Anda di halaman 1dari 328

Philosophy of Psychology

Mario Bunge and Ruben Ardila

Philosophy of Psychology

With 33 Illustrations

Springer-Verlag
New York Berlin Heidelberg
London Paris Tokyo
Mario Bunge Ruben Ardila
Foundations and Philosophy Departamento de Psicologia
of Science Unit Universidad Nacional de Colombia
McGill University Bogota
Montreal H3A lW7 Colombia
Canada

Library of Congress Cataloging in Publication Data


Bunge, Mario Augusto.
Philosophy of psychology.
Bibliography: p.
Includes indexes.
I. Psychology-Philosophy. I. Ardila, Ruben,
1942- II. Title.
BF38.B85 1987 150'.1 86-26124

© 1987 by Springer-Verlag New York Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer-Verlag, 175 Fifth Avenue, New York, New
York 10010, USA), except for brief excerpts in connection with reviews or scholarly analy-
sis. Use in connection with any form of information storage and retrieval, electronic adapta-
tion, computer software, or by similar or dissimilar methodology now known or hereafter
developed is forbidden.
The use of general descriptive names, trade names, trademarks, etc. in this publication,
even if the former are not especially identified, is not to be taken as a sign that such names,
as understood by the Trade Marks and Merchandise Marks Act, may accordingly be used
freely by anyone.
While the advice and information in this book are believed to be true and accurate at the date
of going to press, neither the authors nor the editors nor the publisher can accept any legal
responsibility for any errors or omissions that may be made. The publisher makes no
warranty, express or implied, with respect to the material contained herein.

Typeset by TCSystems, Shippensburg, Pennsylvania.

987654321

ISBN-13: 978-1-4612-9118-3 e-ISBN-13: 978-1-4612-4696-1


DOl: 10.1007/978-1-4612-4696-1
Preface

This book is about some topical philosophical and methodological prob-


lems that arise in the study of behavior and mind, as well as in the
treatment of behavioral and mental disorders. It deals with such questions
as 'What is behavior a manifestation of?', 'What is mind, and how is it
related to matter?', 'Which are the positive legacies, if any, of the major
psychological schools?', 'How can behavior and mind best be studied?',
and 'Which are the most effective ways of modifying behavioral and
mental processes?'
These questions and their kin cannot be avoided in the long run because
they fuel the daily search for better hypotheses, experimental designs,
techniques, and treatments. They also occur in the critical examination of
data and theories, as well as methods for the treatment of behavioral and
mental disorders.
All students of human or animal, normal or abnormal behavior and
mind, whether their main concern is basic or applied, theoretical or em-
pirical, admit more or less tacitly to a large number of general philosophi-
cal and methodological principles. For example, they presuppose that
mind is (or is not) distinct from brain function; that understanding the
nervous system is (or is not) necessary to explain behavior and mind; that
animal research is (or is not) necessary to advance the understanding of
human behavior and mind; that statistics are (or are not) indispensable to
assess the efficacy of treatments of behavioral and mental disorders; that
psychology is (or is not) an autonomous discipline; that psychology has
much (or little) to learn from artificial intelligence-and more.
Some of these strategic principles guide research or practice, whereas
others mislead them. As long as they remain tacit they are mere dogmas.
Whereas some such dogmas may be fertile, others are bound to be barren
or even harmful to the search for truth and efficacy. A principle used in
scientific research or in professional practice becomes a hypothesis the
moment it is rendered explicit. From then on it can be subjected to exami-
nation and evaluation, whereas before that it was out of consciousness,
hence beyond criticism.
VI Preface

