodes Using an Automatic Focusing System”

January 1953 “The Microlaterolog”

21 See Journal of Petroleum Technology, January 1949-


22 The four studies listed were published in Journal of Petroleum Technology.

233
The History of a Technique

For Schlumberger field engineers, however, living close to the prob¬


lems of production, it was no longer possible to restrict themselves to
providing precise values in ohms or millivolts. At last, in 1949, Tixier was
given the green light to publish two articles, after he had disclosed their
substance months before:

“Evaluation of Permeability from Electric Log Resistivity Gradient

and, more importantly,

“Electric Log Analysis in the Rocky Mountains”23

Finally, in 1954, the Ridgefield Research Center, a pillar of official


thinking, published under the names of H. G. Doll and M. Martin an
article entitled:

“How to Use Electric Log Data to Determine Maximum Producible Oil


Index in a Formation”24

The title was self-explanatory; the subject matter was no longer physical
parameters, but barrels of oil.
In the same year André Poupon in Ridgefield, Milton Loy in Califor¬
nia, and Maurice Tixier in Houston jointly published an important study:

“A Contribution to Electrical Log Interpretation in Shaly Sands”25

Aside from the description of new instruments, Schlumberger con¬


tributions to specialized literature between 1950 and 1957 may appear
modest, but the role the Company played in the dissemination of informa¬
tion was, on the contrary, far from unimportant. This dissemination
developed in stages. Training courses for Schlumberger engineers were
first organized in the United States. After 1948 confirmation to a post
required prior technical tests with the emphasis on log interpretation.
Although initially this indoctrination took place in the various centers,

23 Both articles appeared in Oil and Gas Journal, June 16 and 23, 1949.
24 Oil and Gas Journal, July 5, 1954.
25 Journal of Petroleum Technology, June 1954.
Technology, June 1954.

234
After 1945

where a senior engineer was assigned to train the beginner, the need for a
more uniform approach soon became evident, and a training school was
opened in Houston under the direction of André Allégret. By 1956
Surenco in South America and SPE-Overseas in Europe had their own
schools. Meanwhile, in 1950, 2 week refresher courses were conducted in
Houston; a third week, entirely devoted to interpretation, permitted each
Schlumberger engineer to invite one of his clients. Success was such that
the attendance soon outgrew the facilities, and separate courses had to be
offered to clients only. Even this soon proved inadequate, and in-service
training had to be organized in each division, where the Schlumberger
staff, assisted by specialists from Houston, held full-fledged seminars.
Such an educational approach had a wide influence.26 It provided an
opportunity for direct contact with clients, industry-wide, to discuss their
specific problems, to assess the practical performance of new tools, and to
evaluate the methods proposed by various authors.
It seems fitting to complete the record by mentioning the contribu¬
tions of some of those who left Schlumberger, either to open their own
offices as consulting engineers, to join oil companies as logging experts, or
to become competitors. Whether they left on friendly terms or slammed
the door behind them, they took with them all they had learned, but this
did not affect their relations with their former colleagues.
Guyod was hired by a servicing company. Charrin and Castel created
a company of their own. A. A. Perebinossoff became a logging expert for
Socony. R. G. (Bob) Hamilton, a long-time Schlumberger leader in the
Mid-Continent, became a logging consultant in Tulsa, Oklahoma. Louis
Chombart, who, with Maurice Lebourg, had introduced an automatic
recorder into the United States, opened an office in Wichita, Kansas,
where he became an authority on carbonate rocks; among others opening
their own offices were Bob Seale in Dallas, Texas, because he wanted to
give more time to hunting and fishing, his true vocations, and Bob Kelso
in Houston. John Walstrom, who among other activities had contributed
to the writing of Doll’s paper on S.P., joined Standard Oil of California as
a logging expert and became a leader in the profession. While in Ven¬
ezuela, Leendert de Witte, more of a physicist than a field engineer, left
Schlumberger for Continental Oil Company, where he made significant

26 It was estimated in 1958 that over 9000 participants had attended the seminars during
the 3 preceding years.

235
The History of a Technique

contributions to the basic study of the interpretation of shaly sands. Frank


Millard and Ray Braeutigam took their Schlumberger experience to Car¬
ter Oil and Sinclair, respectively. Hamilton Johnson, after 10 years with
Schlumberger, first joined the competition and later turned to teaching.
There were many more who left Schlumberger for a variety of
reasons. By establishing logging services within the client companies,
organizing lectures and seminars, and publishing many articles, all of these
men made major contributions to the interpretation and dissemination of
logging interpretation techniques.

The new resistivity logs

The idea had prevailed between 1945 and 1950 that any problem of
interpretation could be solved on the basis of the S.P. and resistivity
measurements. Though this concept proved to be overoptimistic and
even erroneous, it nevertheless made an essential contribution to the
science of logging.
Before developing new tools and putting them into operation, it had
been necessary to gain a clear understanding of the various rock types
involved, to grasp the mechanism by which porous formations are invaded
by the mud filtrate, and to speculate both on the best way to obtain the
desired measurement with no or minimal corrections and on the benefit
of measuring the resistivity of the invaded zone in addition to that of the
virgin zone. In a short time this kind of thinking spread beyond the
Schlumberger research group. The field engineers played an important
role, along with the geologists and engineers of the client companies
(some of whom were already specializing in log interpretation) and the
universities and technical institutes. Although Doll no doubt deserves full
credit for the basic theory of resistivity determination in a permeable
formation, which took into account the parameters—borehole, depth of
invasion, shoulders—it is also true that he sought the advice and
experience of those dealing daily with interpretation. Adjusting theory to
practice was basic to his design of the array of tools which was to give
Schlumberger a substantial technical advantage and reaffirm its leadership
in the field.
There are two kinds of difficulties in measuring the true resistivity of
a porous formation, that is, the resistivity of the undisturbed zone beyond

236
After 1945

the invaded zone. In the case of fresh mud the invaded zone is much less
conductive than the virgin zone and acts as a screen to conventional
resistivity measurements; conversely, with salt muds the invaded zone is
much more conductive than the virgin zone, and most of the current flows
into the former while barely entering the latter, which thus escapes
measurement. For a three-dimensional current flow these two cases rep¬
resent the versions of a well-known electrical problem: how to evaluate a
resistance (X) that is inseparable from a spurious resistance (Y); the
solution is, when possible, to connect X and Y in parallel ifX is small with
respect to Y, and in series if X is large.
Induction and Laterolog, respectively, were the solutions. Realiza¬
tion, however, was difficult, especially for induction, with its requirement
for high frequencies and electronic controls almost beyond the pos¬
sibilities of that time.
Schlumberger engineers, as well as their clients, placed almost as
much importance on the resistivity of the invaded zone as on that of the
undisturbed zone. The reason was that the invaded zone, which has all the
petrophysical characteristics of the formation, could easily be analyzed
because it was saturated by water of known characteristics, namely, the
mud filtrate, samples of which could be taken at the surface. Experience
with the short normal, which under favorable conditions reads close to the
invaded-zone resistivity, confirms the advantage of knowing this value in
all circumstances.
It was realized that some oil always remains trapped, even in a zone
well flushed by the filtrate. Moreover, the actual values of this residual
saturation could be determined by laboratory analyses of conventional
cores and sidewall samples.
Therefore analysis of the invaded zone seemed highly promising; and
since the technical problems involved were much less serious than those
in the virgin zone, this became the first objective of the postwar program.
From this thinking and research came the microlog.

The Microlog Technique

In 1945 the RJRt ratio method (Æ, for the average resistivity of the
invaded zone, R t for the true resistivity) was already in current use. It had
been introduced in the Rockies by Tixier, and with a few extrapolations
its value was soon proved by the direct application of Archie’s formulae.

237
The History of a Technique

In cases of a thick layer, a shallow invasion, and a rather low resistiv¬


ity contrast between invaded and virgin zones, it was correctly assumed
that a normal sonde of sufficiently long spacing was adequate for the
measurement of Rù as for Ri, what was needed was a sonde of spacing
short enough to include only the invaded zone in the measurement. Yet a
normal sonde cannot be shortened beyond a certain point without the Rj
measurement being overwhelmed by the influence of the borehole mud.
An elegant solution proposed by Doll was to apply the electrodes against
the borehole wall by means of a rubber pad, which would also insulate
them from the mud. The spacing between electrodes being substantially
shorter, the measurement would extend only to a very small volume
immediately against the wall and entirely within the invaded zone. As
standard electrical logging equipment could be used, there was no serious
electrical problem to overcome. This was not quite the case, however, for
mechanical problems: many developed because of the behavior of the
sonde and pad in the hole, but without critically interfering with the
completion of field tests with an experimental model.
Since electrical logging equipment allowed for the simultaneous re¬
cording of several curves, three small round electrodes were inserted in
the pad 1 inch apart. The objective was the recording of three curves with
different depths of investigations: one normal AM, (1 inch), one normal
AM2 (2 inches), and one lateral AMiM2 (Fig. 69). Furthermore the ex¬
perimental model could simultaneously record the borehole diameter
through a makeshift electrical device actuated by the spring of the pad.
The first practical tests took place in May and June 1948 in Gulf Oil
New Mexico wells. The results were analyzed in Houston by a group of
experts under Doll, and all participating have retained a vivid memory of
the event. Bricaud, who had conducted the field tests, was so elated that
he called that day one of the brightest days of his life. Allowing for the
favorable conditions under which the tests took place, the outstanding
feature of the logs was the nearly perfect concordance among the various
resistivity curves, with the exception of those zones indicated as perme¬
able by the S.P., over which a large and constant separation between
curves could be noted; the curve with the deepest investigation reading
the highest. The separation and parallelism of the curves had a very
characteristic aspect. This result, obvious as it may seem today, afforded
some headaches at that time, all the more so because there had been an
error in the presentation of the experimental logs: the measurement of

238
After 1945

239
The History of a Technique

the diameter suggested a widening of the hole in the section with curve
departures, whereas a correct reading would have indicated a narrowing.
Bricaud had simply reversed the diameter scale. It remained for Doll to
explain both the curve departures and the borehole diameter reduction in
front of the permeable zones. The cause was the mud cake—the clogging
layer built up against the wall of the permeable formation, by the solid
particles suspended in the mud. The thickness of the mud cake that
separated the pad from the borehole wall—0.1 to 1 inch—was by no
means negligible with respect to the spacing of the electrodes, and a
miniaturized two-media problem reappeared.
To be sure, the mud cake was nothing new to mud control specialists,
nor should it have been to Schlumberger engineers, who daily collected
some of it on sidewall samples. Yet there was surprise and perhaps even
some discouragement. The two-media problem was no happy omen.
Since experience with conventional logging and departure curves gave
scant hope that it would ever be feasible to obtain the Ri value through
adequate corrections, the question arose as to whether the idea of measur¬
ing the resistivity of the invaded zone directly should be abandoned. No
one, at that time, could have anticipated what a remarkable future lay
ahead for the method just devised.
Even though the desired resistivity was not always measurable, and
there was little hope for a direct evaluation of the formation factor and
hence the porosity, the fact remained that the new device located the
permeable zones with an accuracy heretofore unknown. Indeed, the S.P.
curve identifies permeable zones in many cases, and Doll’s recent study
had shown how to increase its accuracy. Yet the very same study pointed
to a large area where the S.P. gave uncertain results. This was typically the
case for consolidated rocks like sandstones and limestones, for which the
new method proved fully applicable. The much too specific name “Forma¬
tion Factor Fogging” was changed to “Microlog,” a vaguely scientific-
sounding and hence commercially attractive designation.
For the detection of mud cakes and, therefore, permeable zones, the
Microlog is a first-rate instrument. In consolidated formations, where the
best that can be expected from the S.P. is to delineate permeable zones
with an accuracy of a few feet and where its ill-advised use can lead to
intolerable thickness overestimations, the Microlog locates the bound¬
aries within an inch. Because such accuracy facilitates the understanding
and interpretation of other resistivity measurements, the Microlog may be

240
After 1945

said to have played, in consolidated formations, the same role as S.P. in


sand and shale sequences.
After rather unspectacular field tests on the Gulf Coast, where S.P.
usually worked well, the Microlog was tried in Oklahoma with such
success that Gulf Oil used it on a wildcat in Alberta Province, Canada. To
everyone’s surprise the new technique revealed a thick, permeable, oil¬
bearing section that had hitherto escaped the geologists in their examina¬
tion of cores and cuttings. The incident created quite a commotion.
Tixier, then operating in Wyoming and Colorado, requested and obtained
the prototype; he proposed Microlog services to his clients with im¬
mediate success.
The Microlog was fully consecrated in Scurry County, western Texas,
where" an important discovery in 1949 was followed by feverish drilling
activity. The reservoir was a rather shallow limestone reef where oil had
migrated at random according to local variations of porosity and perme¬
ability. As the depth of the oil-water contact was well known, all that
remained to do to reach the oil was to find permeable zones high enough
in the structure. As often happened in shallow United States fields, most
wells in Scurry County were drilled by independent producers, some of
whom quickly ran out of funds and needed bank support for the produc¬
tion phase. As security for the loan, it became customary to submit the
Microlog diagrams to the bankers. Even though the local banker knew
little about geophysics, he understood enough to recognize the positive
separation of the curves and correctly evaluate the thickness of the
permeable beds. From that point on, in the way of bankers, he computed
the size of the loan. The petroleum world was greatly impressed by such
trust in the new method on the part of Texas bankers, who were not
known to rush headlong into rash ventures.
At any rate, the story shows the ease with which the Microlog can be
interpreted qualitatively—an ease that facilitated its success, and it is
worth noting that, in association with other measurements, the Microlog
continues to operate almost in its original form. Only two curves are now
recorded, the deep (micronormal) and the shallow (microlateral), with the
same spacing as initially: 1 inch between electrodes. The presentation has
also remained the same over the years, as its users have opposed any
change. One marvels at such lasting success: perhaps there were some
lucky decisions, but what passes as luck in such cases is actually a reward
of full technical background. What this success story mainly conveys is

241
The History of a Technique

that by 1948 complete mastery had been achieved in conventional resis¬


tivity measurement systems.
Of course, this was only a first step, as no one was content to consider
the Microlog merely as a mud cake detector, however valuable that
function might be. The hope still was that, with the help of two mea¬
surements having different radii of investigation and appropriate depar¬
ture curves, it would be possible to evaluate the resistivity, now called R x0
of the fully flushed zone. This approach, although applicable in some
specific conditions, did not lead to a general solution for R x0. The correc¬
tions were so large that the results could not be accurate and did not
justify the hard work required. A technique that would give a good
approximation of Rxo by direct reading was needed. The striking similarity
with conventional resistivity logging suggested to Doll that, since the
Microlog had proved the value of sidewall pad measurements, the princi¬
ple of focusing, lately developed for the “Laterolog” (see p. 257), would
also be applicable to a pad device. This was to become the “Mi-
crolaterolog.” However, it required a downhole electronic technology that
could not be realized before 1952. Meanwhile, the Microlog was im¬
proved over the years, the main effort bearing on the mechanical system
of this forerunner of the increasingly numerous and important pad-
mounted tools.
The year 1952 saw the first hydraulic pads, invented by André
Blanchard, who, from his years with Michelin, had retained a fondness for
rubber. To allow for better contact with the borehole wall, he had de¬
signed an oil-filled rubber pad. The system proved to be of interest
inasmuch as the absolute measurements remained substantially un¬
changed, while the mechanical behavior was better, and the results were
more consistent.
The next step was the improvement, in 1954, of the sonde itself.
Until then it had been a rather crude tool made of a mandrel centralized
by two opposite spring blades, one of which carried the pad (Fig. 69). The
system was simple but had serious drawbacks. Because the pad rubbed
against the wall on the way down when no recording was being taken, as
well as on the way up, it wore out rapidly. In deep and difficult wells a
damaged pad could mean a lost job. The first solution was completely
makeshift; the springs were tied against the sonde body with a piece of
rope, expected to be worn out by friction and to break by the time the
tool reached the bottom. Then (it was hoped) the two springs would be

242
After 1945

released, the pad pressed against the wall, and the log recorded on the trip
up the hole. Satisfactory as this remedy may have been to engineers who
recalled the pioneering days of “La Pros,” it was plain tinkering. The
Engineering Department produced something better, the so-called
“warhead ’: the sonde was locked in closed position when being lowered
downhole, and the springs were released before logging by electrically
firing a small explosive cartridge from the surface.
Even this was hardly satisfactory. It did not eliminate unnecessary
pad wear when the sonde was pulled out after logging, or when it had to

Figure 70. Motorized microlog sonde.

243
The History of a Technique

be lowered again to check or repeat a log. There was another very serious
shortcoming. It took only a few cuttings caught between the sonde body
and the springs to block their release or, alternatively, their closing, and
on the way up, the system could become wedged in a keyseat or restric¬
tion of the borehole.
The true solution was to replace the springs by a remote-controlled,
bidirectional, articulated arm. The downhole environment, however,
caused complications. The problem was solved in 1954 by M. Lebourg
and R. Q. Fields (Fig. 70) in Houston. Through a hydraulic system a small
electric pump controlled the opening and closing of the tool, while an
appropriate kinematic configuration ensured that the pad remained paral¬
lel to the borehole wall. This was the first in a series of “power sondes.” Its
cost was about eight times that of the conventional sonde, a feature that
the financially responsible field engineers did not appreciate. But it did
become the starting point of a technology that, unobtrusively, was to give
Schlumberger an unquestionable superiority and a lead difficult to over¬
take. Over the years many and diverse tools made their appearance that
called for tight contact between their specialized components and the
borehole wall. All of them proved to be outstanding in performance and
dependability.

Induction Logging

In a first stage, general considerations on resistivity measurements


had led to the development of the Microlog. The same considerations
showed also the need to know accurately the resistivity of the undisturbed
zone. It might even have been logical to tackle the latter problem first; yet
the existing system, however imperfect, gave substantial results, and
improved measurement of true resistivities beyond the invaded zone
posed far tougher practical problems than those encountered in develop¬
ing the Microlog, which, in fact, was nothing but an extension of the
resistivity log.
As long as the invasion remained moderate, the conventional mea¬
surements—short and long normal, lateral—gave satisfactory results, as
on the Gulf Coast and in Venezuela, Trinidad, and California. The objec¬
tives there were highly porous sands drilled with fresh mud, where the
discrimination of salt water and oil beyond the invaded zone was almost
always possible with the long normal. Systematic difficulties occurred

244
After 1945

when drilling became deeper in these same regions, and also when a more
thorough interpretation was attempted where the more consolidated
sands turned into sandstone (northern Texas and Louisiana, southern
Oklahoma, etc.)- In both cases the depth of invasion increased as porosity
decreased.2. Longer spacings would have reduced the vertical resolution,
and no great improvement could be expected from increasingly rigorous
corrections: even in thick beds the laborious use of departure curves
brought dubious results. The truth was that in such difficult cases, when it
was impossible to reach a correct fluid analysis, Schlumberger was hardly
better off than its competitors: without fluid analysis the only remaining
use of logs was well-to-well correlation, for which competitor logs were
equally good. Under such conditions, technical superiority was not a
decisive commercial advantage.
The overriding problem was undoubtedly the fact that, according to
the salinity of the mud, the invaded zone could be either more or less
resistive than the undisturbed zone; hence any resistivity contrast could
be encountered. Theory, confirmed by experience, shows that one and
the same measuring system is not suitable in both cases, and the question
may be asked why the system requiring what seems to the layman the
most difficult technical solution was developed first. There were several
reasons for this. Not least was the “challenge” of the oil-base muds (see p.
161), introduced in the 1930’s and appearing for a while to have great
potential for expansion. As electrodes insulated by oil cannot conduct
current, the technique precluded conventional logging—an aspect that
would have been a heavy blow for Schlumberger. As a matter of fact, only
cost and operational difficulties28 prevented these muds from spreading.
To make a resistivity log it was necessary to replace the electrical
contact with the mud by induction, which, by virtue of the time-varying
electromagnetic fields, would permit the circulation of currents in the

27 This is only an apparent paradox, as filtration is controlled by the mud cake; whatever
the formation, the amount per unit of time of the liquid filtering through is constant.
Consequently, the lower the porosity, the greater is the volume required to absorb a given
amount of filtrate.
28 Anyone who has worked on a well drilled with oil-base mud appreciates the "tidiness”
of operating with water-base mud. There is another drilling method that normally excludes
conventional logging: percussion or cable-tool drilling of empty holes, but scant attention
has been paid to this system, which was in declining use even in the 1930’s, whereas rotary
drilling with oil-base mud threatened to spread widely.