Explicit principles, in short, are not only guides to research and prac-
tice. They can also become objects of research, in particular of concep-
tual analysis, theoretical systematization, and empirical checking. A goal
of the present study is to ferret out and examine some of the philosophical
hypotheses and methodological rules held or used more or less tacitly by
contemporary psychologists.
Our exercise is not one in academic futility: it should be of some help to
psychologists as well as philosophers. To the former because bad princi-
ples, particularly when hidden, are roadblocks, whereas good ones expe-
dite research and praxis, and they can occasionally reorient them in prom-
ising directions. Our exercise should be useful to philosophers because
the philosophy of mind will continue to be obsolete, boring, and barren, as
long as it remains out of touch with the forefront of research and practice.
Our book, then, is not one of philosophical or armchair psychology but
a work in the philosophy and methodology of psychology. We do not wish
to usurp the job of psychologists but to study it from a certain viewpoint.
In fact, our task will be to analyze psychological research and practice in
the light of philosophy and methodology, and with the hope that such
examination will in turn enrich both philosophY and psychology. We
agree that philosophical psychology is at best the precursor and at worst
the enemy of scientific psychology, but submit that the philosophy of
psychology can be its ally.
This work is the outcome of the joint effort of a research psychologist
(R.A.) and a physicist turned philosopher (M.B.). The former wrote chap-
ters 10 and 12, and the senior author wrote the rest. Each author takes full
responsibility for his own contribution, and neither endorses fully that of
his partner.
The authors undertook this venture on the strength of five beliefs. (1)
Psychology has an extremely rich but largely untapped philosophical and
methodological problematics. (2) Some of the philosophical and method-
ological principles at work in psychology are tacit, and hence are held
somewhat uncritically. (3) All the principles that guide or misguide re-
search and practice in any field should be subjected to thorough investiga-
tion. (4) Because such investigation bears on norms that concern both
research and practice, it should be taken seriously by all students of
behavior, mind, and mental health. (5) Psychologists can make solid con-
tributions to such philosophical and methodological studies provided they
become reasonably well acquainted with contemporary philosophy, and
philosophers can do as much as long as they become reasonably conver-
sant with contemporary psychology. This being a tall order, it is best for
psychologists and philosophers to cooperate with one another.
Mario Bunge Ruben Ardila
Foundations and Philosophy Departamento de Psicologia,
of Science Unit, Universidad Nacional de Colombia,
McGill University Bogota, Columbia
Montreal, Canada
Acknowledgments

I am indebted to the Social Sciences and Humanities Research Council of


Canada for a research grant in support of this project. I thank the follow-
ing persons for having supplied valuable information, comments, or criti-
cisms on a variety of psychological or neuroscientific problems over the
last few years: Ruben Ardila (Psychology, Universidad Nacional de Co-
lombia), the late Dalbir Bindra (Psychology, McGill University), David
Blitz (Philosophy, Concordia University), Bernard Dubrovsky (Psychia-
try, McGill University), Mike Dillinger (Educational Psychology, McGill
University), Hans Flohr (Neurobiology, Universitat Bremen), Lluis
Garcia i Sevilla (Psychology, Universidad de Barcelona), the late Donald
o. Hebb (Psychology, McGill University), Rodolfo Llinas (Physiology
and Biophysics, New York University), Peter M. Milner (Psychology,
McGill University), Mortimer Mishkin (Neurobiology, National Institute
of Health, Bethesda), Meinrad Perrez (Psychology, U niversite de Fri-
bourg), Ernst Poppel (Medical Psychology, U niversitat Munchen), Viktor
Sarris (Psychology, J.W. Goethe-Universitat, Frankfurt), and Endel
Tulving (Psychology, University of Toronto).

M.B.
Contents

Preface..................................................... v
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vll

I PRELIMINARIES

WHY PHILOSOPHY OF PSYCHOLOGY. . . . . . . . . . . . . . . . . 3

1.1 Influence of Philosophy on Psychology. . . . . . . . . . . . . . . 4


1.2 Philosophies of Mind. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 The Identity Hypotheses. . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Philosophical Presuppositions of Scientific Research. . . 17
1.5 Philosophies of Psychology ......................... 21
1.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2 WHAT PSYCHOLOGY IS ABOUT.. ... ... .. .. ... . .. . ... . 25

2.1 Definitions of Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . 26


2.2 Referents of Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3 The Fragmentation of Psychology and How to Remedy It 30
2.4 Unification in Action. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5 Aims of Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

II APPROACH AND METHOD

3 APPROACHES TO BEHAVIOR AND MIND. . . . . . . . . . . . . . 43

3.1 Approach......................................... 44
3.2 Atomism, Holism, and Systemism . . . . . . . . . . . . . . . . . . . 45
3.3 Nonscientific Approaches to Psychology. . . . . . . . . . . . . . 48
3.4 Toward a Scientific Psychology. . . . . . . . . . . . . . . . . . . . . . 51
x Contents

3.5 Scientific Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56


3.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4 METHODOLOGy....................................... 62

4.1 Method........................................... 63
4.2 Observation ....................................... · 66
4.3 Measurement...................................... 71
4.4 Experiment........................................ 76
4.5 Inference.......................................... 81
4.6 Summing Up . . . . .. . . .. . . .. . . . . . . .. . . .. .. .. . .. . ... . 85

III BRAINLESS PSYCHOLOGY

5 MENTALISM.......................................... 89

5.1 Subjective Experience. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90


5.2 Classical Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.3 Gestalt Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 100
5.4 Information-Processing Psychology. . . . . . . . . . . . . . . . .. 105
5.5 Pop Psychology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 111
5.6 Summing Up . . .. . . . . . . .. . . .. .. . ... .. . .. . .. .. ... . .. 115

6 BEHAVIORISM........................................ 116

6.1 Phenomenalism (Black-Boxism) ..................... 117


6.2 Environmentalism.................................. 120
6.3 Operationism...................................... 125
6.4 Intervening Variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 130
6.5 Hypothetical Constructs. . . . . . . . . . . . . . . . . . . . . . . . . . .. 132
6.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 134