245
The History of a Technique

formations and their measurement. At first it was only a dream—a Jules


Verne fantasy, according to Doll, because the technology required for
such a system was simply nonexistent. This did not prevent the devotion
of a great deal of thought to the matter. It soon became apparent that
induction logging, mandatory for oil-base mud drilling, might also be the
way to skirt the “electrical barrier” of a resistive invaded zone in water-
base muds. This is one of the basic features of induction logging, as
stressed by Doll in his 1949 paper (see p. 233). In conventional logging
the current lines cross boundaries between regions exhibiting different
resistivities; hence it is impossible to assess individually the influence of
each portion in the volume measured because a change in resistivity
affects the current lines not only where it occurs, but also over their whole
path. On the contrary, the currents generated by induction being
circular and coaxial with the sonde, whose axis, in turn, is that of the
borehole, the cylindrical symmetry of the formation about the hole results
in each current line remaining within a medium of constant resistivity, at
least in horizontal or subhorizontal beds. Furthermore; as long as fre¬
quency and conductivity are not too high, these various elementary cur¬
rent loops have only a negligible mutual action, and the effect proper of
each region may be considered separately, the total signal being merely
the sum of the individual signals. Once the share of each portion of the
volume involved in the measurement—borehole, invaded zone, transi¬
tion zone, undisturbed zone, shoulders—is exactly known, correcting
the readings to obtain the true resistivity becomes much easier than with
the departure curves used in conventional logging.
Still better, this theoretical simplification made it possible to envisage
more refined systems than the two coils. It became feasible to foresee and
calculate the characteristics of a system of coils increasing or decreasing
the proportion of the signal emitted from a given region: the system is
called “focused,” and it can be focused vertically or radially in space. To a
certain extent, the lateral depth of investigation of the instrument may be
increased while maintaining a good vertical resolution; in other words it
became possible to read “in depth” inside a rather thin layer. At least in
theory, these totally new possibilities gave induction logging a vast
superiority over any alternative hitherto known, particularly in boreholes
drilled with low-salinity mud. The physical properties mentioned above
and the ensuing mathematical simplifications led to an early development

246
After 1945

of the theoretical basis of induction logging by Francis Perrin29 and André


Blanchard, followed in 1948 by André Poupon.
During World War II, as noted earlier, Doll applied the technique of
induction to a jeep-mounted mine detector (see p. 200). Heartened by
this success and well informed concerning the possibilities of modern
electronics, he started to develop the induction sonde (Fig. 71). The
transmitter coil, usually fed with a 20 kilohertz current, produces a
variable magnetic field generating eddy currents, which follow circular
coaxial paths in the formations surrounding the borehole. These currents
create a secondary magnetic field, which, in turn, induces an alternating
voltage in the receiver coil. The intensity of the eddy currents and of the
signal induced in this coil increases when the formation resistivity de¬
creases. Contrary to conventional logging, which records resistivity, in¬
duction logging measures its reciprocal, that is, conductivity. The theoret¬
ical unit is the siemens-m/m2, the reciprocal of the ohm-m2/m (called the
mho by electrical engineers as a reminder of the link between resistance
and conductance units). To avoid decimals in the majority of cases where
formation resistivity exceeds 1 ohm, it was decided to use the one-
thousandth submultiple, or millimho.
Although only a limited knowledge of Faraday’s laws is needed to
understand the mechanism of induction, it may be useful to point out
some difficulties. The only voltage of interest is that induced by the eddy
currents generated in the ground, yet the voltage resulting from the direct
coupling between transmitter and receiver coils can be 1000 times
stronger. The approach then is to generate a voltage equal to that of the
direct coupling and subtract it from the total signal. However, this sub¬
traction still leaves a residual error signal, caused in particular by the
thermal drift. To sort out the useful signal one must use the fact that the
voltage induced by eddy currents lags 90 degrees behind the one induced
by direct coupling. Thus, since the system requires great stability under
thermal and mechanical stresses, the simplest induction sonde is depen¬
dent on refined electronics.
Because of this complexity Doll had to win a psychological battle
before designing his first version of an induction log. Except for his close

29 To F. Perrin goes the credit for the concept of the “geometrical factor,” permitting
calculation of the response of the sondes and precise analysis of the induction logs.

247
The History of a Technique

Figure 71. The principle of induction logging. (From H. G. Doll, “Introduction to Induction
Logging and Application to Logging of Wells Drilled with Oil-Base Mud,” Journal of Petro¬
leum Technology, June 1949). (Courtesy SPE of AIME).

248
After 1945

co-workers,30 no one was on his side. On the minds of management, field,


and research staffs alike was the considerable technical and financial effort
just invested in the new truck, the new chronological system, and the new
recorder; they wanted a breathing spell before embarking on other and
more costly technical ventures. What must also be mentioned is that the
advantages of induction were not quite as obvious as they became later,
and the field engineers were ill prepared for its electronic complexity.
Many thought it foolhardy to substitute a complicated, expensive, and
without doubt unpredictable apparatus for the simplicity of three elec¬
trodes mounted on a mandrel. The saving grace was the wells drilled with
oil-base mud, where, at least as an auxiliary, induction found its rationale.
They ^constituted, not a large market, but one substantial enough to be
taken seriously. What probably tipped the balance was the fact that
Henquet, the new assistant general manager, had worked for a long time
in California and was aware of the serious oil-base problem there. In
1945, objections and doubts notwithstanding, the prototype was put on
the drawing board.
The first induction log was recorded on May 3, 1946, by O. H.
Huston, assisted by R. T. Wade, C. Bailey, and R. Theis, in the Humble
Faulk No. 7 well in the Hawkins field near Tyler, Texas. As was true of
the logs of September 5, 1927, at Pechelbronn and March 6, 1929, at
Cabimas, there was something almost touching in this unassuming little
curve. Drilled with oil-base mud, the well did not allow for corroboration
of the results by a conventional log; the only possible comparison could
have been with a poor log taken with scratchers (see p. 161). The induc¬
tion (Fig. 72) indicated the presence of an oil zone toward the bottom of
the hole, yet even today one finds it difficult to believe that this log was
the first in a series that was to revolutionize the industry. Considering that
in normal muds conventional logging in the area gave results of high
quality, one can only pay tribute to Doll’s optimism and faith and his
justified conviction that induction would win. Although marketed in some
centers from 1948 on, the log was hardly utilized except in oil-base muds
until 1952. In fact Doll’s paper of June 1949 was specifically captioned
“Introduction to Induction Logging and Application to Logging of Wells
Drilled with Oil-Base Muds” (p. 233), but it again recorded and em-

30 Especially O. H. Huston and G. K. Miller, who took a prominent part in the design and
realization of the original equipment.

249
Figure 72. The first induction log.

250
After 1945

phasized his conviction that the system would be equally valuable in


water-base muds.
It is worth noting in the same paper the progress achieved since
1946. A sample log is compared with a good conventional log of the same
well: considering the scale of the induction log, the correspondence is
excellent. Conductivity, represented on a linear scale, increases from right
to left, while resistivity is by convention always shown as increasing from
left to right. The two curves thus move in the same direction, although in
terms of resistivity the scale of the induction log is reciprocal and the
variations of the smaller resistivities are emphasized.31 This somewhat
unusual presentation in no way affects the quantitative usefulness of the
result^. As long as induction was confined to wells drilled with oil-base
mud, this question of scale hardly mattered; but when it came to water-
base muds, direct comparison with conventional logs was required.
However, this came later. It took several years to produce more
dependable and better focused sondes with low sensitivity to a borehole
filled with conductive mud. Meanwhile, Doll had reconciled the field staff
to his ideas, and the engineers began to appreciate the advantages of
induction over the long normal under the operating conditions prevailing
along the Texas-Louisiana Gulf Coast, where most of Schlumberger’s
activities were concentrated. Commercial success, however, was far from
immediate. First, in fields with sands of high porosity and shallow in¬
vasion, satisfactory results were obtained with the long normal,
supplemented by the lateral when needed—if the formations were thick
enough. Indeed, the advantage of induction would have been clearer in
formations 6 or 10 feet thick, but the use of the long normal had become
so deeply rooted that the thickness effect could be assessed even without
the assistance of departure curves, or allegedly so. On the other hand, the
induction log required a special run, costly in time and money. It was

31 The difference between 1 and 2 ohms is

1000
1000 - = 500 millimhos.
2
Between 100 and 200 ohms the difference is only

1000 1000
millimhos
100 200

and between 1000 and 2000 ohms it is 0.5 millimho.

251
The History of a Technique __

therefore out of the question that the induction would be substituted for
the conventional log, on which the S.P. immediately gave the lithology,
and the comparison of the short and long normal provided a good qualita¬
tive appraisal of the production potential.
After being unreservedly accepted in oil-base mud drilling, induction
began, after 1952, to spread to some conventional wells. This happened in
northern Texas and Louisiana, where the reservoirs consist of sandstones
of much lesser porosity and deeper invasion than in the southern parts of
these states. The greatest benefit of this heartening initial success was the
conversion to induction, first of Schlumberger unbelievers and then of the
Company’s clients. Only a good S.P. and a short normal complementing
the induction were required to relegate the conventional log to a museum
piece.
The solution was far from easy. One characteristic of the induction
sonde is that its response is distorted by the slightest metallic influence,
whether from electrode or conductor. Only in 1956 was a downhole tool
completed that combined an S.P. electrode, a conventional 16 inch short
normal, and a five-coil induction sonde with vertical and radial focusing.
The presentation of the log was modified to conform with that of the
conventional electrical log. On the right-hand track, the induction curve
was recorded on a linear conductivity scale; it replaced the lateral. On the
middle track, the reciprocal of the induction appeared as a linear resistiv¬
ity curve. This curve was dotted and replaced the (dotted) long normal. It
was obtained through a “reciprocator,” an electronic system which con¬
verted the induction signal arithmetically.32 As in the past, the short
normal was recorded in the middle track and S.P. on the left (Fig. 73).
In spite of its higher cost, its more difficult operation, and the still
quite satisfactory conventional systems, this new form of induction log¬
ging soon prevailed all over the Gulf Coast. Any remaining hesitation was
overcome by superior results and the confirmation of theory by practice:
once the major oil companies saw the light, all the others followed suit.
Such unanimity, however, had its drawbacks. All the former doubters
became enthusiasts and demanded induction even in situations where the
new system was obviously less efficient: whereas induction is highly
suitable for measuring a conductive formation behind a resistive barrier, it
is much less so in the case of a resistive formation behind a conductive

3- R — 10001C, with R in ohms and C in millimhos.

252
After 1945

Figure 73- Comparison between induction-electrical log and conventional log.

invaded zone. It was not long before Schlumberger and the client in¬
terpretation specialists reverted to a more rational approach: that of
utilizing the Laterolog, by this time commercial, in cases where induction
was deficient.
Induction logging attained its definitive form in 1956, and since 1957
its use has become worldwide wherever conditions have been favorable.
Such was its success that competitors had to develop their own induction
systems and rediscover not only the technique but also the know-how
required for making and operating the equipment. They succeeded and
partially overcame their lag, but Schlumberger, as for conventional log¬
ging, had pioneered the service which, in a short 10 year span, had
changed the course of the logging industry.

253
The History of a Technique

The Laterolog Technique


From the beginning of logging the technicians had been greatly
preoccupied with bed resolution, particularly that of very thin beds. As
early as 1927 Conrad had devised a system which, through the use of
elongated, so-called guard electrodes, forced the current of a central
electrode to penetrate laterally into the formation it faced. As shown
earlier (see p. 124 and following), by opposing the spread of the current
this device improved the delineation and detail of the formations. Conrad
had described another system, where point electrodes would give the
same effect, but unlike the former, it never went beyond the conceptual
stage.
The failure of conventional logging to provide a sharp representation
of the beds is found mainly in cases where there is a great contrast
between the resistivity of the formation and that, much lower, of the mud.
The boundaries between layers become blurred, and even the lithological
interpretation of the log is dubious. In consolidated sandstone or lime¬
stone reservoirs, even with fresh mud, there is a large resistivity contrast.
The situation becomes considerably worse when the shoulders are evapo-
rites of practically infinite resistivity and the well is drilled with salt-mud.
Western Kansas is a typical case where Paleozoic limestones lie under
hundreds of feet of rock salt, too deep to justify the cost of an inter¬
mediate casing. Cave-ins produced by dissolution of the salt can be
prevented only by allowing the drilling fluid to reach saturation. Its
resistivity is then less than 0.1 ohm, which is a ratio of 1 to 1000 when the
reservoir resistivity is 100 ohms. Similar conditions prevail in the impor¬
tant Permian Basin of western Texas, except that the salt occurs in several
thinner beds. There, again, conventional resistivity logs are poor; and for
lack of salinity contrast between mud filtrate and formation water, both
salt saturated, the S.P. is useless.
Little activity took place in these regions until 1950, but, as in most of
the Mid-Continent, an important potential market was there ready to be
exploited. None were more convinced of this than the few Schlumberger
engineers who were struggling to promote their logs in these problem
areas that made the Gulf Coast look like an El Dorado by comparison.
They were well aware of the weaknesses of the conventional system, and
tended to think that the Engineering Department considered logging to
be destined exclusively for the Gulf Coast. According to the Schlum-

254
After 1945

berger tradition of uninhibited comment, they were rather outspoken in


their views.
When Marcel visited the United States in 1945, he bore the brunt of
their criticism. Since experience in the region demonstrated that the
lateral curve showed at least some stratigraphic features, provided that its
peculiarities could be sorted out, he conceived the idea of combining two
lateral sondes head to tail. The assembly, being symmetrical, gave a
symmetrical response on lithology and eliminated the most serious
shortcoming of the lateral sonde. However, it was quickly evident that the
apparent thicknesses of the resistive beds were reduced at their top and
bottom boundaries by the basic sonde spacing. To preserve the lithologi¬
cal usefulness of the system, it became necessary to reduce the basic
spacings A0X and A02 (01 and 02 being the middle of MjNj and M2N2,
respectively), thus lessening the depth of investigation to the point of
making the system sensitive to variations in the diameter of the borehole
(Fig. 74). The elimination of the lateral’s distortion had also removed its
main quality, the depth of investigation. All this was confirmed by experi¬
ence. Whereas this “limestone sonde” had been immediately commis¬
sioned in the Permian Basin, its only use was found in log correlation.
Although some clients continued to request it even after 1950, it re¬
mained confined to western Texas, a fate that is not very exciting for a
logging technique.
The solution lay elsewhere. The delineation of resistant beds in a well
drilled with a salt-saturated mud was incompatible with the spreading of
the current toward the more conductive shoulders; it was therefore neces¬
sary to direct or force the current into the formation of interest or, to use
the term that was soon to characterize all the new measurement systems,
to “focus” it. This is what the guarded monoelectrode sonde had tried to
achieve. The two voltage measurements required were ill adapted to the
continuous recordings of the 1950’s, but this was not an insuperable
problem. More serious was the fact that the effectiveness of the focusing
depended on the selection of the shunt. At this point Doll came forward
with a solution that was both original and technologically feasible. To
prevent the current from escaping through the conductive borehole, he
thought of plugging it—electrically, of course.
There was actually nothing new in the idea of electrically plugging
the hole, as Doll himself had already introduced it in his analysis of the

255
The History of a Technique

Figure 74■ Schematic view of the limestone sonde.

tridimensional S.P. current. In his 1948 paper he had defined “static S.P.”
as the potential difference that would be created if the current were
interrupted by two imaginary “electric plugs” at the top and bottom of the
bed. In another study he had gone so far as to realize such plugs in the
so-called “Selective S.P.” system, directly derived from his reflections on
S.P. in formations of very high resistivity, with the purpose of locating
porous zones in such formations. Meanwhile, an easier solution to the
problem had been provided by the Microlog, but the principle of the plug
remained: to plug the current between the two electrodes M and N,

256
After 1943

auxiliary electromotive sources are used, with their outputs adjusted so as


to balance the potential difference between M and N.33 This was the
system that would be used for the automatic focusing of measurements in
salty muds.
The sonde array is made of seven electrodes (hence the name
“Laterolog 7”): three current electrodes, A 0, A x, and A 2, and two pairs of
measure electrodes, M XM ' j and M 2 M ' 2 (Fig. 75). To prevent the current
emitted by A0 from spreading vertically and to force it into the resistive
formation, the borehole must be electrically plugged, at the level of pairs
MxM\ and M2M2, by emitting from A, and A2 currents (called “bucking
currents’’) adjusted to permanently balance the potential difference be¬
tween the electrodes of each pair. In fact, the result is even better than
merely plugging the hole, since the plug overflows into the formation,
where it maintains the current emitted from A0 in a well-delimited hori¬
zontal sheet. As for conventional logging, resistivity is measured on the
basis of the potential common to M x M\, M 2 M'2.34 A little masterpiece of
simplicity of ingenuity, Laterolog 7 was designed by Doll in almost
finished form. It then became possible to study its characteristics, since
the only problem was to superimpose the effects of several point elec¬
trodes and to optimize their geometry. The same held true for the
calculations of the currents, which, as could be expected, had to be very
strong at the focusing electrodes in the difficult case of a thin and resistant
layer between conductive shoulders. Serious technical problems arose
because for the monitoring of these heavy currents it was necessary to use
the very small potential differences appearing between M and M ',35 which
had to be reduced and maintained at a value near zero. In 1950 electronics
was not sufficiently advanced to solve this “gain” problem by downhole
amplification at frequencies intentionally equal to those of conventional
electrical logging. The possibility did not even present itself that, for the
measurement of resistivity, anything other than the conventional and
proven pulsated current could be used. The decision was a fortunate one;
although the pulsator was taxed to the limit of its capacity, it allowed time
for the rapid and technically sound realization of the first Laterolog.

33 The limiting case is MN infinitely short; the potential gradient in the vicinity of M is
then said to be null.
34 In practice, electrodes 7W, and M2,M\ and M2 are short circuited, as well as A, and A2.
35 This is the potential difference existing between the short-circuited electrodes M1M2
and M\ AT2.

257
The History of a Technique

Figure 75. Schematic view of laterolog 7. (From H. G. Doll, “The Laterolog: A New
Resistivity Logging Method with Electrodes Using an Automatic Focusing System,” Jour¬
nal of Petroleum Technology, November 1951). (Courtesy SPE of AIME).

258
After 1945

Because of the difficulty of adjusting the bucking current downhole,


all operations took place above ground, making it necessary to send
the weak signal MAI' to the surface and to inject the strong bucking
current from the surface. From the perspective of the 1970’s, the
operation may seem somewhat primitive. However, it gives some
insight into a rather dependable technology borrowed from the arma¬
ments industry as it had developed at the end of the war. The MM' signal
was electronically amplified and drove a small motor in either direction.
The motor actuated a rheostat (similar to a manual operation) which, in
turn, controlled the bucking current. A mobile pointer on the control
panel indicated the motion of the rheostat and informed the engineer on
the damping of the system. If the pointer vibrated, the system was
underdamped; it was overdamped when the pointer moved slowly. A log
with sharp contacts required a slight underdamping, obtained by adjusting
the “gain.”
Since the bucking current was pulsed, it could influence the monitor¬
ing signal by induction. This is the pernicious effect of direct coupling,
highly instable because of self-amplification. Only by a thorough study of
the chronological system and of the inductive coupling between cable
conductors was it possible to work out a solution and balance the circuits.
Of course it would have been very desirable to amplify the MM' signal
within the sonde; but, as already noted, a sufficiently stable electronic
amplifier at such low frequencies and under prevailing conditions was
beyond reach. A partial remedy was to place in the sonde a “square wave
transformer”; the name brought shudders to electricians, but the device
boosted the signal 10 times.
All this required the highest precision. To mention only the pulsator,
it had, on the one hand, to be perfectly balanced so as to avoid any stray
signal; on the other hand, it was expected to switch currents up to 2.0
amperes on and off without fireworks. This required absolute cleanliness.
The answer to these constraints was the high-precision cam pulsator,
developed in 1957. A further complexity introduced into the sonde is
worth mentioning: in addition to the square wave transformer, it was
fitted with a relay through which the electrode configuration could be
changed from the surface, the purpose being to switch from Laterolog 7 to
the limestone sonde—the very tool it was supposed to supersede!36

36 The switch may have been without much technical interest, but afterwards the relay

259
The History of a Technique

The first experimental log was recorded on April 11, 1949, north of
Houston. Though the results were satisfactory, they did not arouse much
enthusiasm because they were obtained in a fresh mud well and offered
little advantage over a good conventional log. In August of the same year,
a demonstration involving several instruments, including the Laterolog,
was organized in an important Gulf Oil well near Ardmore, Oklahoma, in
the presence of high-level experts: here again, nobody was convinced of
the Laterolog’s superiority over conventional logs.
So far, the objective had been only to test the tool operationally. The
real testing started in western Texas in February 1950 and continued in
Canada in May. Finally, the Laterolog was put into operation in Kansas in
August 1950. Wherever drilling took place with salty mud, the success
was overwhelming, and conventional logging was quickly eliminated.
When Doll published his November 1951 article in the Journal of Petro¬
leum Technology, the Laterolog was already an established technique.
Such success was like a whiplash to competitors, who in the same
year produced their own version of the tool under the name “Guard
Electrode Log.” The system was not unlike that of Conrad’s old guarded
monoelectrode sonde. Focusing was achieved by nullifying the potential
difference between the central electrode and the guard electrodes. The
technology was all electronic; but either because the components then
available lacked stability, or perhaps because it was more difficult to
determine the potential difference between current electrodes than be¬
tween measure electrodes, the competitors had tough going until they
were able to demonstrate technical performance which matched that of
Laterolog 7. In any case, when confronted later with the problems of very
thin layers, Schlumberger was to experience the same difficulties.
There was, however, great interest in the guard electrode log in
Kansas, where the productive horizons of the Pennsylvanian are com¬
posed of thin beds. It allowed for a narrower A0 current sheet than did
Lateroldg 7, actually a small advantage but enough for the competitors to
advertise it ad nauseam. R. D. (Bob) Ford, then manager of the area that
included Kansas, intervened strongly in Houston; his message, loud and
clear, was persuasive to the extent-that a Laterolog with guard electrodes,
equivalent to the competitive model, was placed on the drawing board.

made it possible to have a log at a "compressed scale,” that is, linear as to resistivity on one-
half of the strip, and as to conductivity on the other.

260
After 1945

This did not constitute plagiarism in any way since it was a mere resump¬
tion of the 1929 idea of the guarded monoelectrode.
Doll authorized the project without great enthusiasm. An electronic
Laterlog with a narrow current beam (“Laterolog 3”) was made, but with
results far less satisfactory than those obtained with Laterolog 7. An
improvement made in 1953 consisted of inserting small measure elec¬
trodes between current and guard electrodes, making it possible to nullify
the potential gradient as in Laterolog 7. Laterolog 3 became, in fact, a
Laterolog 7 in disguise.
However, Laterolog 3 offered a possibility that, strangely enough,
passed unnoticed initially. The upper guard electrode was long enough to
receive a gamma ray recording cartridge. This subject will be discussed
again under nuclear logging; it suffices to say here that gamma rays, in any
kind of mud, provide good detection of shale beds, and in this sense the
gamma log prevails over S.P. whenever the latter is suppressed by the
salinity of the mud. The combination of the Laterolog and gamma log
quickly gained ascendancy in Kansas and in the Permian Basin of Texas.
Still another factor made the study of Laterolog 3 fruitful: this technology
was to prove indispensable for the development of the “Microlaterolog,”
the natural extension of the Laterolog. In fact—and this constituted a
substantial practical advantage—an electronic cartridge was soon produced
that could serve for either Laterolog 3 or the Microlaterolog.
The success of the Laterolog technique in the United States made it
the standard resistivity log in salty mud. However, because it remained
confined for such a long time to areas where these muds prevail, its
commercial success did not match that of the induction log. Only much
later, and initially in the eastern hemisphere, did the Laterolog regain its
standing. No longer is the sole reason for its use the fact that it gives a
significant resistivity log in saltwater mud, which neither the conventional
nor the induction log can do: it is also suitable in situations where the
resistivity of the invaded zone is lower than that of the undisturbed zone. In
such cases an accurate evaluation of the latter requires measurement in
series, which is precisely what the Laterolog provides.

The Microlaterolog Technique

It has been explained how the existence of the mud cake between the
pad of the Microlog and a porous formation interferes with measurement

261
The History of a Technique

of the resistivity of the invaded zone. To prevent the current from


escaping through the conductive medium of the mud cake and to force it
into the formation to be measured, a system similar to the Laterolog is
needed. Seven electrode devices are placed on the pad. Since the prob¬
lem is no longer to block only the vertical leaks, but rather to have a
system electrically focused in any direction, the electrodes have a circular
pattern (Fig. 76).37 If, by modulating the output of the guard ring A1} the
potential difference created between M1 and M2 by the steady current
emitted from electrode A0 is balanced, this current is forced to flow
through the mud cake into the formation. The spacing of the electrodes
determines the outward shape of the beam emitted from A0; for a correct
measurement of the invaded zone, it must remain cylindrical for several
inches, with a diameter somewhere between those of rings M, and Af2—
about 2 inches—before opening up. It is assumed that the beam will flare
at some 3 inches from the borehole wall, and this value, which represents
the depth of investigation of the instrument, can be modified by changing
the spacing of the rings.
With the focusing system of the Microlaterolog, the contact between
pad and borehole wall is, basically, a less critical problem. A nearly perfect
contact is positively required for the Microlog only; nevertheless the
Microlaterolog has had the benefit of all the mechanical improvements
added to the former, for example, the hydraulic pad and the powered
sonde.
Whereas the Microlog and the Laterolog are systems in which the
downhole electrical equipment consists merely of measure and current
electrodes, the sonde of the Microlaterolog required electronics. Difficul¬
ties did not prevent its early realization. The first tests were conducted in
1950, and marketing was successful in Kansas in 1951. In October 1952
Doll disclosed its characteristics and potentialities to the American Insti¬
tute of Mining Engineers in Houston.
Technically successful as it was, the Microlaterolog did not arouse the
same enthusiasm as the Microlog had 2 years earlier, nor did it develop
into the same profitable business. The reason was that, whereas its appli¬
cation was confined to quantitative interpretation, the Microlog had be¬
come the basic tool for the location of permeable beds in consolidated

They are not actually continuous circles, but concentric circular strings of intercon¬
nected electrodes.

262
After 1943

Figure 76. Pad of the microlaterolog and schematic distribution of the current lines. (From H.
G. Doll, “The Microlaterolog” Journal of Petroleum Technology, January 1933) (Courtesy
SPE of AIME).

formations. However, the Microlaterolog was to open new horizons to


the specialists in interpretation, as it allowed a more thorough elaboration
of the ideas put forward by Tixier in 1942 (see p. 203).
Instead of considering a rather vaguely defined invaded zone, the
emphasis was now on the zone of maximum displacement of the original
fluids by mud filtrate. In this 2 to 4 inch thick section—precisely that
affecting the Microlaterolog—all the interstitial water has been replaced
by the mud filtrate; as to hydrocarbons, if any, most have been flushed
away, leaving in place only the amount which cannot be moved, usually

263
The History of a Technique

termed irreducible. The resistivity of this zone, so effectively measured


by the Microlaterolog, is called Rxo. If Sxo stands for the proportion in the
pores of a mud filtrate of resistivity Rmf> the Archie equation becomes

FR mf
R xo

whereas, applied to the undisturbed zone, it remains

FRW
Rt C2
«-> 1

Therefore

(1)

It is sometimes possible to obtain a fair idea of the residual oil proportion


(1 - Sxo) by a laboratory analysis of various cores. Then Sw may be
calculated through equation 1.
When Sxo is unknown, it is necessary to formulate a hypothesis based
on an approximate empirical relation between Sw and Sxo. Such a relation
will be similar to that observed by Tixier in the Rockies, except that,
under the stricter conditions defined above, it. will be closer to reality.
Assuming that

= (Sw)115

it follows that

/ RxolR mJ $18

l Rt/Rw !
At this point the essential data for its evaluation, the oil and water
percentages in the pores of the reservoir, are known.
Reverting to equation 1, if the term under the radical equals 1, that
is, if

Rxo _ Rt
Rmf Rw

then

= 1 and Sw = S
5 xo

264
After 1945

which generally means that

Jw J xo 1

This will be the case for a water-saturated formation, where the


invasion did nothing but substitute the mud filtrate for the original water.
But even if Sw and Sxo differ from 1, their equality means that the
water : oil ratio was not affected by the invasion—in other words, no oil
has been flushed by the filtrate. The fact that there was no flushing of
the oil indicates a virtually nonproductive reservoir. It will
happen, for instance, with an oil so viscous that its recovery through
conventional methods is not feasible.
The same method leads to an interpretation technique proposed as
early as 1952. Its later application, however, was devised by Doll and
published jointly with M. Martin (see note 24, p. 234). It is known that
from the porosity ((/>) a satisfactory value can be deduced for the forma¬
tion factor (F):

F - (3)

The Archie equation then becomes, for the undisturbed and invaded
zones, respectively,

Rn
R, = 4>2S'w
•2

R XO Rmf
<t>2S2xo

Here 4>SW, the product of the porosity and the water percentage,
represents the quantity of water by unit of formation volume. Likewise,
4>Sxo is the amount of filtrate in the invaded zone. The amount of oil
displaced by the invasion is thus

Y = 4>Sx0 ~4>SW

by unit of volume, or, on the basis of the above equations,

T =

The interest in the formula will be readily understood as it informs

265
The History of a Technique

not only on the amount of oil in the undisturbed reservoir, but also on its
mobility or its “producibility.”
It is still a puzzle as to why the Microlaterolog, with all these assets,
was not an immediate commercial success. Without doubt its advanced
technical features made it more difficult to operate than the Microlog;
hence it lacked what the field people called “sex appeal.” Yet, as early as
1950, quite a number of logging experts were using it profitably and were
even able to elaborate novel methods of interpretation based on the
Microlaterolog.
It must be acknowledged that the reason for its half-failure was the
often dubious quality of the results. Serious practical difficulties existed.
The first was the lack of dependability characteristic of the electronics of
that period because they were being employed in a definitely hostile
environment. A further obstacle was inherent in the Microlaterolog sys¬
tem: to read Rx0 beyond the mud cake, focusing is required, but it must
not be too deep lest the resistivity of the undisturbed zone be encom¬
passed. Yet the thicknesses of both mud cake and invaded zone are
variable. The former is contingent on mud quality: a good mud rarely
forms a mud cake thicker than 0.2 inch, though at that time 0.8 inch and
over was not uncommon; the latter, being a function of porosity, is
reduced in a very porous reservoir. It is a tightrope exercise; a focusing is
needed that suits every case even though it cannot pretend to apply to the
whole range of thicknesses.
There was another and more subtle problem of design. The pad must
be suitably curved to fit the borehole. Clearly, focusing, being determined
by the tridimensional geometry of the whole system, is affected by this
curvature; but since a borehole is never perfectly cylindrical, the pad is
continuously changing its shape to keep in contact with the wall.
Aside from electronics, the Microlog has encountered the same
problems but was content with being an excellent indicator of even very
thin permeable layers and hardly pretended to quantitative results. The
Microlaterolog, on the contrary, aimed at the highest accuracy. Its lack of
favor with certain experts arose from its failure to fulfill all that it had
promised.38 Finally, there were almost simultaneous developments of
porosity measurement processes other than resistivity, which made it

38 At that time, Rt itself was not always above suspicion, and it is likely that a failing Rxo/R,
ratio has, sometimes incorrectly, been attributed to a faulty Rxo.

266
After 1943

possible to calculate F through relations like equation 3 above. The


problem of calculating F, or eliminating it between two equations with
the Rxo of the Microlaterolog, became much less stringent.
Years passed before improved electronics, better focusing, and drill¬
ing muds of higher quality restored the Microlaterolog to its deserved
place among the Schlumberger tools. The theory of interpretation set
forth above and its application to the design of the Microlaterolog have
been fully justified by experience.

The nuclear logs: gamma rays, neutrons

If may be safely stated that all the practical systems of resistivity


logging originated with Schlumberger, and, furthermore, that the credit
for these systems and also for S.P. goes to two men: Conrad Schlum¬
berger and Henri Georges Doll.
However, geophysics is more than resistivity. When, after 1945, the
accelerated development of petroleum exploration and production
brought to the staffs of the major oil companies—Esso, Gulf, Mobil, Shell,
Texaco—physicists assigned to the systematic study of such parameters as
could lead to a better evaluation of the oil fields, the policy of the
companies was not to keep a proprietary interest in the results of such
studies, but to license their patents to service companies. When these
patents were offered for licensing, Schlumberger, being best qualified to
make a success of a new technique, was given priority but not exclusivity,
since the latter would have been contrary to American law. This was the
circumstance under which the array of Schlumberger techniques came to
include some developed outside the Company.

Gamma Ray Logs

Natural radioactivity varies from one rock to another. It is very weak


in pure sands or limestones and reaches a maximum in shales, where there
is a concentration of radioactive uranium, thorium, and potassium-40
salts. The energy of the gamma rays the rocks emit is such that they pass
through steel without a great loss: for instance, a typical well casing will
absorb only 30 percent of the radiation.
The log of this radioactivity in an open-hole section will differentiate

267
The History of a Technique

marls and sands as does S.P., but what is important is that radioactivity is
not affected by the nature of the mud, whereas S.P. is generally useless in
salt-saturated mud. Other substantial advantages are, in many cases, the
capability to see through the casing and to record a log in a cased
borehole.
The interest of geologists and geophysicists in radioactivity is nothing
new. In 1909 J. Joly had published an article entitled Radioactivité et
Géologie (Radioactivity and Geology).39 In 1921 Richard Ambronn had
measured the radioactivity of cores taken in an oil well at Celle, near
Hanover, Germany. It is known that before World War II Russian
geophysicists had built a logging device based on gamma rays, but no
results were published. Finally, on October 29, 1938, natural radioactivity
was recorded for the first time, under the name “Gamma Ray Log,” in a
well at Oklahoma City. The apparatus had been designed by a group of
physicists40 of Engineering Laboratories in Tulsa, Oklahoma. With the
support of Socony Vacuum, the group established a service company,
Well Surveys, Inc., which marketed the log from 1939 on.
The detector, based on a rather old technique, was an ionization
chamber, a gas-filled enclosure made weakly conductive by the gamma
radiation. Like all gamma ray systems, it required a number of electronic
components in the sonde.
The results confirmed the theory: whatever the mud, even in a cased
well the logs were comparable with the S.P. Practical applications were
found immediately, especially for the resumption of production in old
wells where no logging had ever taken place, and which could now be
studied by correlating the gamma ray log with S.P. logs in surrounding
wells. The same was true for the control of perforating depths.
Mention has already been made of how difficult it is to measure
accurately the depth of a well and, in particular, to position the perforator
at exactly the same depth where a bed has been identified by electrical
logging under mechanical conditions (tension and elongation of the cable)
quite different from those in the cased well. The necessary correction was
now provided by the clear correlation between the open-hole S.P. and the
gamma ray log in the cased well.41

!il Reprinted in Geophysics (Silver Anniversary Program), Society of Exploration Geophysi¬


cists, 1955.
40 W. G. Green, S. A. Scherbatskoy, R. E. Fearon, L. M. Swift, and J. Neufeld.
41 The method is actually rather complex. Even when a horizon has been located with

268
After 1945

Wartime personnel and equipment shortages delayed the full success


of the gamma ray measurements and depth control technique. Well
Surveys, Inc., merged with Lane Wells, specialists in perforation, and
placed Schlumberger in an awkward position. True, in 1940 Doll had
proposed marking the productive zones with radioactive bullets. Once the
formation had been identified on the electrical log, and before the casing
had been run, a bullet carrying a small amount of radioactive substance was
fired into the formation, using a regular perforator. The bullet was very
accurately positioned by means of an S.P. electrode lowered simultane¬
ously, as for sidewall sampling. In a formation of average consistency the
bullet penetrated deep enough not to be displaced during casing and
cementing. It provided a marker that could be detected through the
casing' and the detector used was simpler than that for measuring the
natural radioactivity of rocks.
The method proved less valuable, however, than had been antici¬
pated. A first problem was that the decision to use the gamma ray log
could be postponed until the very time of perforating, whereas the firing
of the radioactive bullet had to be scheduled before casing was run. The
difference was, in effect, either doing two operations (firing and locating)
or running a single gamma ray log. Also, some additional data could be
expected from the gamma log, whereas the location of markers was
limited to depth control. Finally, as the bullets contained a radium salt
with a half-life of 1590 years, the zone surrounding the marker was made
radioactive and no measurement (gamma ray) of its weak natural radioac¬
tivity could be made.42 Although commercial prospects were never

gamma rays, a residual uncertainty over the depth arises from differences of weight,
volume, and location between the sonde and the perforator. Such uncertainty would be
eliminated if locating the horizon and perforating could take place in the same run. There
are two reasons why this cannot be done: first, it is too hazardous to attach an electrically
detonated pyrotechnical device and a sonde requiring a high voltage at the end of the same
cable; and, second, the recoil of the perforator would destroy the electronic components
of the gamma ray sonde. An intermediate reference is thus required, and it is provided by
the collars of the casing. Since 1948 this reference has been a passive magnetic device,
requiring no current and presenting no hazard. A first detector, fitted on the sonde, locates
the collars on the gamma ray log, and their virtual location on the electrical log is deduced
by correlation with the S.P. A second detector, mounted on the perforator, locates the
collars again. To position the perforator, all that is needed is to interpolate depths between
two consecutive collars, roughly 30 feet apart, with the cable measuring wheel. The
accuracy reached is within an inch.
42 With isotopes currently available today, this objection would no longer hold.

269
The History of a Technique

outstanding, there is one area where the system has prevailed. Whenever
the problem is to detect with an accuracy within a fraction of an inch the
ground subsidence that occurs or may occur in California, Japan, or
Holland as a result of gas or oil recovery from thick unconsolidated beds,
the radioactive bullet is the answer.
By the end of World War II it was obvious that gamma ray logging
had a bright future, especially since a new radioactivity technique was
emerging: “Neutron Logging” with its anticipated contribution to the
evaluation of reservoir rocks. In 1945 an agreement had been concluded
between Schlumberger and The Texas Company for the development of a
gamma ray logging patent. The director of research for Texaco, G. R.
Herzog, a Swiss physicist well known for his work on cosmic rays, was
made available to Schlumberger. His specific plan was to use a Geiger-
Miiller detector for counting the gamma rays. The problem of creating an
electromotive force of some 1000 volts in the sonde while maintaining
good insulation at a bottom-hole temperature of 300 degrees F was
solved, and the system, put on the market in August 1946, proved wholly
satisfactory.
The absolute value of radioactivity has little quantitative use; it may
sometimes be necessary to compare two readings at different levels, but
this requires only proportionality between the deflection of the gal¬
vanometer and the natural radioactivity. Nevertheless, this was not good
enough for Schlumberger, with its propensity for measurements of the
highest quality.
Therefore the Company introduced a calibration technique that
made the log independent of the particular features of the sonde or the
control panel.43 This calibration system was abandoned, however,
when the American Petroleum Institute devised another, which became
standard for all the service companies, even though the Schlumberger
system seemed simpler and less artificial.

The Neutron Log

Together with the proton and the electron, the neutron is one of the
elementary particles of the atom. It is usually associated with the proton in

43 See A. Blanchard and J. T. Dewan, “The Calibration of Gamma Ray Logs,” The
Petroleum Engineer, August 1953.

270
After 1945

the nucleus. Its mass equals that of the hydrogen atom, and it is electri¬
cally neutral.
How the neutron was discovered reads like a “whodonit.” The race
for the truth began at the end of 1930 with Bothe and Becker in Germany
and ended on February 17, 1932, with Chadwick’s triumph at the Caven¬
dish Laboratory. To postulate the existence of the new particle, Chadwick
had used, among others, the results published barely a month earlier
(January 18, 1932) by Frédéric Joliot and Irene Joliot-Curie of the Institut
du Radium in France.
No less extraordinary was the speed with which the adaptation of the
neutron to lithological studies was proposed and achieved. Robert E.
Fearon was granted a patent in 1938, another was granted in 1940 to
Folkert Brons, and in 1941 the Italian physicist Bruno Pontecorvo
published in Oil and Gas Journal an article entitled “Neutron Well Log¬
ging.” Well Surveys, Inc., undertook the first tests that same year.
Meanwhile study of the new particle, whose role in nuclear weapons
could already be foreseen by those with the requisite knowledge, had
taken giant strides. It seems rather surprising that the neutron was consid¬
ered for logging, but the highly penetrative character of a flux of these
particles did augur well for logging through the casing. On the other hand,
as a consequence of neutronic activity, high-energy44 gamma rays that a
gamma ray sonde can detect are emitted. As Bothe and Becker had
demonstrated, neutrons are generated at the rate of several billions per
second by bombarding a beryllium target with alpha particles provided by
a naturally radioactive substance like polonium or radium.
However advanced nuclear physics may have been in 1945, its first
application to logging was highly empirical. It was soon possible, with the
log obtained, to distinguish shales from sandstones or limestones, but for
a rather long time use of the neutron log was strictly qualitative, comple¬
menting the natural radioactivity log and enhancing lithological interpreta¬
tion and well correlations. The gamma-neutron combination somehow
recalled the S.P.-resistivity association, except that it was now possible to
“see” through the casing.
The free neutron, a neutral particle, collides with atomic nuclei
without interference from electrostatic forces. Ejected at a velocity of
thousands of kilometers per second, it collides with various nuclei; these

44 This property distinguishes them from the much less energetic natural gamma rays.

271
The History of a Technique

collisions slow it down to the velocity due to thermal motion at ambient


temperature, or about 2 kilometers per second. It can then be absorbed
by the nucleus of an atom, and this absorption results in the emission of a
high-energy gamma ray, or “capture” ray. The fundamental interest of the
neutron log lies in this slowdown, studied by Fermi as early as 1934. The
neutrons sustain their highest deceleration when colliding with hydrogen
atoms because the latter, having the same mass as neutrons, absorb rather
than reflect energy at each collision. If, on the contrary, neutrons collide
with much heavier atoms such as calcium or silicon, whose mass is about
25 times higher, they “bounce” without any significant loss of energy.40
In a hydrogen-rich medium the neutrons are thus slowed down and
captured in the vicinity of the emitting source, and the gamma rays
generated by this capture will be too far away to be sensed by a rather
remote detector 1 to 2 feet away. On the contrary, if the medium is poor
in hydrogen, the neutrons will proceed further before being captured, and
a large number of gamma rays will be emitted in the vicinity of the
Geiger-Miiller counter. In other words, the number of gamma rays
counted varies in inverse ratio to the hydrogen concentration.46 In the
formations drilled through by a borehole, the presence of hydrogen is
almost solely caused by the fluids, and consequently the amount of
hydrogen per unit volume of rock is a function of porosity. The latter can
be evaluated by counting the gamma rays of capture if the pores are filled
with oil or water, both of which have about the same molecular concentra¬
tion in hydrogen. The same is not true, however, of gas, the hydrogen
concentration of which is much lower; a gas-bearing horizon will there¬
fore appear on the neutron log as less porous than if it were filled with oil
or water. With the help of an additional porosity measurement (core,
sonic or density log, etc.; see pp. 307 and 304), it thus becomes pos¬
sible to distinguish not only between gas and water—which resistivity
already does—but also between gas and oil—which resistivity cannot do.
In addition to being a significant parameter characterizing the reser¬
voir, porosity makes possible the determination of the formation factor
(Archie formula) and, provided that the resistivity (Rw) of the interstitial
water is known, the deduction of the resistivity (R0) of the 100 percent-
water-saturated rock. If the electrical log gives a resistivity value of Rt

45 This is only a rough mechanical image of a phenomenon belonging to nuclear physics.


46 There are processes to evaluate the slowdown of the neutrons other than counting the
gamma rays of capture (see p. 304).

272
After 1945

higher than R0, the rock must, in principle, contain hydrocarbons in a


proportion that can be obtained from the ratio Rt/R0-
The importance of the neutron log becomes clear from these re¬
marks. The slowdown of free neutrons is a function of the hydrogen
concentration, which in turn is a function of the porosity. Knowledge of
this porosity at every point makes it possible to determine at any depth
the excess of the measured resistivity over what it would be with 100
percent water saturation and thereby to establish the presence of oil and
quantify it percentagewise within the pores. With this percentage, and
with porosity once more intervening, the volume and value of the oil in
place are known.
Those47 with the necessary background were aware of such poten¬
tialities, but realization had to wait for several years. The practical difficul¬
ties were great. It soon became apparent that not only porosity, but also
variations in the borehole diameter, the position of the sonde, and the
nature of the mud substantially affected the results. The need to obtain
identical logs irrespective of equipment mandated calibration, and since
this had to allow for particular features of the sonde as well as the source,
it was always a delicate and sometimes uncertain operation.
Schlumberger entered the field in 1949 with an ambitious
$2,000,000 program. The problem was to catch up with the competition,
design precise and dependable instruments, set up a first-class service, and
integrate the results of the new log into the general technique of interpre¬
tation. A group of experts was hired (including John Dewan, Denis
Tanguy, and Jay Tittman), who quickly made the basic decisions as to the
source,48 the type of detector, the spacing, and other relevant features.

47 The relation between porosity and neutron slowdown had been considered in 1949 by
Clark Goodman, a technical consultant for Schlumberger. In the same year R. Fearon
referred to it in Nucleonics (June 1949). One of the first studies of Jay Tittman when he
joined Schlumberger in 1951 was to tackle the theoretical basis of this relation.
48 Schlumberger initially decided in favor of radium-beryllium, although this combination
had serious shortcomings as compared with polonium-beryllium: the price was very high,
and the gamma radiation required thorough protection of the operators. However, its
half-life period is 1590 years against only 140 days for polonium. Whereas in the United
States radioactive sources could be delivered within a few days, the situation was different
when customs formalities interfered. The delays for polonium could then exceed one or
two longevity periods, and a problematic recalibration was required, if the product had not
become altogether useless. The selection of the source was therefore determined by the
fact that Schlumberger’s operations were international.

273
The History of a Technique

Special laboratories were established in Ridgefield and Houston for han¬


dling radioactive materials and designing the tools. A system of safeguards
and tests had to be organized to comply with U. S. Atomic Energy
Commission standards. Periodic inspections of the sources detected any
possible leaks in the stainless steel block in which they were sealed.
Instructions were given on fishing jobs, should the source be lost in a well.
Schlumberger’s version of the neutron log was introduced in the
United States in 1952. From the very outset it proved comparable in
quality to competitive services, especially since it had the advantage of the
field engineers’ competence in interpretation. Observations made in 1953
in the Permian Basin provided practical data on porosity determination.49
Concerned with giving the client the means to determine a physical
parameter irrespective of the conditions under which the measurements
were made, Schlumberger published correction charts that are to the
neutron log what departure curves are to resistivity.50 Moreover, by using
two logs with different spacings between source and detector (which gave
different depths of investigation), it became possible to identify gas¬
bearing horizons. The method was especially successful in eastern Ven¬
ezuela, thanks to the outstanding teamwork of Pierre Majani, Frank
Dennis, and Mike Grosmangin.51
Neutron logging developed steadily between 1952 and 1957, even
though it did not fulfill the hope that it might become a universal porosity
measurement method. In addition to the difficulties already listed, there is
a rather large lithological effect, foreseeable but complicating interpreta¬
tion, and in areas where the lithology was known empiricism often pre¬
vailed over rational procedure.
At Schlumberger, as elsewhere, nuclear physicists had much earlier
proposed and tried to reduce the lithological effect by other methods than
counting the gamma rays of capture to detect the slowing down of the
neutrons. Such methods were developed later, and neutron logging
realized its full value only when it became integrated into a set of mea-

49 The linear relation between the logarithm of the porosity and the neutron log reading
had been recognized by Shell's S. H. Rockwood, who had conducted many comparisons
between porosity values measured on cores and neutron log readings.
50 J- V Dewan, “Neutron Log Correction Charts for Borehole Conditions and Bed
Thickness, Journal of Petroleum Technology, February 1956.
M. Grosmangin and E. B. Walker, “Gas Detection by Dual-Spacing Neutron Log in the
Greater Oficina Area, Venezuela,” Journal of Petroleum Technology, May 1957.

274
After 1945

surements the interpretation of which was so complex as to require the


help of the computer. In effect, from 1958 on the neutron log was no
longer the only one of its kind: for a time the sonic log and then the
density log rounded out the array of porosity tools.

Auxiliary operations

Dipmeter Surveys

The S.P. correlation dipmeter often gave excellent results. From the
outset it almost always proved superior to the electromagnetic anisotropy
dipmeter, as it indicated not only the azimuth but also the dip angle of the
strata.
The operating technique was to select on the conventional log a
number of sections, each about 30 feet long, such that the dips would be
geologically significant and the S.P. curve would allow correlations. The
selection was usually made in collaboration between the client’s geologist
and the Schlumberger engineer. Computation of the dip requires knowl¬
edge of the inclination of the borehole and the orientation of the instru¬
ment, and these data were given by the photoclinometer kept in fixed
geometry with the dipmeter sonde. Once the tool had been lowered to
the deepest level of a section selected for study, it was necessary to wait
for about 1 minute until the needle of the compass and the ball of the
inclinometer stabilized and could be photographed. While the sonde was
then slowly pulled to the upper level of the section, the three dipmeter
curves were recorded. There was then another stop, another wait, and
another photograph. As a comparison of the photographs taken at the two
levels almost always showed that the instrument had rotated during the
recording, computation of the dip was possible only on the assumption
that the rotation had been proportional to the vertical travel. This is the
reason why the length of the sections was limited: a substantial rotation
would have required an exceedingly hazardous interpolation.
Correlating the dipmeter curves was theoretically easy, but actually
required great skill. Wherever the dipmeter operated, there were
specialists analyzing the logs received from the field, correlating curves,
making trigonometric calculations to determine the azimuth and dip
angle, and preparing a graphic presentation of the results. These

275
The History of a Technique

specialists were recruited from among draftsmen, assistants, and en¬


gineers removed from the field for reasons of health or especially hired
for the job (there were a few women in the group), and some became
outstanding experts. The quality of the service provided was due to their
competence as well as to the proper selection of the sections and the
accuracy of the field measurements. Such a strong organization, along
with highly specialized teleclinometer equipment, gave Schlumberger a
practical monopoly in dipmeter operations.
Southern Louisiana was the area of greatest activity and success for
the S.P. dipmeter. In this region of salt dome tectonics, where knowledge
of the dip is essential, the S.P. curves are excellent and highly diversified,
and very sharp deflections yield unequivocal correlations.52 Doll’s study
on the S.P. specified the conditions to be met for sharp deflections: what
is needed is not only good resistivity contrast between drilling mud and
interstitial water, but also permeable formations and shoulders of rather
low resistivities. Such conditions also prevail in southern Texas, Califor¬
nia, eastern Venezuela, Nigeria, and Indonesia. On the other hand—
either because the drilling takes place with saltwater mud (Kansas, Per¬
mian Basin, offshore), or because the resistivities are high (northern
Louisiana, northern Texas, Oklahoma)—there are many regions where
the S.P. dipmeter works inadequately, or not at all.
There is also a geological factor that renders the S.P. inadequate as a
universal dip-measuring parameter. The sharpest deflections (i.e., those
giving the least equivocal correlations) are the ones marking the sand-
shale contacts. Yet it often happens that these contacts, which are in a way
stratigraphic accidents, are not in the plane of the structural dip. On the
contrary, while nicely layered shales allow for excellent dip values, S.P.
does not distinguish one layer from another in a typical shale stratum, so
that the S.P. dipmeter is blind to the most significant sections. This is the
reason why, despite its complexity and limitations, the electromagnetic
dipmeter, extremely sensitive to the anisotropy of the shales, continued
to operate for years in certain regions.
From 1945 on, the above considerations determined the path to
follow: resistivity curves were necessary to obtain universally good corre-

52 At the depth scale of 1:20 generally used in these recordings, the maximum theoretical
curve displacement for a 5 degree dip in an 8 inch well is less than 0.7 millimeter. This
already requires precision in evaluation, but the dips to be measured and the displace¬
ments are sometimes even smaller.

276
After 1945

lations. Again, the solution came from Doll. He conceived an apparatus


that combined three lateral sondes with a single common-current elec¬
trode. The measure electrodes M and N were placed in pairs on three
arms at 120 degrees, with AIN spacing of about 1 inch. The current
electrode A was on the lower part of the mandrel, 3 feet from the plane
defined by the centers of the three MN pairs. Three detailed resistivity
logs, suitable for correlations, were thus anticipated.
When this sonde was designed, only the textile-braided quadricable
was available, and it was necessary to work out simple53 and dependa¬
ble bottom-hole electronics so as to record the three resistivity signals
independently. The results were immediately conclusive. Excellent corre¬
lations, could be obtained from the resistivity curves in areas where
nothing could be expected from S.P. In southern Oklahoma, where the
initial tests took place in 1946, clients reacted positively to the new
service. From January 1947 it was accepted throughout the region and its
use was widespread. From there it promptly expanded first in the United
States and then to Venezuela and Trinidad. Beginning in 1952, resistivity
dip measurements became a standard, worldwide procedure.
More improvements were desirable because of the importance of
selecting the proper sections and their boundaries. Should these selec¬
tions be made a priori, or would it not be preferable to record the whole
open hole and then select the sections after examination of the results?
However, this was not possible at the time because of the stops required
for the operation of the photoclinometer. This device therefore had to be
replaced by a type of inclinometer that functioned continuously, if possi¬
ble, or at least required no stops in the upward movement of the sonde.
From 1940 to 1942, during the German occupation, the Paris center
had worked out a photoclinometer, fluid filled and consequently mechan¬
ically damped, whose elements could be photographed without interrupt¬
ing the upward motion. A gas bubble moving inside a liquid under an
upward-convex glass replaced the steel ball of the cup. Without interrupt¬
ing the motion of the sonde, as many pictures could be taken as the film
magazine permitted. In spite of such minor troubles as the formation of

53 The inclination of the borehole and the orientation of the electrodes were still obtained
by the photoclinometer, whose operation remained unchanged. A relay made it possible
to switch from the dipmeter to the photoclinometer function. With electronics the two
could have been performed simultaneously, but since the mechanical system required a
stop to photograph the needle and the ball, simultaneity was not essential.

277
The History of a Technique

secondary bubbles, the instrument could have become operational except


for the fact that it used film. Intermittent inclinometry data constituted no
major drawback provided that the interval between photographs could be
reduced. The major limitation was the sensitivity of the film to tempera¬
ture. Whatever the precautions, it frequently happened that, after operat¬
ing in a deep well, the film was found to be completely blackened; this was
especially exasperating since it was discovered only after the operation. It
was better to forego the film and transmit the inclinometry data to the
surface, a practice that, in addition to actually yielding the required
information permits monitoring the instrument behavior.
The first device meeting this requirement was realized in Ridgefield
in 1951, and a small series of 10 was put into operation in the United
States in 1952. Called the C.D.M., for “Continuous Dipmeter Microlog,"
it became known as the “de Chambrier dipmeter”54 or “The Thing" by
field engineers marveling at such complex and novel electronics. Indeed it
contained a fluxgate compass borrowed from the aircraft industry, and an
elementary electronic computer which, by combining the readings of the
compass and the inclinations of two pendulums mounted at right angles,
calculated the north-south and east-west components of the borehole
deviation. As to the dipmeter curves, a considerable advance had been
made by using three Microlog arrays, as their vertical resolution was much
better than that of the old resistivity curves and permitted more correla¬
tions of higher accuracy. The system was further perfected in 1954 with
the replacement of the Micrologs by Microlaterolog-like curves. Also, the
mechanics of the sonde were improved: instead of fixed electrode sup¬
ports selected according to the nominal diameter of the borehole, there
were now three mobile pads applied to the borehole wall by springs, thus
making it possible to record simultaneously the diameter variations of the
hole.
However ingenious, the teleclinometer lacked dependability, was
very difficult to maintain, and required bulky and complex surface equip¬
ment. These drawbacks did not prevent the de Chambrier dipmeter from
being accepted by clients, proof that continuous dipmeter surveying was
the solution of the future, and that discontinuous measurements were no

54 Pierre de Chambrier was an engineer and electronics specialist who had expended great
effort on dipmeter problems, especially during the several years he had spent in eastern
Venezuela.

278
After 1945

longer acceptable in regions where knowledge of the dip is particularly


important.
At the same time, the Paris center had begun to develop a tool
capable of similar performances, but mechanical in nature. The initiative
came from Marcel. In a way it could hardly have been otherwise, for the
device was so innovative that nothing less than his full technical and moral
authority had been required to launch the project. The guiding idea was
to measure the rotations of the deviation pendulum, the relative bearing
pendulum, and the compass enclosed in the teleclinometer by having
them actuate rotating potentiometers designed for minimal frictional
torque. Imagine a short magnetized bar oriented in the terrestrial mag¬
netic field and actuating a potentiometer: the torque is so weak that it
does not seem able to overcome the friction of the cursor on the rheostat.
Yet in 1950 the Swiss firm Ohmag was already making potentiometers
that a compass needle could actuate. Conventional measurement of the
resistances of the potentiometer easily gives the position of the cursor,
and thus the angle between the magnetic north and the direction of
reference of the instrument. What is true for the torque of the compass
also applies to the torques of the deviation pendulum and the relative
bearing pendulum when they are only 1 or 2 degrees off their positions of
equilibrium, and thus their slightest angular displacements can be mea¬
sured. These are the elements of the teleclinometer that was called the
“Poteclinometer.”
The instrument was tested in Pau in 1953 and then brought to the
United States and Venezuela in 1954 for intensive trials. Edmond
Boucherot and Bob Canup adapted to the poteclinometer the three-pad
Microlaterolog sonde used for dipmeter curves. In October 1954 Rene
Roussin in Paris resumed study of the mechanical design, which had made
little progress since the death of Marcel in 1953. In cooperation with
Ohmag, the potentiometers, under the name “Micropotentiometers,’’
were improved and manufactured by Schlumberger under a licensing
arrangement. The friction effect was further minimized by a small motor,
giving a tacking motion to the whole mechanism. Its sensitivity was such
that the compass had to be developed in the open countryside because the
Paris subway trains deflected it by some 10 degrees.
From 1957 on the instrument was produced in its final form (Fig. 77).
It not only replaced the de Chambrier instrument in the United States,
but also, since it no longer required highly specialized operating per-

279
The History of a Technique

Figure 77. Poteclinometer.

280
After 1945

sonnel, soon supplanted the photoclinometer and the resistivity dipmeter


all over the world.

Thermometry

The battery cell thermometer had been employed in the field since
1938, mainly for thermal location of the cement tops in cased wells (see p.
179), and had given rather satisfactory results. Another application, less
dependable because multiple intervening factors made any quantitative
interpretation, even in broad figures, very difficult, was the detection of
gas entries in the mud column. Certain cases demanded the study of
minute temperature variations. However, the technique proved reason¬
ably successful in the Permian Basin, where gas entries are important.55
On the other hand, it was used to excess in Kansas to detect low-potential
and low-gas : oil-ratio oil-bearing strata. Oddly enough for a thermome¬
ter, it had temperature limitations. The bottom-hole temperature induced
a rather sharp drop in the voltage of the cells enclosed in the sonde, and
in the potential variation measured at the surface it was difficult to
distinguish between the change in thermometric resistance and the drop
in battery voltage.
To achieve more dependable results, Boucherot undertook in 1948
the study of a “Deep Well Thermometer.” He produced an instrument
that withstood the highest bottom-hole temperatures and was more sensi¬
tive, quicker to operate, and less bulky than the previous model. The
solution was a bold one. The thermometric wire was dipped bare into the
drilling mud—hence a minimum calorific mass and a quicker response.
The problem had been that, in order to render negligible the leaks
through the mud, a low thermometric resistance was needed. As the trend
of the time was to use resistivity logging circuits for all purposes, the
resistance bridge constituting the thermometric sonde was fed with pul¬
sated current. The signal was amplified by placing at the head of the sonde
one of those unorthodox square current transformers already described in
connection with the Laterolog. With no substantial change in operating
method or interpretation of results, the deep well thermometer was put

55 The problem was to produce a gas-free oil, as at that time gas had no commercial value
and was flared on the spot. However, reservoir conservation rules strictly limited the
gas : oil ratio. Only the reservoir energy of the gas was involved, but it remained essential
to locate its entry.

281
The History of a Technique

into operation in 1950 and lived up to expectations. Even though the


thermometric wire had to be insulated by a special coating for operations
in highly conductive muds, its basic qualities were such that it led to
consideration of a type of production logging to be discussed later.

The Borehole Caliper

It may seem strange that a tool to determine borehole diameter had


not been produced by Schlumberger at the outset. To measure a geomet¬
ric value by an electrical method would have been well within Marcel’s
talents. Perhaps he was too remote from the field, and the engineers did
not call his attention to the significance of the problem. At any rate the
first caliper was put into operation in 1938 by Halliburton, a leader in the
cementing of casing. It had been proved by thermometric measurements
that the level to which the cement rises in the annular space between
borehole wall and casing is always less than it would be if the borehole
diameter were equal to that of the drill bit. As such equality is nearly
always prevented because the borehole is exposed to erosion by the
circulation of the mud and the whipping of the drill pipe, the annular
space must be accurately known to determine the amount of cement to
reach the desired level. The war delayed the success of the Halliburton
caliper until 1945.
The caliper put on the drawing board by Schlumber in 1946 reached
the market in January 1947 under the name “Drill Hole Gauge.” The
sonde utilized three long, flexible spring bands that contacted the
borehole wall, their lower ends fastened to a rod free to slide inside two
coils on the sonde mandrel. The spring band diameter (hole diameter) was
calibrated in terms of rod position, which, in turn, was a function of the
coupling effect between the two external coils. Here again conventional
resistivity logging circuitry was used. The primary coil was fed by a
sawtooth alternating current produced at the surface from pulsated cur¬
rent by a combination of inductance coils and capacitors; accordingly, the
output of the secondary coil was a pulsated voltage56 easily accommo¬
dated by the conventional resistivity measurement system. Since the
inductive system removed the need for any mechanical contact, the design

>h The induced voltage is proportional to the derivative of the inductive current with
respect to time, and the derivative of a sawtooth curve is a square-pulse curve.

282
After 1945

was very simple. Furthermore, by varying the metallic assembly of the


rod, one could make the signal proportional to either the diameter or its
square. In the latter case, the log could be read in Trd2/4, in other words, in
terms of the area of the supposedly circular cross section. A mere integra¬
tion of the curve as a function of depth then gave the volume of the
borehole. However ingenious this solution, clients preferred a linear
scale. The presentation most in demand, the so-called “Mae West,” con¬
sisted in tracing two symmetrical curves, which gave the impression of a
cross section of the borehole by a plane passing through its axis.
Later experience demonstrated the need to record the borehole
diameter simultaneously with certain measurements (Microlog, Micro-
laterolog, dipmeter, etc.) so as to better control their quality. Although
this recording, which aimed solely at the sonde-borehole geometry, was
less dependable than results obtained with the caliper proper, it often
gave a first approximation of the volume of the hole and of the expected
depth of the cement top. From that point on, special caliper mea¬
surements were reserved for problem cases: unusual borehole shapes,
large diameter increases, the need for exceptional precision, and so on.

Sampling

In the southern United States and eastern Venezuela especially, the


Sample Taker or Sidewall Sampler had been remarkably successful since
its introduction in 1936.
The recovery rate had been considerably increased by the substantial
improvements made over the following decade in the bullets, retrieving
wires, and loading methods.57 However, geologists and engineers were no
longer content to look at the cores, smell or lick them, or even examine
them with a steroscopic microscope or under ultraviolet light. They were
now demanding a more systematic analysis of the samples, particularly for
porosity and permeability. Certain laboratories provided such mea¬
surements, but these still had to be interpreted taking into account the
variation sustained by both parameters under the impact of the bullet on
the borehole wall. Furthermore, there was agreement among specialists
that the samples were too small for a complete analysis that would include

57 Up to 80 samples for every 100 shots fired was the recovery rate in favorable areas.

283
The History of a Technique

paleontology, clay percentage, oil saturation, oil composition, and other


relevant factors.
The problem was tackled in Houston and Paris simultaneously. After
many discussions that, incidentally, were indicative of the kind of compe¬
tition existing between the two centers, the Paris model prevailed.58 This
was the last major contribution of Marcel, who had not conceived it, but
had pursued its realization step by step. The C.S.T. (“Chronological
Sample Taker”) could take 30 samples in each run, against 18 for the old
sidewall sampler, and their volume was 2.5 times greater. What was really
new was the loading and igniting system, devised by Boris Schneersohn
and Jean Planche (Fig. 78). Igniter needles, already tested on the bullet
perforator and since extended to other models, allowed for the use of
cartridges whose powder, carefully controlled as to weight and moisture,
ensured a uniform fire power. Selectivity was achieved in an original
manner: the firing of each shot armed the following one, thus eliminating
any ambiguity as to the chronology of the firings. To this end, Schneer¬
sohn and Planche had first considered utilizing part of the deflagration
gases somewhat like an automatic weapon; but Marcel raised the objec¬
tion that the gas ports would be quickly eroded, and proposed a solution
of his own: a steel wire, pulled by each bullet fired, armed the following
one. Improvised as this solution appeared, it proved to be the right one.
A testimonial to the quality of the C.S.T. is the fact that it has
undergone no major change since it was put into operation in 1951 in the
United States and Venezuela. Only accessories—bullet bottoms, wires,
and igniter needles—were gradually perfected to the point of ensuring
optimum performance in areas best suited to sidewall sampling. However,
despite displays of ingenuity and progress in steel metallurgy, sample
taking remained as unsuccessful as in the past in hard formations such as
sandstones or limestones. The conviction finally prevailed that percussion
sampling was impossible in such formations, and that the problem had to
be attacked from another angle.
As early as 1947 Schlumberger had offered the service of a mechani¬
cal sampler. It had been conceived, designed, and built in Houston by
Maurice Mennecier, an enthusiastic and touchy perfectionist, whose basic
idea was simple: a hollow-core barrel with a diamond drill crown was
driven by an electric motor, and carved in a borehole wall was a core 1

58 The Houston model had to be limited to small-diameter wells.

284
After 1945

*-/ ^} CM. (*• Xct -â

A*. A-C^ UKÜ. UA-vJ

.X> «h* 4-Ae. ^ * G-u. ÿ A/W" A ^

J^/-o X ^Un &A< .

/ '

Figure 78. Sketch by Marcel Scklumberger of a proposed needle-igniting system.

285
The History of a Technique

inch in diameter and 2 inches long. Practical difficulties were great:


anchoring, feed, core recovery, and so on. Only a few units of this
ingenious mechanical prodigy were built; they met with some success in
the sandstones common to northern Texas and southern Oklahoma.
However, the complexity of the device, along with the skillful handling of
the winch required, and chiefly the difficulty of sampling in limestones
where porous samples disintegrate, prevented wider use. Manufacturing
was discontinued, and about 1955 the service ceased. Not until 1970 was
the concept of a mechanical sidewall sampler resumed in another form.
Another kind of sampling should be mentioned: the recovery of fluid
samples from the pores of permeable rocks. Several methods are conceiv¬
able.
In 1951-1952, at Marcel’s behest, study was undertaken in Paris of a
“tester” or “suction” bullet. Fired by the sidewall sampler, it was to take a
sample of the fluid through a system of instantaneously opening and
closing valves actuated by the impact on the formation. In 1952 Jacques
Delacour conducted a few successful tests in the United States, but the
project was shelved when it appeared that the few cubic centimeters that
could be tapped under the most favorable conditions were inadequate for
any thorough diagnosis.
To increase the volume of fluid recovered, another device had been
studied. This consisted of firing into the formation a bullet retained by a
flexible tube connected at the other end to a container in which, through a
system of valves, the liquid could accumulate. From similar projects
undertaken in Paris and Houston, two instruments were born. The first,
marketed between 1955 and 1957 under the name “Fluid Sampler,” could
take five samples of 330 cubic centimeters in a single run. This was
deemed inadequate, and the other, designated as the “Formation Tester,”
prevailed. It took only one sample, but the volume exceeded 5 gallons.
The formation tester (Fig. 79) involved a flexible pad forced tightly
against the permeable formation to be tested. Once it was well sealed, a
bullet or the jet of a shaped charge pierced the pad and penetrated the
formation, thus connecting it with the sampling container. All the stages
of the operation were recorded at the surface, particularly the flow line
pressure versus the sampling time. Thus, in addition to a good sample of
the fluid, these data, which under certain assumptions permit calculation
of the formation permeability with an acceptable approximation, are
obtained.

286
Figure 79■ Formation tester.

287
The History of a Technique

Of all the instruments of its kind only the formation tester has
survived to this day. However, in spite of many alterations and improve¬
ments, it remains of limited use because of its mechanical complexity,
operational difficulties, and maintenance requirements. It has found cer¬
tain application in southern Texas, in Argentina, and in Nigeria, but has in
no way eliminated the drill stem test, which requires a more demanding
and hazardous endeavor but yields much more comprehensive results.

Perforation

The perforation technique had barely evolved during the period


from 1938 to 1945. The only noticeable change in Schlumberger field
operations was the introduction of firing by bursts, which was more in the
nature of a palliative of the shortcomings of the igniting system than actual
progress. In 1948 Schlumberger, as well as its competitors, became
equipped to lower and set bridge plugs59 in the casing. Until then, the
operating companies themselves had set the plugs, lowering them at the
end of the tubing or the drill pipe. However, since the operation very of¬
ten took place immediately before or after perforating, it seemed logical
to take advantage of the cable to save rig time and improve setting depth
accuracy. To enter the field, Schlumberger made an agreement with
Baker Oil Tools, a manufacturer of bridge plugs and associated setting
equipment.60 In 1949 the encouraging reaction of clients led to augment¬
ing this service with a more delicate operation: the setting of production
packers.61 Since in either case the problem is to lower an assembly almost
as wide as the inside diameter of the casing, the operation can be jeopar¬
dized by the slightest (casing) burr. Unfortunately any gun perforation
produced a burr; hence it was out of the question that a packer or a plug
be run through a perforated section, and this made the service much less
attractive. The problem was put to the Paris center, which was then in
charge of all perforating developments. The solution was quickly found,
but it took time to render it fully operational. It was to place in the gun, in
front of the bullet, a small “no-burr” plate, thanks to which the perfora¬
tions were as neat as if they had been made with a drill. Penetration was

59 A plug to seal a borehole at a given depth.


60 Sealing results from submitting the packing of the plug (hardened elastomer) to very
high stresses, generated in the borehole by the combustion of a special slow powder.
61 A packing gland between casing and tubing.

288
After 1945

not affected, because the energy consumed to perforate the plate was
compensated for by an increase in pressure build up and muzzle velocity.
More serious problems, however, soon arose. One was that the
duration of operations involving hundreds or even thousands of shots had
to be shortened; more importantly, the bullet penetration in consolidated
sandstones and limestones was inadequate, and there was need for in¬
creased firing power. The first problem was one of accessories, of proper
organization of the loading shop, and of better personnel training, rather
than of the perforator itself. This aspect concerned Kuwait, Iraq, and
Venezuela more than the United States, where large shooting jobs were
few and, in any case, not handled by Schlumberger. The operational speed
was first increased in Kuwait. A truck-mounted shop equipped to reload
the guns on the spot as shooting went on was driven to the well site. The
layout was excellent, and every convenience was available: guns were
assembled and disassembled in record time by pneumatic tools. The
tightest safety measures were enforced, for with perforators firing in burst
any accidental ignition could have meant catastrophe. This equipment had
been designed and built in Paris by Marcel, but its operation in the field
was conducted by, among others, Louis Bordât, Mike Grosmangin, Joe
Moffet, Jacques Priotton, and Ahmud, the Kuwaiti foreman who was
unanimously credited for the liaison achieved with the local manpower
and for the high efficiency obtained.
The second problem, penetration, became especially acute in 1948.
The inadequate penetration in consolidated rocks was revealed not only
by tests made on cores, but also by the fact that, in bailing the bottom of
the hole, drillers would retrieve parts of bullets, and even whole bullets,
that had not penetrated the casing. On the contrary, in sand-shale series,
perforation at the correct depth resulted in oil production if the cement¬
ing had been properly done, and in water production otherwise. The
efficiency of the perforator was obvious. Yet in 1948 Gulf developed a
new cementing process through which the seal between casing and forma¬
tions was substantially improved, with the result that even in sand-shale
series penetration did not always suffice to bring the well in.
By that time the service companies had already undertaken the study
of more powerful perforators. Schlumberger had been working since
1945 on a new gun, and in 1947 entered the field of shaped charges—an
area where others had a substantial lead. The oil companies themselves
were increasingly unhappy with conventional tools and submitted the

289
The History of a Technique

various perforators to surface tests. These were not very meaningful in


absolute value, since surface and bottom-hole conditions were quite unre¬
lated. Nevertheless, they did allow a rough comparison of the various
tools. During one of these tests Schlumberger had one of the poorest
scores, and something spectacular had to be done. Even though its new
gun perforator was not in final shape, it was advanced enough for a surface
demonstration, and in December 1948 Planche was dispatched to Hous¬
ton for this purpose.
The new perforator, to be known later as the “Supergun,” was
technically highly elaborate. Theory had suggested, and experience
confirmed, that during the fraction of a second that followed firing the
stresses developed during propagation of the shock wave were above the
elastic limit of the alloy steels from which the body of the perforator and
the gun barrels were built, but without any resulting damage. The threads
had been given a special profile; the powder load density was very high;
the seal plugs were integrated with the bullets; and the firing was actuated
by the needle system of the sidewall sampler gun, a design that had made
it possible to reinforce the bottom of the chambers. Whereas the bullet
velocity was formerly 750 meters per second, with the supergun it
reached 1000 and ultimately 1500 meters per second.
The supergun was an outstanding success, first in Venezuela, where it
was introduced in 1949, and then in the United States in 1950. It was
initially used as a special tool in conjunction with the conventional per¬
forator but gradually supplanted the latter throughout the world.
Indeed, it was finally surpassed by shaped charges, but not before the
1960’s. As to penetration, for a long time it could successfully compete,
especially in sand reservoirs, where it had the advantage of fracturing the
cement behind the casing and thus enhancing production.62 Operation¬
ally, however, shaped charges were much more convenient, and the
supergun was actually the ultimate in what could be achieved with a bullet
perforator. Only marginal improvements resulted from further intensive
engineering work, whereas with shaped charges the prospects seemed
unlimited.
The technique of shaped charges derives from the armament indus¬
try. Its origins lay in an 1856 discovery by Munroe that a cavity in an

62 If perforation takes place close to the water-oil contact, such fracturing may also cause
water intrusion. But if the firing is positioned much above that contact, the advantage is
such that it justifies the development of fracturing shaped charges.

290
After 1945

explosive charge concentrates the energy of deflagration (the Munroe


effect). How to multiply this effect was discovered a few years later: when
a conical cavity is dug at the end of a cylindrical charge and lined with a
thin metal, the hole in the target will have a narrow opening, but a much
greater depth. Analysis of this phenomenon shows that the shock wave
disintegrates the cone, and that the particles of the metallic lining act like
a jet with a velocity of 6000 to 10,000 meters per second, a velocity at
which the developing pressures are far above the mechanical resistance
characteristics of the best steels. Under the impact the target behaves like
a plastic body.
Great enthusiasm was aroused by the discovery: the armor of iron¬
clad vessels seemed obsolete. Disappointment followed; under water,
shaped charge torpedoes did not work because the jet formed only when
the cavity was filled by a compressible fluid such as air. Not until World
War II did the proliferation of armored vehicles bring about, on both the
Allied and the German sides, the shaped charge projectiles, known as
Panzerfaust or bazookas, that could pierce the heaviest armor. Thinking
then turned naturally to a similar solution for the perforation of casings.
The only difficulty was the need to maintain air in the cavity and hence to
use sealed containers. This feature was tightly protected by patents.63
Preliminary tests were conducted in Texas in 1946, but not until 1947 did
a major development take place with the founding by Robert H. McLe-
more and R. C. Armstrong of the Welex Jet Perforating Company and the
manufacture of the charges by du Pont de Nemours. The first commercial
perforation took place in Mississippi to the advantage of Gulf Oil Corpo¬
ration.
Great improvements in the charges themselves have since been
made, but their operation remains essentially unchanged. The charge
carrier is a steel cylinder (about 6 feet long and 4 inches in outer diame¬
ter), in which threaded openings have been machined at intervals. The
charges are housed in the cylinder, the base of the cone facing the
opening, into which is screwed an aluminum or steel plug that will be
perforated and destroyed by the firing. The charges are generally distrib¬
uted on a helix so as to have four perforations per foot in four azimuths at

63 The first patent of Clyde O. Davis (1942), licensed to du Pont de Nemours, did not
include this feature; but that of Muskat (Gulf, 1945) specified that the cavity of the cone,
lined with a ductile metal, is filled with a compressible fluid.

291
The History of a Technique

90 degrees from each other. They are simultaneously ignited by a detonat¬


ing cord, which is itself primed by a standard blasting cap (Fig. 80). Each
firing is a burst of 24 to 30 shots. As nearly all the energy of the explosive
is concentrated in the perforating jet, the stresses sustained by the charge
carrier are much lower than those in the bullet perforator. The result is
much less weight and easier handling, loading, and access to the well.
Provided that the operation requires a substantial number of holes and
that the spacing of the openings in the charge carrier is suited to the type

292
After 1945

of production contemplated, the shaped charge perforator is an almost


ideal tool.
Remarkable performance was attained as soon as the shaped charges
became operational. Surface demonstrations, much easier than with the
conventional perforator, showed that mild steel targets 4 inches thick
were pierced through and through, and that penetration in hard
sandstones or limestones was at least 2.5 times that of a bullet. Under the
circumstances the bullet perforator, in spite of such advantages as flexibil¬
ity of bullet-by-bullet firing, fracturing of the cement, and lesser cost, was
doomed.
The development of shaped charges had caught Schlumberger flat-
footed. At a time when the Paris center, still in the process of postwar
reorganization, was undertaking responsibility for the whole perforation
technique, new choices were already emerging in the United States.
However, while still concentrating its efforts on the supergun as a tool
suitable for difficult operating conditions, Schlumberger had the sense to
keep abreast of what was going on elsewhere in the field. In 1946 a study
was commissioned from the French Direction des Poudres, which in¬
cludes leading experts in the field of explosives, and somewhat later
Schlumberger entered into an agreement with the Brandt Company.
These steps enabled Boris Schneersohn, Alexis Venghattis, and Claude
Baks to realize, in 1947, their first shaped charges. In 1948 Schlumberger,
in collaboration with the French-German Laboratory in Saint-Louis (near
Basel, Switzerland), undertook basic research from which it acquired the
capacity to develop and make its own explosives.
The first shaped charge perforations by Schlumberger took place in
Kuwait in 1949. The results were so conclusive that bullet perforation
soon became obsolete. This experience was repeated in 1949 in Ven¬
ezuela, where competitors with their newly imported shaped charges
were giving the supergun a tough time. Finally, in 1950, once the question
of patents had been settled, Schlumberger was the only company in the
United States to bypass du Pont de Nemours, the exclusive supplier of all
other perforating companies.
Schlumberger’s original contribution was all the more appreciated
since a special engineering section had just been established in Houston
with one of its first objectives that of adapting perforation to a new
technique which the major oil companies, Esso in particular, were de¬
veloping under the name “Permanent Well Completion.” This production

293
The History of a Technique

technique consisted of arranging, from the outset, for all completion and
work-over operations involved in the lifetime of an oil well (water shutoff,
recementing, reacidization, reperforation) to be feasible without inter¬
rupting production, that is, without “killing” the well by pumping heavy
mud, and then removing the tubing. To make perforation an integral part
of this new technique required not only a small-diameter perforator but
also a cable small enough to pass through the packing gland of the
blowout preventer. The quick solution was to scale down the 3§ inch
shaped charge so that it could be used in a lfi inch expendable aluminum
tubing carrier. The effort resulted in a charge which shot upward at a 45
degree angle with a shot density of five charges per foot. Later on, a string
of sealed frangible aluminum capsules, each with its own shaped charge,
replaced the tubular gun. These were of particular interest in operations
outside the United States because of their convenience and portability.
The lfi inch overall guns were able to pass through most production
tubings; as to the cable, it was a single-conductor, armored type about ^
inch in diameter, wound on a small truck-mounted winch. Gillingham
deserves the credit for having recognized the appeal of such equipment
and taken the initiative in its manufacture. Even though widespread
commercial success was not immediate, Schlumberger had led the way in
a new and revolutionary technique.

294
Phase Two: 1957-1974

A general view

To set 1957 as the beginning of the contemporary history of the


Schlumberger processes is no arbitrary decision. Even though no major
technical developments took place during that year, it was marked by a
fundamental change in the legal setup of the Company: 1957 saw the
consolidation into a holding company, Schlumberger Limited, of Société
de Prospection Electrique (Europe and Africa), Schlumberger Overseas
(the Middle East and Far East), Schlumberger Surenco (Latin America),
and Schlumberger Well Surveying Corporation (North America). This
restructuring of hitherto more or less autonomous companies into an
international holding company required lengthy and laborious negotia¬
tions with French and American fiscal authorities. Once achieved, it put
the group on a rational organizational basis, no longer tied to family
interests. A few years later the Schlumberger stock was listed on the New
York Stock Exchange. Although members of the family still hold a
substantial number of the shares, they no longer make kinship requisite
for top management. Jean Riboud, the president of the holding company
since 1965, is not related to the Schlumbergers.
Technical and commercial development was deeply affected by the
reorganization. Intensive diversification ensued, especially in the field of
electronics.1 On the other hand, Paris, after some slowdown following the
death of Marcel in August 1953, resumed a leading role, as shown by the

1 This activity is mentioned here for the record only, as it does not fall within the scope of
this work.

295
The History of a Technique

establishment of a new center at Clamart in the outskirts of Paris. Al¬


though less extensive than the Houston center, it handles in a modern
plant an increasing share of the joint research and development program
and manufactures a substantial part of the equipment. A number of
problems of coordination and standardization arise from the distribution
of activities among Ridgefield, Houston, and Paris. It often happens that a
field crew receives equipment manufactured partly in Houston and partly
in Paris, and its assembly in the field is expected to proceed without a
hitch. On the other hand, as regards both design and workmanship, there
is no doubt that the friendly competition between Houston and Paris
contributes to the success of many projects.
The research and development budget ($11,000,000 in 1970)2 al¬
lows for the most ambitious studies. Electronics and informatics give
Schlumberger direct access to highly advanced technology and specialized
know-how. Although basic contributions were perhaps less frequent than
in the postwar years, a definite increase in efforts directed toward improv¬
ing the quality of instruments and services followed in the wake of the
1957 restructuring.
Time, accuracy, and environment are of the essence. The cost of
standby time in drilling, especially offshore, is such that time becomes the
overriding factor. This, in turn, creates a demand for equipment that can
be set up quickly, is foolproof, and combines an array of tools whereby
several measurements can be made in a single run. The increasing demand
for accuracy in interpretation calls for redesigning both the electronics
and the downhole mechanics of the equipment. Finally, each component
requires special study when the equipment must function under extreme
conditions, such as pressure and temperature at depths of 30,000 feet,
polar cold, and storms at sea.
Schlumberger benefited greatly from advances in electronics and, in
the first place, from the development of transistors. Especially for sondes,
their value is obvious. With transistors, volume reductions become possi¬
ble that even miniaturized tubes would not allow. This is an essential asset
for a tool housed in a cartridge whose overall diameter never exceeds 4
inches. Transistors are much less fragile than tubes, sensitive as the latter
are to shocks and vibrations, not only because of the glass, but mainly

2 This figure covers only traditional Schlumberger activities. An almost equal amount is
earmarked for other undertakings of the holding company.

296
After 1945

B *•

Figure 81. Printed circuit and modular element of cartridge.

297
The History of a Technique

because their components (cathode, grid, and anode) have a rather high
inertia and easily become deformed. Furthermore, the transistor needs
very little power, a valuable asset for circuits operating at the far end of a
cable several miles long and ill suited for power transmission. In the latter
case it was necessary to await the advent in 1954 of silicon transistors,
which can withstand temperatures of 350 degrees F.
The printed circuit technique (Fig. 81) was, in its turn, put to ample
use. In this case, unlike the audiovisual industry, dependability, together
with easier troubleshooting and, mainly, volume reduction, was more
important than cost cutting. On the other hand, the first integrated
circuits, in which all the elements are assembled on a crystal of a few
millimeters, were developed in I960, and only their early inadequate
resistance to temperature delayed for a time their incorporation into
Schlumberger instruments.

The modern tools

Resistivity Tools

By 1957 the development of resistivity measurements had almost


attained its high point. Schlumberger’s unique experience, plus the basic
improvements that had been added, left little room for further changes
except to increase the dependability of the equipment through better
selection of the components, to refine the features of the instruments, and
to combine them in order to obtain maximum logging capacity in a single
run.
The last pulsated current systems gradually passed out of use in the
1960’s. The 20 year old chronological system was superseded by theoreti¬
cally superior and at the same time equally convenient operational tech¬
niques.
A new 6 FF 403 induction sonde was produced with a greatly in¬
creased depth of investigation. The short normal was replaced by
Laterolog 8, which had the same shallow depth of investigation and,
through appropriate focusing, was less affected by variations in borehole

3 The code name for a six-coil, vertically and laterally focused sonde with 40 inch spacing
between main transmitting and receiving coils.

298
After 1945

diameter. Later, a single tool combined two induction sondes, one the 6
FF 40 and the other much less deep, as well as the Laterolog 8. In
conjunction with the Microlaterolog, this tool proved useful in the qual¬
itative interpretation as well as the quantitative study of resistivities.
Francis Perrin had based his calculations on the assumption that
concentric elementary currents had a negligible mutual effect (see p. 247).
This is not always correct because, at the frequency used, a reaction
resulting from the skin effect can be observed in low-resistivity zones.
Although appropriate charts made it possible to correct the reading, an
automatic correction at the time of recording was more rational, and the
surface apparatus was modified accordingly.
The year 1971 saw the integration into a single tool of an induction
sonde, a sonic device (see p. 307), and a very shallow resistivity instru¬
ment, the “Spherical Focused Log” (S.F.L.). The primary use of the S.F.L.
is to replace the Laterolog 8. Its focusing system forces the current
emitted by the central electrode to spread as in a homogeneous isotropic
medium. The equipotential surfaces are then spheres. The borehole effect
is eliminated without any increase in the depth of investigation, thus
providing an accurate resistivity measurement of the invaded zone. The
induction-sonic-S.F.L. combination was a distinguished technical
achievement; the tool was particularly suited to the characteristics of the
Texas and Louisiana coasts.
In regard to the measurement of Rt, it is worth putting on record that
the advocates of Laterologs 3 and 7, up to recent years, could not agree as
to which was the better system. Although Laterolog 3 is easy to combine
with radioactivity instruments, Laterolog 7 gives a better determination of
the true resistivity, especially in cases where Rt is unusually high. One
might say that, roughly, Laterolog 3 is better suited to the conditions of
American fields and Laterolog 7 to those of the eastern hemisphere. This
is the reason why Houston refined Laterolog 3, and Paris Laterolog 7.
Laterolog 3 became a conductivity measuring instrument. The prob¬
lem was not merely to convert the ohm scale into mhos, which could have
been done by a reciprocator, but to obtain a measurement signal linearly
proportional to conductivity: the current emitted by the central elec¬
trode being constant, voltage varies as resistivity; conversely, if the volt¬
age is kept constant, conductivity is measured by the current variations.
The latter system is used in Laterolog 3.
From 1964 Laterolog 7 included a long-contemplated downhole

299
The History of a Technique

amplifier. Next came the analysis of the so-called “Delaware effect, first
observed in a western Texas basin on a sand overlain by a thick anhydrite
bed of practically infinite resistivity. The effect is eliminated by a special
arrangement of the electrodes, unless mutual induction generates pertur¬
bations in the opposite sense:4 the “anti-Delaware effect.”
During this period detailed theoretical studies were available, di¬
rected at defining a Laterolog suitable for conditions, frequent in the
eastern hemisphere, of a true resistivity much above that of the invaded
zone. André Poupon was the first to become fully aware of the problem
and was able to enlist the interest of Doll, who lent his support and
original views. Jean Dumanoir was put in charge of the mathematical
treatment, and after many transatlantic discussions specifications for the
tool were set. Two systems were combined: the deepest possible
Laterolog, and another one with guard current return electrodes close to
the transmitting electrodes, an arrangement that focuses the current
throughout the significant part of its flow through the invaded zone. This
shallow investigation system is called a “Pseudolaterolog” (Fig. 82). How¬
ever attractive the combination, it appeared that two resistivity mea¬
surements are not always enough to give a good definition of the invasion
profile and to determine Rt accurately; an additional Microlaterolog-type,
low-penetration measurement was needed on the borehole wall. As the
design of a tool able to make these three measurements simultaneously
seemed difficult, a phased approach was adopted. With the measurement
of Rxo left to the conventional Microlaterolog, an instrument was created
that, in two successive runs and without leaving the borehole, recorded
both the Laterolog and the pseudolaterolog curves. As a wide range (0.2
to 40,000 ohms) was desired, a technique had to be devised that would
associate the features of both measurement systems: conductivity and
resistivity. Under the name “constant power technique,” it consists of
maintaining the constancy, not of the voltage (V0) or the current (70) of the
central electrode, but of their product, that is, the power.
The results, bolstered by those of the Microlaterolog, proved quite
satisfactory. The presentation of the three logs in logarithmic scales
allowed in particular for the determination of their ratios, which are

4 Schlumherger Documentation, Log Interpretation, Vol. I: Principles, Houston, 1972; J. Suau,


L Outil Dual Laterolog Rx0, Revue de l'Association Française des Techniciens du Pétrole, No.
223, January-February 1974.

300
After 1943

44444^1

Figure 82. Principle of the Double Laterolog. Left: Deep Laterolog. Right: Pseudolaterolog.

essential data for the “quick look” (rapid qualitative interpretation), as


well as for a complete quantitative analysis.
The final version, the “Dual Laterolog-/?x0,” was completed in 1972.
Three resistivities are recorded simultaneously: deep, medium, and very
shallow. The last of these is no longer dealt with by the Microlaterolog,
which is excessively affected by thick mud cakes, or even by the Proximity

301
The History of a Technique

log,5 which operates satisfactorily only if the invasion is rather deep;


rather, it is measured by a miniaturized, pad-mounted “Spherical Focused
Log” (Micro S.F.L.), which is hardly affected by the mud cake in spite of
its very shallow depth of investigation. Moreover, by combining the
values of the focusing and measuring currents, the Micro S.F.L. provides
an indication of the thickness of the mud cake and a quality factor for the
Rxo measurement.
The fact that in a Dual Laterolog the two systems use in common the
electrodes A0, Mlf M [, M2, and M'2 entails equal thicknesses of the current
sheets, and therefore of the vertical definition of both measurements;
even in thin layers the responses are in complete concordance, a great
advantage in getting clearer results and exploiting them quantitatively.
However, as had been foreseen, since the Micro S.F.L. measurement gives
a much finer vertical resolution than the other two, it results in an often
unwelcome degree of detail. This led to “smoothing” the Micro S.F.L.
curve by a weighted average, automatically computed over an interval
with a thickness equal to that of the Laterolog-Pseudolaterolog current
sheets.6 The result is equal vertical resolution for the three curves, which,
in addition, are recorded at the same depth by means of a memorizing
system. Finally, the tool also measures the S.P. through an electrode
located far above the metallic mass by which it could be affected. This S.P.
is also memorized and recorded at the same depth as the three resistivity
logs. When the S.P. reading is unsatisfactory, as often happens under
conditions in which the Laterolog is the most effective tool, a gamma ray
log7 is substituted, the appropriate cartridge then being housed inside
electrode A2.
The simultaneous recording of multiple parameters, the logarithmic
scale presentation, and the smoothing and memorizing are all operations
made possible by modern electronics. Perhaps less spectacular, though
equally essential, is the mechanical sophistication of the equipment. Like

5 Through a focused pad-mounted system, the Proximity log measures the resistivity of the
zone fully invaded by the mud filtrate, but is less affected by the mud cake than the
Microlaterolog, from which it is derived. It was put into operation in 1958 and has
gradually replaced the Microlaterolog wherever there are thick mud deposits.
6 What is actually computed is an average of conductivities.
7 Another solution is the recording of a neutron log (C.N.L.; see p. 304) or a compensated
density log (see p. 304-305), but the quality of the pseudolaterolog is affected by the
excentralization these tools require.

302
After 1945

all tools using pads, the Micro S.F.L. requires tight contact with the
borehole wall. Correct alignment is achieved through a system of four
hydraulically collapsible legs, which also ensures the centering of the
bottom of the sonde and the measurement of the borehole diameter.
The Dual Laterolog-i^j.0 has confirmed in practice the theoretical
considerations on which it was based, thus justifying the effort invested in
its design.

Nuclear Tools and Techniques

Only after 1957 did nuclear logging techniques begin to show prom¬
ise of successful development. Every service company concentrated its
efforts in this new field, either on instrumentation or on interpretation.
Although not a pioneer, Schumberger built on its experience and re¬
mained in the leading ranks.
One of the first objectives, as was true for the whole nuclear industry,
was to design better detectors. As radioactivity phenomena are of a
random nature, the number of events occurring within 1 second—
emission of natural or induced gamma rays, neutrons, or other particles—
is variable. A valid measurement requires an average over a time span
that, according to the theory of probability, will be longer as the events
are more widely spaced. All the counters used heretofore, whether ioniza¬
tion chambers or Geiger-Muller counters, were of poor efficiency as they
counted only a minor part of the rays passing through them. To obtain
significant measurements, logging required either overlong counters,
which worked against the degree of resolution, or slow logging speeds
that extended the duration of the operation. The scintillation counter (a
photoluminescent crystal coupled with an electron multiplier) was wel¬
comed as the ideal detector for logging, but the question of downhole
temperature had to be dealt with first. Without waiting for the manufac¬
turers’ improvements, Schlumberger tackled the problem by undertaking
an exhaustive study of the components that could serve in such an
instrument. In I960, with the assistance and advice of André Lallemand,8
Jean-Pierre Causse designed and built a scintillation counter combining

8 A French astronomer, the inventor of cells with high-sensitivity electron multipliers,


which he developed at the Paris Observatory for the electronic camera he had conceived in
1937.

303
The History of a Technique

good performance under high temperatures with excellent mechanical


strength, which proved highly suitable for logging. Its standards are so
high that it has found wide application outside logging, especially in space
technology; it is part of the equipment on a number of American satel¬
lites.
The neutron logging technique evolved along several lines. A basic
change was the use of a criterion other than the number of gamma rays of
capture to measure the neutron population in the vicinity of the detector,
with the result that the present systems measure the epithermal neutrons,
that is, those with an energy level slightly above 0.4 electron volt (eV), or
the thermal neutrons, 0.025 eV. A further step was in the direction of
lessening the effect of the mud. To this end the whole tool was initially
excentered so as to rub against the borehole wall, an arrangement soon
improved by introducing a neutron source and a detector into a kind of
elongated skid hydraulically pressed against the wall. In an even more
recent system the “Compensated Neutron Log” (C.N.L.) eliminates the
mud effect by making simultaneous measurements on two counters lo¬
cated at different distances from the source; the relation between the
readings, automatically established by the recorder, determines a porosity
index that is practically independent of the borehole. The C.N.L. utilizes
very powerful sources—up to 16 curies—such as plutonium-beryllium or
americium-beryllium, which allow for increased spacing between source
and detector. Thus, while maintaining a counting rate adequate for a
satisfactory iogging speed, the investigation is deeper.
Together with improved techniques, new methods were elaborated.
As measurement of density by gamma ray absorption was used in various
industries, it seemed logical to apply it to logging. The idea was that the
porosity of a rock could easily be deduced from its global density, pro¬
vided that the densities of the matrix and of the pore-filling fluid were
known.9 The matrix density is well known for various formations: for

9 If Pb is the measured global density, </> the porosity, pma the density of the matrix, andpr
the density of the pore-filling fluid:

Pb = Pma( 1 - (f>) + prf>

Hence
Pma Pb
cj) =
Pma Pf

304
After 1945

typical rocks it varies between 2.65 and 2.87. As to the fluid—since


theory contemplated a shallow depth of investigation involving the in¬
vaded zone it usually is a mud filtrate with a density in the vicinity of 1.
A small amount of residual oil, with a density between 0.8 and 1, hardly
affects the results; only the possible presence of gas will lead to an
exaggerated porosity estimate. From thereon the road seemed clear, and
Lane Wells was the first to tread it. Its tool was a sonde containing a
radioactive source and a detector, pressed against the borehole wall by a
spring blade. Shielded against the direct flux of the source, the detector
counted the gamma rays after their diffusion into the surrounding forma¬
tions. The theoretical basis was found to be correct, but mechanical
trouble was serious inasmuch as the mud cake effect was important, and
loss of contact with the wall led to aberrant results.
Schlumberger tackled this type of logging in 1957. On the instigation
of Majani, then manager of the Maracaibo Division, the first tests were
conducted by Tittman, a nuclear physicist of the Ridgefield center. One
year later, the “density” or “gamma gamma” log appeared. In the begin¬
ning it had to compete with the sonic log on the Texas and Louisiana
coasts, but its intrinsic qualities finally won out. Much of its success was
due to Schlumberger’s special lead in hydraulically controlled power
sondes. Another 6 years passed, however, before a final solution to the
mud cake problem was found.
As for the previously mentioned C.N.L., the locating of a second
detector close to the gamma source provides a double reading from
which, in the course of recording, the correct density values can be
automatically computed. For reservoirs filled with liquid, the subsequent
porosity determinations are entirely satisfactory. Although this is not true
for residual gas, the difference between the porosity given by the neutron
log and that provided by the density log, which is overestimated, is in
itself a valuable indicator of the presence of gas, a criterion widely used in
interpretation.
Another innovation in the nuclear field required a much more am¬
bitious program. For a long time there had been interest in the rate of
decay of thermal neutrons as a function of time: this rate increases with
the amount of chlorine in their path. The reason for this is that, of all
common elements, chlorine has, by far, the greatest neutron absorption
power. Thus the rate of decay of thermal neutrons depends on the

305
The History of a Technique

amount of sodium chloride or, ultimately, salt water.10 The striking anal¬
ogy with other measurements also dependent on the amount of salt water
in a nonconductive medium gave reason to hope that a log could be
obtained that would be similar to the resistivity log but could be recorded
through the casing. It was not for Western scientists to lead the research
in this field: in 1958 the Russians published their first experiments.11
Conventional sources cannot be used to measure the rate of neutron
decay as a function of time; a discontinuous source that generates a large
number of neutrons within a few microseconds is needed. The decay is
measured within a few milliseconds following the interruption of the
emission. The process is repeated a great number of times within a
second. The measurement is simplified because of the fact that, the decay
being exponential, two values suffice in principle to give the time con¬
stant. A log is obtained with a pulsated source and a detector for gamma
rays of capture or thermal neutrons. From 1964 Lane Wells used this
system in the United States under the name “Neutron Lifetime Log.”
Schlumberger’s version, put into operation in 1966, was called the
“Thermal Decay Time Log” (T.D.T.).
The great difficulty here lies in “manufacturing” neutrons in large
quantities within a very short time span. The most suitable method is to
use a particle miniaccelerator to bombard a tritium target with deuterium
atoms. Yet to generate neutrons the nuclear reaction calls for energy
involving a potential difference of some 100,000 volts, an apparently
insuperable obstacle considering the reduced space and high tempera¬
tures at the bottom of a borehole. The problem was solved, not only for
standard sondes about 4 inches in diameter, but also for the lfè inch
sondes required to go through the tubing. This was a remarkable feat,
particularly for the latter, which are used to run through the tubing under
pressure and log the formations behind the casing while the well is
flowing. One interesting application is the repeated observation at
scheduled intervals of the saltwater level in a producing horizon.
Schlumberger’s contribution to the pulsated neutron technique was

1(1 The conventional neutron log studies the distance covered by the particles before their
collisions with hydrogen atoms render them thermal or epithermal. In the present case
what is at stake is the life duration of the thermal neutrons before absorption.
11 B. G. Erozolimski, R. L. Voitsik, N. V. Popov, and A. S. Chkolnikov, “New Methods of
Logging Bore Holes Using Pulsed Neutron Sources ''Neftyanoie Khozyaistvo, Vol. 36, 11
(1958).

306
After 1945

important not only for the instrumental progress achieved, but also for
the practical utilization of the results. The Ridgefield and Houston centers
manufactured small series of instruments dependable enough to be put
into service in the Middle East, but their correct operation was made
possible only by the international technical organization. An original
method had been devised for precision measurement of the neutron rate
of decay;12 interpretation was supported by Schlumberger’s long experi¬
ence.13
The function of the T.D.T. is not to evaluate a reservoir before casing
is set. In the current state ol the technique, its depth of investigation is not
sufficient to make it competitive with the various resistivity logs. The
quantitative interpretation of the results, however, makes it an outstand¬
ing instrument for production control insofar as it determines drainage
efficiency and assists in optimizing production methods. Such is its role
today in the most productive fields of the U.S.S.R., the United States, and
the Middle East.

The Sonic Log

The origin of the sonic log is somewhat curious. It was developed as


an adjunct to seismics; initially it contributed nothing to log interpreta¬
tion, although it was to become one of its essential elements. However, its
history begins much earlier.
In this connection a return to surface prospecting is appropriate.
Schlumberger’s field activities emphasized electrical methods, but since
1935 seismic prospecting had enjoyed the lion’s share in the study of deep
structures. It was known that seismic (or sound) waves propagate at
velocities varying with geological formations. The thought then came
naturally of an acoustical log, similar to the resistivity log, and on June 1,
1934, a patent was filed by Schlumberger under the title Procédé et Ap¬
pareillage pour la Reconnaissance des Terrains Traverses par un Sondage
(“Process and Apparatus for the Reconnaissance of Formations Drilled
Through by a Borehole”) (Fig. 83).

12 J. S. Wahl, W. B. Nelligan, A. H. Frenttrop, C. W. Johnstone, and R. J. Schwartz, "The


Thermal Neutron Decay Time Log,” Society of Petroleum Engineering Journal, December
1970.
13 C. Clavier, W. Hoyle, and D. Meunier, “Quantitative Interpretation of Thermal Neu¬
tron Decay Time Log J Journal of Petroleum Technology. June 1971.

307
The History of a Technique

RÉPUBLIQUE FRANÇAISE.

MINISTÈRE DU COMMERCE ET DE L’INDUSTRIE.

DIRECTION DE LA PROPRIÉTÉ INDUSTRIELLE.


--£*$*<}“-

BREVET D’INVENTION.
Gr. 8. — Cl. 1. N° 786.863

Procédé et appareillage pour la reconnaissance des terrains traversés


par un sondage.

Société dite : SOCIÉTÉ DE PROSPECTION ÉLECTRIQUE (Procédés SCHLUMBERGER)


résidant en France (Seine).

Demandé le 1er juin 1934, à 16 heures, à Paris.


Délivré le 17 juin 1935. — Publié le 11 septembre 1935.
[Brevet d’invention dont la délivrance a élé ajournée en exécution de l’art. 11 S 7 de la loi du 5 juillet 1844
modifiée parla loi du 7 avril 190a.]

N 786.863 Société dite : PL unique

Société de Prospection Électrique

(Procédés Schlumberger)

308
After 1945

On the basis of acoustic velocities of 1000 meters per second ob¬


served in sands, and 5000 in hard limestones, the patent covered “the
reconnaissance process of the formations drilled through by a borehole
. . . which . . . consists in determining the velocity of sound propaga¬
tion in the various formations penetrated, such velocity being measured at
various depths in the open-hole section of the borehole” (translation from
the French). Also covered by the patent was the possibility of measuring
in the same operation the intensity of the sound received, which would
give “an indication on the absorption caused by the various formations
drilled through” (translation). In fact the patent covered all the sonic
techniques used today. The sonde included a sound generator and, close
to the source, two microphones appropriately spaced vertically. If the
acoustic signal is a sinusoid vibration, the propagation velocity in the
formations is determined by the phase difference between the mi¬
crophone signals; if it is not, a highly picturesque method comes to the
rescue, based on the biauricular technique. It consists in listening, both
ears working separately, to two earphones, each one fed by its own
downhole microphone. The operator, all ears, moves a kind of trombone
slide, thereby lengthening or shortening one of the channels, until he
perceives the sound coming from straight ahead of his honest face (see
Fig. 82, lower right). The position of the slide determines the phase
difference, and the rest of the operation is pursued as in the case of
sinusoid vibration.
With an automobile horn for sound, the device was tested in Pechel-
bronn in 1935. The biaural system proved too inaccurate, and the only
remaining record of that episode is of a practical joke;14 thereafter,
despite a July 1947 research note from Doll, the matter rested until 1952.
However, the surface geophysicists felt an increasing need to know
accurately the velocity of seismic waves in various formations, and it often
happened that, as soon as a borehole had been drilled in an unexplored
area, this parameter was measured point by point.15 A geophone was
lowered to the desired depth at the end of a logging cable, a dynamite
charge was detonated at the surface, and the time was measured between

14 At a certain moment the operator heard, instead of the usual cacophony, these distinctly
spoken words: “Trilobites speaking. Go to hell!” (Trilobites are fossil crustaceans of the
Paleozoic.)
15 There is here a similarity with one of the first goals of downhole resistivity measure¬
ments: to obtain parameters for the interpretation of surface results.

309
The History of a Technique

the explosion and the arrival of the seismic wave at the geophone. The
operation was repeated at some 20 different levels in the hole; the process
was slow and costly, however, and there was too great a discontinuity
between results for the desired accuracy.
At the instigation of Frank Kokesh, a field engineer transferred to
Houston, better systems were designed. In the first the arrangement was
reversed: with the geophone at the surface, a standard perforator fired 24
blank cartridges at as many depths in the borehole. Much time and many
explosives were saved, but the energy of the shots did not reach beyond a
few hundred meters. The few units built were used in 1952 in shallow
stratigraphic reconnaissance coreholes. In 1955 new tests were made with
the geophone being lowered downhole about 100 feet below the per¬
forator, making it possible to measure, regardless of depth, the propaga¬
tion time over 24 intervals in a single run. This “Long-Interval Velocity
Logging” system was marketed for several years without any great success.
One deterrent was the fact that two instruments 100 feet apart on the
same cable meant a difficult fishing job, should either get stuck or lost in
the hole.
The advantage of all these measurement methods was their relative
facility. Propagation times were short, though not much more so than
those of the conventional seismic method. But the processes were clumsy,
and the results lacked continuity. The geophysical services of the major
companies, Magnolia (Mobil), Humble (Esso), and Shell, resumed the
search for a continuous propagation velocity log and undertook the de¬
velopment of more advanced techniques. Although reverting to spacings
of a few feet between transmitter and receiver allowed for a much weaker
source, it entailed measurements in microseconds, a unit not current in
geophysics. Electronics easily solved this problem in 1952. Two systems
prevailed. In the first, propagation time between transmitter and receiver
was measured; in the other, as in the 1935 Schlumberger patent, the
measurement took place between two vertically spaced receivers, which
in principle eliminated the travel time in the mud and hence the effect of
the borehole diameter.
The first operational equipment (March 1954) was produced by
Magnolia, which licensed its patent rights to the Seismographic Service
Corporation. In 1955 Schlumberger acquired the patent rights of Humble
and thereupon developed a two-receiver system, which was marketed in
1958.

310
After 1945

In addition to the essential contribution of the sonic log to seismic


interpretation, more applications were soon found. Tom Hingle, a pet¬
rophysicist of Magnolia in the Mid-Continent, had observed correlations
between formation factors and the first sonic log readings. Experience was
soon to confirm that the more porous a formation, the lower is its sound
velocity. M. R. J. Wyllie, already known from his study on S.P., deserves
the credit for having, in collaboration with A. R. Gregory and L. W.
Gardner,16 expressed this correlation in mathematical terms. Their “time
average formula” indicates that the time for a sound wave to cover a
certain distance in a porous medium equals the sum of the time spent in
the liquid and that spent in the mineral grains.17
These efforts to measure porosity lent considerable impetus to the
development of the sonic log. It gave excellent results in sands and, as
shown by Tixier, was remarkably compatible with the induction log in
shaly sands.18 In conjunction with the neutron and density logs, and
integrated into the system of parameters fed into the computer for the
complete processing of data, the sonic log has become a basic factor of
reservoir evaluation.
In October 1965 two experts of Shell Oil Company, C. E. Hottman
and R. K. Johnson, presented to the Society of Petroleum Engineers a
paper entitled “The estimation of Formation Pressures from Log-Derived
Shale Properties,” in which they pointed out an interesting possibility re¬
garding sonic log interpretation. While the travel time in shales evolves
steadily with depth, distinct anomalies are observed when the bottom of
the borehole is approaching a high-pressure porous bed. This is a most
useful warning for the driller since it signals the danger of a blowout.
As anticipated in the 1935 Schlumberger patent, by then fallen into
both oblivion and the public domain, velocity is not the only significant
factor: another source of information is the intensity of the sound. It
could not be measured with the initial sonic instruments, but their ab-

16 M. R. J. Wyllie, A. R. Gregory, and L. W. Gardner, ‘‘Elastic Wave Velocities in


Heterogeneous and Porous Media,” Geophysics, Vol. 21, pp. 41-70 (1956).
17 Their formula is 1 _ </> 1 — <p
TV =_vT + —G-
where 0 represents porosity; V m, measured velocity, Vf, velocity in the fluid; and V r,
velocity in the rock grains.
18 Although both logs are affected by shale, a first approximation of the oil content is gained
by proceeding as for nonshaly sands, without either log requiring correction.

311
The History of a Technique

normal behavior signaled zones of very low intensity. It soon became


manifest that such intensity drops occurred precisely where high acousti¬
cal absorption, caused chiefly by horizontal fractures in the rock, could be
expected. Despite the temptation to look here for a fracture indicator,
experience showed that the phenomenon was complex and needed much
closer scrutiny before an unequivocal diagnosis could be formulated.
Far from laboratories and engineering offices, there was an unex¬
pected development. The sonic log had just been introduced in Ven¬
ezuela, where the Maracaibo crews were pursuing a suggested experi¬
ment. Until then all measurements had been in the open hole; would a
sonic log recorded inside the casing have any significance? Pierre Majani
and Alan Rushton were struck by certain recurring anomalies. Intuition,
based on a thorough knowledge of drilling, casing, and cementing pro¬
cesses, led them to the conclusion that the cause of these anomalies was
the condition of the cement rather than the cased formations: a good
bonding between steel and rock dissipates the signal, of which only a
minor part reaches the detector. Conversely, in the case of poor bonding,
the wave travels mainly in the casing, and the detector registers a strong
signal. Should this assumption be confirmed, it would have momentous
consequences: whereas cementing control then rested on a temperature
log that had to be recorded within days after the setting of the casing, it
would henceforth be possible to do the recording months or years after
the well had been completed.
Day-to-day experience seemed to provide confirmation of this view,
but for lack of any other substantiation it was accepted only with reluc¬
tance. Proof required that the equipment be modified so as to record the
amplitude (intensity) of the signal. Again the Maracaibo team did what
was needed: its electronics specialist, Jean Bouffard, traveled to Houston
to work out the change he had proposed. A new service was born:
“Cement Bond Logging” (C.B.L.), which after 1961 superseded ther¬
mometric cementing control.
Even though earlier tests had shown the difficulty of detecting frac¬
tures by sonic logging, the importance of the objective justified further
effort. The analytical approach was to differentiate in the acoustical wave
train between the compression and the shear wave, from which differ¬
entiation might come the ability to locate horizontal as well as vertical
fractures. Geographically limited successes resulted. Although a number

312
After 1945

of specialists hold that the key to the problem is at hand, it seems doubtful
that a solution is imminent.
An altogether different application of the acoustical phenomena was
proposed in 1956 by Dowell, specialists in oil well cementing, acidizing,
and fracturing. It was a sonar-type device, designed to measure the
volume of the huge caves obtained by dissolving rock salt in order to
provide space for hydrocarbon storage. On the basis of the same idea,
Mobil Oil, in 1967, took out a patent covering an instrument having the
rather misleading name “Borehole Televiewer.” The target was no longer
caves but conventional boreholes, and what was measured was not the
propagation time but the intensity of the reflected pulses. It was like
observing the borehole wall in monochromatic light, through the mud and
even the mud cake if any: if it were hard and smooth, it would appear as
shiny, whereas fractures would show in dark.19
Schlumberger, having acquired the license for the Mobil patent,
made it operational, notably in western Texas, Libya, and Iran. A continu¬
ous log is obtained when pulling the sonde up the hole, while a rotating
acoustic beam scans a low-pitch helix all along the borehole wall. In
compact limestones, nonhorizontal fractures emerge sharply, as do the
perforations in a cased well. This is the only method known to locate them
unequivocally.
However important these side applications, the main use of the sonic
log remains the measurement of sound velocity and, through the Wyllie
formula, the computation of porosity. The tools were greatly improved,
and in 1963 the Schlumberger sonic equipment was thoroughly rede¬
signed. The present sonde includes two ultrasound generators and four
receivers. An average is taken between two measurements, one on an
ascending, the other on a descending, wave. This eliminates errors arising
from diameter variations of the borehole or from the inclination of the
sonde in relation to its axis.

19 The question has often been raised as to how useful it would be to televise, or take color
films, in boreholes. However interesting, a picture of a borehole wall would be less
informative than its resistivity, density, sound velocity, and so on. Moreover, a borehole is
usually filled with mud or crude oil, and since the real subjects of interest (i.e., the porous
and permeable borehole walls) are plastered with mud cake, nothing much would be
revealed. The television cameras that have been designed for highly specific purposes have
found little application.

313
The History of a Technique

Miscellaneous Techniques and Tools

The continuous dipmeter with its three Microlaterolog pads had


quickly achieved general use since its creation in 1955 and had opened
the way to a new interpretation technique. Thorough studies in southern
Louisiana showed that valuable data could be obtained not merely from
the dip itself but, above all, from its variations relative to depth. Further
developments included the identification of the structural dip, the detec¬
tion of fault crossings, the location of old channels in deltaic complexes,
and even the determination of the direction of the current at the time of
sedimentation. Such interpretations were due to the collaboration of oil
company geologists and Schlumberger engineers, of whom A1 Gilreath
was the first to formulate the basic principles.20 Ray Campbell, then at the
Ridgefield Research Center, participated in working out the theory, ac¬
cording to which a more thorough interpretation of the dips required a
much higher number of determinations per meter than heretofore.
Rather than perfecting the existing dipmeter, the need was for a
system with the highest possible resolution. Michel Gouilloud was put in
charge of the project and achieved an outstanding new correlation dipme¬
ter. Its telemetry includes the poteclinometer; a fourth pad confirms the
results of the other three while ensuring a better positioning of the tool in
the borehole. The microresistivity curves are as detailed as possible, and
since five of them are recorded simultaneously (the fifth being used for
the evaluation and correction of the speed variations of the sonde as it is
being pulled out), the number of data transmitted by the cable created a
problem that only sophisticated electronics could solve. The sonde being
powered, the opening, the closing, and even the pressure of the pads on
the borehole wall are controlled from the surface. After the instrument
was put into operation in 1966 under the name “High Resolution Dipme¬
ter” (H.R.D.), its mechanical and electrical dependability was enhanced
by several changes. An important step forward was the manufacture by
Schlumberger in 1970 of the first sonde made entirely of titanium, a metal
preferable to the conventional stainless steel because of its better
mechanical behavior, its light weight, and its corrosion-resistant and non¬
magnetic qualities (Fig. 84).

20 J. A. Gilreath and J. J. Maricelli, “Detailed Stratigraphic Control Through Dip Compu¬


tations," Bulletin of the American Association of Petroleum Geologists, Vol. 48, No 12 pp
1902-1910 (December 1964).

314
After 1945

So many correlations can be derived from the extremely detailed


microresistivity curves that they are beyond the scope of conventional
analysis. From the outset the high-resolution dipmeter was conceived for
computer processing, which required recording on magnetic tape and
adequate programming. Magnetic tape recording was not limited to the
dipmeter; in I960 it was already clear that data processing was to play a
leading role in the interpretation of logs of all kinds. Programming was
and continues to be the subject of intensive research from log correlations

315
The History of a Technique

proper to the analysis and presentation of results. A few magnetic re¬


corders were operating as early as 1961, but they came into general use
only with the advent of the H.R.D., and from that time forward the
optical log of the dipmeter merely monitored the operation. The dipme-
ter has become today an indispensable tool for the geologist in the
exhaustive study of sedimentation and structural problems.
An interesting outgrowth of the dipmeter is the use, since 1973, of
the four-arm sonde as a borehole caliper under the name “Borehole
Geometry Tool.” As the cross section of a borehole is more frequently
elliptic than circular, the four arms allow for the measurement of the two
axes of the ellipse, and consequently a more accurate computation of the
volume. The (optional) use of the poteclinometer gives the azimuths of
these axes and reveals the distortion and rotation of the ellipse relative to
depth; heretofore the geometric representation of the borehole had never
reached such a degree of accuracy.
Whereas the C.S.T.-type sidewall sampler has undergone few
changes in recent years, the formation tester has been substantially im¬
proved. Fitted with a second pad and shaped charge, its sampling of small
intervals (1 foot) is more representative than a single-hole test. In addi¬
tion, the two-pad tool can be used for taking a fluid sample through the
casing, as is desirable in some cases. A drawback is that the perforations
left in the casing provide openings through which—at inopportune times
—water, oil, or gas can enter the borehole. The conventional remedy is to
plug them by a squeeze job involving a special truck equipped with
powerful pumps, although often only a few quarts of cement is required.
Schlumberger therefore designed a system whereby the tester, once
anchored in place, discharges into the perforations the amount of quick¬
setting cement necessary to seal off the holes. With this design the tool
becomes more complicated, longer, and heavier. The operation is delicate
and the choice of cement critical; ultimately, success is contingent mainly
on how good the initial cementation between casing and formation was. In
any case, the operation is worth trying; if it fails, it is not too late to
summon the cementing truck.21

21 This same Schlumberger tool is used for an altogether different purpose: the injection
into incompetent sand of resinous compounds that bind the grains without excessively
reducing the permeability of the reservoir, and prevent them from being carried by oil or
gas through the production string and eroding it.

316
After 1945

The coring of hard formations reverted in 1970 to diamond tools, but


instead of the Mennecier system (see p. 284) two small circular saws,
slightly offset at an angle from one another, travel vertically along the
borehole wall and carve a 4 or 5 foot triangular prism with a concave base.
The mechanism is delicate and complicated; much know-how is needed
for its anchoring in the well, the automatic feed of the saws, the recovery
of the sample, and so on. The elegant mechanical concept, however,
would have made Marcel happy.
From 1957 to 1970 perforation made substantial progress. These
advances derived from the technology of explosives and shaped charges
and also from better adaptation of perforators to completion methods.
Today bullets have been completely superseded by shaped charges.
With powerful, high-temperature-resistant explosives and conventional
carriers, penetrations of 20 inches through casing, cement, and formation
can be obtained that, in almost any case, exceed the depth of the invaded
zone. Different liners are used in the shaped charges in order to ensure
penetration, fracture the cement, or even crack the rock, according to
what is needed. Perforation below the tubing, as worked out for the
permanent well completion technique, is now frequently used, not so
much to avoid killing a well under production as to optimize its output
from the beginning. Indeed, a conventional perforation entails the risk of
a blowout unless the well is filled with a mud of such high density that its
hydrostatic pressure overbalances that of the reservoir. The trouble,
however, is that this heavy mud may invade the porous formation and
reduce production. On the contrary, when an oil well is fitted with
wellhead tubing and packers, it can be perforated safely even if the
reservoir pressure is very high. This significant advantage explains the
existence of the many “through tubing” perforator models, using small-
diameter shaped charges rendered increasingly powerful by constant
engineering efforts. They represent an important contribution to the
improvement of well completion methods.
Multiple completion, that is, the simultaneous production of several
beds in a single well with one tubing for each reservoir, has given rise to
specialized techniques. Such wells may include the exploitation of up
to five pay zones, each with its individual tubing. The problem is then to
perforate the casing at the desired level without damaging the tubing or
tubings that carry the production from the lower beds. To determine the
azimuth of the firing, a radioactive pellet is lowered inside each tubing to

317
The History of a Technique

the desired level; the pellets are located from inside the casing by a
directional detector coupled with a swiveling perforator firing in a single
direction. Another system comprises a gamma ray densimeter, which is
also directional and locates the tubings by the increased density read in
their direction. These are all exceptional operations, difficult to execute
and requiring an extremely complex apparatus, along with a flawless
interpretation of measurements.
Traditionally, Schlumberger’s activity in the field began with the
drilling of an oil well and ended with its completion. As soon as the well
was connected with the pipeline, the crews had to withdraw without the
opportunity to put their experience and versatile equipment to further
use. There seemed to be a loss of potential business in this routine. One
possible direction for such expanded activity was analysis of the flow at
the various producing levels of an oil well. Optimum production of one or
several reservoirs mandates that, when necessary, the behavior of a well
be corrected. This, in turn, requires information on the amount and
nature of the fluids flowing into the well at every level. The greater the
production, the more important the problem, as is typically the case in the
Middle East. However, it was in 1957, in Brunei, that at Shell’s instiga¬
tion, and thanks to the perserverance of Simon Noi'k, Schlumberger
became interested in the question. On his own initiative, Noik had built,
with locally available facilities, rudimentary tools promising enough to
convince Paris that the lead was worth following. It soon became apparent
that, in addition to instruments, the primary need was for a theoretical and
experimental study of the flow of fluids in vertical tubes. A laboratory
equipped for the simulation of the whole range of diphasic flows as they
occur in petroleum production was built—probably the only one of its
kind in Europe.
A flowmeter suffices in the case of monophasic production: the total
flow is measured point by point, the inflow at each level being determined
by the derivative of the flow curve as a function of depth. The situation is
no longer simple when the flow becomes diphasic or triphasic (water-
oil-gas) because the flowmeters are affected by the nature of the fluids
driving them. A system of measurements then becomes necessary in
which the percentage of each phase is determined level by level and from
which, in turn, the amount of each fluid entering the well can be calcu¬
lated at each level. One difficulty is that in order to lower the tools below
the tubing small diameters (about lyg inches) are required; another

318
After 1945

is that the measurements must not interfere with the phenomena being
evaluated.
Schlumberger flowmeters are of the spinner type. The percentage of
each phase is obtained by also measuring the density and the dielectric
constant of the mixture: the density reveals the ratio between gases and
liquids, and the dielectric constant that between water and hydrocarbons.
Once more, the research program proved more complex than anticipated,
especially as the solution to the problem calls for two series of instru¬
ments, one for outputs exceeding 500 barrels per day, another for lesser
productions. In the latter case, fluid velocity is boosted by narrowing the
flow section with an inflatable bag. The measurements are translated into
simple physical parameters: electrical capacitance, revolutions per second,
frequencies, and so on. However, there is another original method, also
credited to Noik: determination of the density on the basis of the hy¬
drostatic pressure differential between two levels (Gradiomanometer).
Other tools complete the array, in particular a special caliper, and a
high-resolution thermometer which locates gas leaks in the tubing and
fluid flows behind the casing.
In spite of the importance of production logging a long time was
required for it to prevail. The interpretation technique, which had to be
created from scratch, showed the possibilities and limitations of the
system. Although its advantages are no longer questioned in the case of
large producers, the fact is that most oil wells are small producers for
which such a service, however beneficial, remains too complicated and
costly. Yet it seems logical to think that Schlumberger’s initiative, im¬
mediately followed by the activity of its competitors, has marked the
beginning of an irreversible trend and that with the benefit of new
measurement techniques, among which the Thermal Decay time log is
already operational, this type of logging will become a key element of
hydrocarbon production.

A Note on Offshore Drilling

Even to a public not especially aware of petroleum problems, there is


something exciting in offshore drilling with its great economic invest¬
ment, its quality of adventure, and its use of bold techniques. Con¬
sequently, the question may be asked why a special chapter has not been
devoted to it. Indeed, for Schlumberger the problem was not new: by

319
The History of a Technique

1930 drilling in southern Louisiana had moved from the “bayous” to the
coast, and from there to the shallow waters of the Gulf of Mexico. In
Venezuela, too, in the 1950’s drilling platforms could be found in the
middle of Lake Maracaibo. Solution of the problem had simply required a
switch from the truck of the O.S.U.-C (see p. 224). This unit was normally
installed permanently on the drilling platform and included the equip¬
ment common to all services: cable, winch, cabin, optical recorder, and
generator. As a radio network and air or sea transport were available, no
time was lost in moving the crews and their equipment to the site.
Far greater difficulties arose with deep-sea drilling, and to master
them a new generation of platforms was needed. On the other hand,
logging was not affected by any major new problem. Technical features
were gradually adjusted. There was, first, the question of the optimum
location for the O.S.U.-C. The prohibitive cost of standby time on
offshore platforms required as many measurements as possible in a
minimum of logging runs; combination tools had to be designed to meet
this objective. Along the same lines the first radio transmissions of logs
were studied to devise the fastest possible delivery of logging results
Finally, specialized crews had to be set up to meet the specific require¬
ments of the platforms and their environment—a minor task for a com¬
pany whose motto was “Wherever the Drill Goes, Schlumberger Goes.”
Though there are colorful elements in the ways to reach the platforms and
in the work and life on them, logging services—whether in the North Sea
or off the shores of Alaska, Nigeria, or the Sunda Islands—continue to
offer the same routine efficiency as they do in Iran or Oklahoma (Fig. 85).

Presentation and interpretation of logs

Despite a wealth of new and improved services, the outstanding


achievement of the 1960’s was the progress in interpretation.
Any log reading involves two stages: an immediate and a deferred
interpretation. The former determines the decisions to be taken forth¬
with. This is the “quick look”; it usually gives only approximate results,
but takes few computations and can be applied at the well site. The second
interpretation seeks a thorough reservoir engineering study and requires
long and difficult calculations. As a matter of fact, it has been a problem
for all log users to distinguish what must be solved on the spot from what

320
After 1945

Figure 83. Schlumberger unit on offshore platform.

321
The History of a Technique

deserves elaborate processing. Only when such priorities have been estab¬
lished, can substantial progress be achieved.
The quick look had been the basis of Schlumberger’s reputation and
success on the Texas-Louisiana Gulf Coast. In this region of sands with
low shale content and uniform porosity, where there is little variation in
the salinity of the interstitial water, the S.P. and the two conventional
resistivity logs gave quick and safe diagnoses. In other regions, however,
variable porosity or salinity and a higher shale content blurred the effect
of the hydrocarbons on the resistivity—a fact that no interpretation,
however summary, may overlook. Although no good quick-look method
has yet been devised that accounts for the shale content, it has been
possible to eliminate the effects of porosity and interstitial water to the
point that, by the 1960 s, some logs presented in the field lent themselves
to immediate interpretation.
The first of these logs, the so-called Rwa log, followed shortly upon
the introduction of the sonic log as a porosity tool. The method consists in
considering all the formations as being entirely filled with water and
(Rwa) of this water on the basis of the
calculating the apparent resistivity
true resistivity (Rt) and the formation factor (F) provided by the sonic
log.22
The apparent resistivity (Rwa) will be the true resistivity (Rw) of the
interstitial water wherever it fills the formation at 100 percent. If, how¬
ever, Rwa is higher that Rw, and if, furthermore, the S.P. maximum
deflections are approximately constant, indicating that Rw is the same over

22 The following relations are used:

0.62
F =-
<£215

where

= A t - A tma
9 A tf- A tna

Atf and Atma are the interval travel times in microseconds per foot (known in principle) in
the fluid and the rock matrix, respectively, and At is the time measured by the sonic sonde;
and

where Rt is given by a deep investigation resistivity device, preferably an induction sonde.

322
After 1945

the whole section logged, then these abnormally high values of Rwa are
indications of oil or gas.
Although a sonde has recently been built that records sonic and
induction logs in a single run, to obtain an Rwa in the field in the 1960’s
required the combined results of two runs. On the first run the sonic log
was recorded optically and on punched tape.23 On the second run, while
the induction was being recorded, the punched tape was played back
through a simplified computer that calculated R wa from R t and A*, the
constant values Afma and \tf being dialed in; and as it was being calculated,
Rwa was put into analog form and also recorded optically. This was a
complicated exercise, requiring equipment so bulky as to call for an
additional pickup truck. The operation put a strain on the equipment, as
well as on the nerves of the crew because of the devilish noise of the paper
tape puncher.
This service was operative for a few years but was active only in
southern Texas and Louisiana.24 It had serious limitations; for example,
the variations of Rw had to be small, the invasion could not be too deep,
and the sonic log was expected to give a good porosity value. But the
major practical difficulty was to obtain a perfect depth matching of both R t
and A Jogs. In brief, there was hardly any justification for such complexity
when the result was merely to indicate hydrocarbon potentialities without
even giving their approximate percentages. More could be achieved at
less cost.
To avoid these inconclusive calculations, Doll thought of comparing
the various logs by overlaying their transparencies. Sliding one log over
another would readily give the optical depth correlation; and by using
logarithmic coordinates for the various resistivities, an immediate evalua¬
tion of the ratio between two values (especially Rxo and Rt) should result
from measuring the separation between curves. It then takes the mere
translation of a log with respect to a fixed grid to multiply or divide all its
ordinates by a constant. This makes it possible to multiply all the values of

23 A conventional recording in binary code with 8 bits, that is, 128 values, a resolution that
is adequate for a sonic log.
24 More specifically, it involved also a so-called FR/FS method consisting of calculating the
formation factor on the basis first of the resistivities (FR) and second of the porosity as
provided by the values of the sonic log (Fs). In principle, FR/FS = 1 in aquifers, and any
value above 1 indicates hydrocarbons.

323
The History of a Technique

an Rxo log (Microlaterolog, Proximity log, micro S.F.L.) by the ratio


RjRmf- According to equation 2 on p. 264:

5W R JR mf 5/8

r Jr w
the new Rxo log will coincide with the Rt log (induction or laterolog) in
water-bearing horizons where S w = 1, whereas its noncoincidence will be a
good hydrocarbon indication: where the curves are separated, it will
suffice to measure the separation with a special rule (allowing for the Ys
exponent), and a first approximation of Sw will be obtained without
further ado.
As only comparables may be compared, the method demands that
the vertical resolution be the same for both logs, which is not generally
the case. To be valid, therefore, the method requires obtaining the
average logRX0—actually a “sliding average” of conductivities—a routine
operation easily carried out during recording.
The logarithmic scale has other merits, timidly acknowledged in the
early days of the hand recorder. It amplifies differences in the low-
resistivity zones where they are of interest, but compresses them wher.e
they are less so. It thus allows for the clear recording on a single curve of a
considerable range of values (from 0.2 to 2000 ohms), whereas the log
would be difficult to read on an linear scale. What had doomed prior
efforts was the fact that to geologists the shapes of the curves were more
important than their numerical values.
However satisfactory quick-look methods have become, and how¬
ever skillful geologists, geophysicists, and petroleum engineers are in
their use, they are only a beginning for an industry capable of assimilating
huge amounts of information. Quantitative interpretation had long been
reserved for specialists: it consists in the application of the proper for¬
mulae to translate physical parameters into petroleum production lan¬
guage. As computations made by slide rule and charts were slow and
arduous, their practice was limited to zones that logs and local stratigraphy
indicated as being promising. Yet a real danger existed that a productive
zone might be overlooked for lack of a systematic study of the log over its
whole length. Computers were now available, however, and most of the
laborious routine work seemed to lend itself to programming. Automated
interpretation was the preocccupation of the day, but far from being easy,

324
After 1945

as some believed, it involved certain thorny problems that required sev¬


eral years of research to master. The first thing necessary to make the log
usable by the computer was to digitize it, that is, to substitute for each
curve a series of numbers recorded on punched cards, punched tape, or
magnetic tape; programs that varied with the logs, the geological features,
and the information sought were then drawn up.
Digitization had two aspects, depending on whether existing or fu¬
ture logs were involved. The first group, available in optical form only, was
usually processed by specialized organizations operating more or less
automated equipment; Schlumberger had recourse to these services, as
did all the oil people. As to the future, the Company had been gaining
experience by recording data in digital form together with some conven¬
tional logs since the mid-1950’s. As a result of this experience, the
dipmeter log, recorded on magnetic tape and processed by computers,
became operational in 1961.
From that time on, abetted by the fad for computers, most oil
companies tried to set up their own automated interpretation programs.
Many interesting publications resulted from their research. But as long as
digital recording in the field remained confined to the dipmeter, progress
continued to be of limited scope. The oil companies were most eager to
see digitization spread to all kinds of logs, but the environment of the oil
well is such that it took Schlumberger years to succeed.
Various decisions, both as to recording technique and format of
magnetic tape, were preceded by prolonged studies, with the participa¬
tion not only of the Houston, Ridgefield, and Paris staffs, but also of
recognized interpretation experts. The service started in the United States
in 1965 and after 1968 became widespread with the introduction of a new
digital recorder. Although the optical log remains in use for quick estima¬
tions on the well site, the magnetic tape is becoming the principal docu¬
ment. In addition to being suited for computer processing, it has substan¬
tial advantages. When it is played back, not in the field but in the silence of
the office, it is possible with the help of special devices to reconstitute an
optical log, whose scales and presentation can differ from the original. On
the other hand, the optical recorder with its nine galvanometers can
record nine parameters at best, whereas the magnetic recorder has up to
22 channels, thus considerably enlarging the scope of combinations that
can be achieved with the downhole equipment.
The magnetic tapes recorded on the well cannot be directly pro-

325
The History of a Technique

cessed by the computer; they first need “editing,” that is, the depth and
unit (ohms/m, grams/cm3, microseconds/ foot, etc.) scales must be recon¬
stituted, and then “normalizing,” in other words they must be corrected
for environmental factors25 (borehole diameter, mud cake, invasion, etc.).
Only after the tapes have been normalized by Schlumberger is actual
interpretation possible; it is for the client to decide whether to undertake
the programming himself or entrust it to Schlumberger. This program¬
ming?6 has been the subject of a vast amount of research and is today the
bailiwick of a highly specialized staff. Yet experience reveals that, in spite
of more and more automation, the operator’s intervention remains ines¬
capable when it comes to basic choices. However, since the computer
willingly accommodates a mass of calculations, one can invoke complex
relations and repetitive processes heretofore unthinkable. By using the
whole gamut of logs, it usually becomes possible not only to evaluate
the porosity, the oil saturation, and the amount of oil flushed from the
invaded zone, but also to estimate the shale content and the density of the
hydrocarbons, and even to draw up a “permeability index” from relations
between capillary pressure, permeability, irreducible water saturation,27
and porosity.
The end product may be of several types. It can be a magnetic tape
called CERT (Computer Evaluation Result Tape), which contains not only
the global results, but also edited and normalized basic data enabling the
client to proceed with his own computations. Most frequently it is an
optical log materializing the results of computer processing, and obtained
through a highly sophisticated electronic camera developed in Ridgefield
(Fig. 86). The conventional aspect of the log is maintained, except that the
parameters provided are no longer couched in the language of ohms and
millivolts, but are expressed in terms of barrels of oil. Today only Hous¬
ton and Paris (Clamart) have the requisite equipment and personnel for
this kind of processing, but local centers with a capacity for increasingly
complex operations are being developed all over the world. With the

20 See p. 227-228. This is the "first phase" of interpretation.


2B SARABAND: A. Poupon, C. Clavier, J. Dumanoir, R. Gaymard, and A. Misk, “Log
Analysis of Sand-Shale Sequences, a Systematic Approach,” Journal of Petroleum Tech¬
nology, July 1970.
CORIBAND: A. Poupon, W. R. Hoyle, A. W. Schmidt, “Log Analysis in Formations with
Complex Lithology,” Journal of Petroleum Technology, August 1971.
27 This is the minimum percentage of water in a reservoir.

326
The History of a Technique

proliferation of computers and constantly improved programming, they


are likely, within a few years, to offer a complete service of automated
interpretation.
Ample advantage has already been taken of the adaptability of
magnetic tape to radio communications for the nearly instantaneous
transmission of offshore logs to the decision centers with an appreciable
saving of time. With the advent of worldwide satellite communications, it
is probable that this method, too, will serve for the transmission of logs
and of their interpretations.

328
Conclusion

Though playing only a limited role in overall geophysical activities


(except in the U.S.S.R.), surface electrical prospecting is still commonly
used in the search for minerals other than petroleum. Many of its methods
have remained basically—but with frequent improvements—those that
Conrad invented during the first quarter of the century. However, his
contribution was made too long ago to find an honorable niche in recent
literature. Had it not been for logging, the name Schlumberger would
hardly emerge from the shadow of bibliographical dictionaries.
Logging was a major success. To be sure, Schlumberger did not
invent all the logs, but what is essential is that the Company brought
logging to the petroleum industry. There was nothing new for physicists
in marking the depths as abscissae and the values of a certain parameter as
ordinates; but the curve that resulted and that gave the measurements
concrete form was a true reflection of stratigraphy, and this was something
genuinely new. To the Texas driller, every log is still a “schlumberger.”
If logging had remained confined to the measurement of resistivities,
it would have become nothing but a remarkable correlation tool. It
achieved its full impact with the contribution of S.P. in 1930, and a highly
productive method was created when the latter was plotted on the same
diagram as resistivity. This was, unknowingly, the first “Synergetic” sys¬
tem: the information provided by the whole exceeded the sum of the data
contributed by each single curve. Similarly, combinations of sondes with
various depths of investigation were soon to provide, in the very frequent
case of sand-shale series, precise answers to the problems of the
operators.
At this point the future of logging was no longer in question. It had

329
The History of a Technique

been good fortune that a technique developed for surface measurements


met the needs of deep exploration. But how would a small company
located in Paris, despite all its pioneering laurels, become aware of, let
alone solve, the many problems that the petroleum industry would
encounter in those far-off poles of its activity that were then the U.S.S.R.
and the United States? The challenge was met because its leaders were
not only researchers but also men of action. They could very rapidly avail
themselves of a vast experience gathered from all over the world; but it
had to be digested, evaluated, disseminated with discretion, and used
wisely for the technical and commercial guidance of the young Company.
During this phase of intense learning and important decisions great help
came from clients prompt to visualize the potentialities of logging, and
appreciative of the strenuous efforts of the Schlumberger international
staff, whether in the field or at the engineering office. Schlumberger’s
place in the petroleum industry was earned by the conception and de¬
velopment of a series of basic inventions fostered by such a climate. It
should be mentioned incidentally that, perhaps for lack of any other
choice, Schlumberger, in delegating almost full autonomy and responsibil¬
ity to its field engineers, was a few decades ahead in the application of
certain tenets of modern management.
Throughout its development Schlumberger earmarked an ever-
increasing share (sometimes deemed excessive) of funds for the design
of tools of the highest standard. The conviction prevailed that, however
reluctantly, petroleum operators would give priority to quality over
economy, and experience confirmed the wisdom of this policy. In keeping
with it and allowing for specific conditions predominant in boreholes,
original concepts shaped the design of every basic tool. Perturbations
arising from the environment were gradually studied so as to eliminate
their effect on the measurements. This took place in the first stage
through departure curves and in the second through the development of
tools that made the corrections automatically and read the real parameters
directly, thereby extending the range of logging to conditions heretofore
unfavorable.
Thus, starting from three cable strands wound on a wooden reel, a
potentiometer, a notebook, and a pencil, logging took less than half a
century to reach the stage of a complex association of mechanics, electric¬
ity. electronics, and automation. Indeed, there have been other such
cases, but few techniques have encountered more obstacles strewn across

330
After 1945

their paths. Born in a country and at a time when specialists in oil


production were very few, tested at Pechelbronn under adverse geologi¬
cal and archaic drilling conditions, promoted in the world by a small group
of engineers who, initially, had only the faintest knowledge of the prob¬
lems and the language of their clients, logging had as its assets the guiding
light of brilliant concepts, the keen perceptiveness of its creators, and a
high sense of teamwork.
Logging—whether by Schlumberger, its competitors, or its clients
—is today a vast industry employing thousands of specialists. One cannot
conceive of any oil company, whatever its size, operating without the
support of its daily harvest of various logs. They are the foundation of a
document that contains every possible bit of information on a well, the
“terminal log,” a wide paper strip showing, side by side, the overall depth
of the well, the parameters recorded as well as the results of interpreta¬
tion, and the data on casing, cementing, perforating, conventional cores
and sidewall samples, cuttings, production tests, and so on. The modest
hand-drawn curves of Pechelbronn have evolved into thick folders filling
drawers and cabinets, from the prefabricated hut in the African jungle or
the Alaskan tundra to the top floor of the skyscrapers where the oil
companies have their executive offices.
Each company disseminates copies of the logs to its geological field
offices, regional offices, central services, and sometimes to its subsidiaries
and partners. Other copies go to the technical secretariat of top manage¬
ment, the drilling and production departments, the research centers, the
economic, statistical, and financial divisions, and finally, in compliance
with the law of most countries, to the authority in charge of mineral
resources conservation. These copies are the basic tool of geologists and
engineers for the computation of reserves and the location of aquifers
and gas caps. The logs are likewise consulted in planning the location
and output of development wells. Drillers refer to them in order to draw
up the technical drilling program for new wells (depth, bit sizes, casing
lengths). They support decision-making in high-level policy meetings of a
company.
When a concern is interested in drilling in a new area, the first thing it
does is to scrutinize the logs of that area or, if there are none, those of the
nearest wells. In the United States most of them are published within a
few weeks and can be obtained in the trade. When an organization is in
the process of resuming operations in a territory already prospected by

331
The History of a Technique

another, the technical arguments put forward in the financial discussions


rest on seismic surveys for surface work and on logs for deep exploration.
When an independent American producer seeks financing, he submits the
logs to bankers to stress the potential return on the investment. In
addition, the banks have their own log libraries (often on microfilm),
where their petroleum advisers seek the documentation they need for
investment decisions.
The innumerable uses of logging should not obliterate what has
perhaps been its greatest contribution to the development of the petro¬
leum industry. Closely tied to the rotary drilling system, in fact designed
and developed for its needs, logging and its auxiliary activities have
become intrinsic elements in its progress. Needless to say, logging is not
the sole source of information for deep exploration, and, especially in
wildcatting, coring remains crucial. Nevertheless the driller, relieved of
the constant burden of evaluating the formations through which he is
drilling, is able to concentrate on the technique and efficiency of the
drilling proper, in the certainty that methods exist through which all the
formations can be detected and correlated with those of nearby wells,
the potential reservoirs evaluated, and the formations and fluids they
hold sampled. It is fair to say that, by relieving drilling of the dis¬
ruptive operations for which it was ill fitted, logging has contributed
to the remarkably high rates of penetration of modern wells.
At the service of an industry that still has vast areas open for explora¬
tion, notably the barely touched resources of the deep seas and the Arctic
regions, Schlumberger must actively continue its research and develop¬
ment efforts along the same general lines as were pursued for the past two
decades. The development of tools for the recording in a single run of the
greatest possible number of parameters will be carried on, the ultimate
objective being to record them all (i.e., about 15) under the increasingly
severe constraints of ever-greater drilling depths. Magnetic tape recorders
will be provided to all crews; in addition, they will be equipped with
computer systems for the simplification and automation of otherwise
complex adjustment processes, a heavy burden on field engineers.
With the ever-expanding use of computers, interpretation has
reached a high level of efficiency over the past 15 years. However, the
intervening factors are so many and so diverse that it will take a long time
to complete current research and realize mathematical programs the out¬
put of which will clearly indicate the relative impacts of the various factors

332
After 1945

involved. It will be recalled also that in many cases the parameters


measured are incomplete sources of information (shale content of reser¬
voirs) or are exceedingly vague (permeability, location of fractures before
completion). However uncertain the prospects of success, the possibility
should not be excluded that these gaps will be filled some time in the
future by inventions as significant as S.P. or induction logging.
Further progress in telemetry will make possible the transmission of
logs between wells, especially offshore, and a growing number of centers
equipped for computerized interpretation. Nonetheless, whatever the
speed and flexibility of such transmissions (in any case it may be some
time before they are introduced into a number of developing countries),
field interpretation will remain indispensable for such urgent decisions
as whére to core or take fluid samples, when to lower a casing, how to
select intervals for perforation, and, more importantly, when to continue
or discontinue drilling. This is the reason why the effort is made to
provide the crews with processing systems that give more thorough
interpretations than the quick look.
Whereas the extrapolation of current trends gives some hint of
short-term developments to come, any longer range forecast would be
risky. Twenty-five years ago, before the discovery of transistors with¬
standing high temperatures, it seemed unlikely that Schlumberger would
some day switch from pulsated current to electronic solutions. Likewise,
nobody could foresee from the first-generation computers the role that
informatics would play in measuring and interpretation techniques. There
are many similar examples. What can be safely stated is that all the
technological developments of the future will be drawn upon by Schlum¬
berger to maintain its lead in the field.
In conclusion, it may be fitting to add that Schlumberger’s technical
and industrial success is also, in a way, a human achievement. From the
outset, a certain spirit motivated a staff of young graduates fresh out of
school: fellowship, the attraction of distant horizons, the feeling of par¬
ticipation in an effort where a common task did not stifle individualism.
This rather exceptional atmosphere, although undoubtedly somewhat
attenuated by years of growth, remains alive. Over the half century since
Schlumberger became a leader in geophysics, it has grown without labor
disputes, while avoiding paternalism in the relationship between man¬
agement and staff. It is to be hoped that the future will preserve this
human factor, without which no technique can make history.

333
Date Due

■3 s; ,m A
Sir
’ [ t1 ' w
if'* % "jf *
4 L i 31L
t ■
4
k/ *//#

W/
T fl

w Oh ' fi
w 'S81
-o
ISBN 0-471-01687-5

You might also like