IV BIOPSYCHOLOGY

7 NEUROBIOLOGy...................................... 139

7.1 Brain & Co........................................ 141


7.2 Plasticity.......................................... 146
7.3 Development...................................... 151
7.4 Evolution......................................... 155
7.5 Functional Localization. . . . . . . . . . . . . . . . . . . . . . . . . . . .. 160
7.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 165
Contents xi

8 BASIC FUNCTIONS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 166

8.1 Movement........................................ 167


8.2 Affect............................................ 169
8.3 Sensation......................................... 173
8.4 Attention.......................................... 177
8.5 Memory.......................................... 179
8.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 184

9 HIGHER FUNCTIONS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 185

9.1 Learning.......................................... 186


9.2 Perception........................................ 194
9.3 Conception........................................ 201
9.4 Cognition......................................... 207
9.5 Intention.......................................... 215
9.6 Summing Up . . .. . . .. . . . . . . . .. . .. .. .. . .. . .. . . . . . . .. 218

V THE SOCIAL ASPECT

10 THE SOCIAL MATRIX OF BEHAVIOR. . . . . . . . . . . . . . . . .. 221

10.1 Psychology: Natural Science or Social Science? . . . . . .. 223


10.2 Culture........................................... 224
10.3 Social Classes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 227
10.4 Socialization....................................... 228
10.5 Cultural Homogenization . . . . . . . . . . . . . . . . . . . . . . . . . .. 231
10.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 232

11 CONSCIOUSNESS...................................... 233

11.1 Distinctions....................................... 234


11.2 Definitions........................................ 235
11.3 Applications....................................... 239
11.4 Hypotheses....................................... 243
11.5 Experimental Tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 247
11.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 250

12 PSYCHOTECHNOLOGY................................ 251

12.1 Clinical Psychology and Psychiatry .. . . . . . . . . . . . . . . .. 252


12.2 Educational Psychology ............................ 255
12.3 Industrial and Organizational Psychology. . . . . . . . . . . .. 257
12.4 Designing Cultures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 258
xii . Contents

12.5 The Goals of Psychotechnology ..................... 260


12.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 261

VI CONCLUSION

13 CONCLUDING REMARKS. . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 265

13.1 Reduction......................................... 265


13.2 Integration........................................ 270
13.3 Explanation....................................... 274
13.4 Prospects......................................... 280
13.5 Philosophical Harvest . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 282
13.6 Summing Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 284

References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 287
Name Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 309
Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 317
I

Preliminaries
CHAPTER 1

Why Philosophy of Psychology?

Originally philosophy encompassed all knowledge, and philosophers were


polymaths. For example, Aristotle worked on problems in physics, biol-
ogy, psychology, and political science, as well as in logic and ethics; and
Descartes was interested in mathematics, physics, biology, and psychol-
ogy as well as in philosophy proper. Nowadays philosophy is a branch of
the humanities, and philosophers confine their attention to conceptual
problems of a certain kind. They do not pass judgment on special matters
of fact, which they gladly leave in the hands of scientists and technolo-
gists.
Contemporary philosophy may be regarded as consisting essentially of
the following disciplines: logic, which is also part of mathematics; seman-
tics, or the study of sense, reference, interpretation, and truth; epistemol-
ogy, or the theory of knowledge and general methodology; ontology, or
the theory of the most basic and pervasive features of the world; and
ethics, or the theory of good and of right conduct.
There are several philosophical styles. The most popular way of philos-
ophizing is to reflect on certain general problems, such as 'What is
mind?', using a mixture of ordinary knowledge (e.g., lay psychology),
scraps of our philosophical heritage, and logic. This style has no appeal
for scientists. Piaget (1971) called it 'autistic philosophY.' In order for a
philosophy to be of some use for science, it must be intelligible (if possible
exact) and compatible with science. For instance, a philosophy of mind
should make use of contemporary psychology as well as of tools of con-
ceptual analysis.
Psychology used to be a branch of philosophy, from which it is said to
have gained independence in about 1850 with the birth of psychophysics.
Why then should contemporary psychologists be bothered with philoso-
phy? Because, whether or not they know or like it, psychologists hold and
utilize a number of philosophical ideas, particularly ideas on the nature of
mind and science. Every psychologist is thus not only a scientist or a
practitioner but also an amateur philosopher, usually malgre lui. Which
should worry nobody except for the fact that tacit knowledge is half-
4 1. Why Philosophy of Psychology?

baked and inconsistent, often obsolete, and never in sight for critical
scrutiny.
There is a further reason for dealing explicitly with the psychology-
philosophy connection, namely that philosophers consume psychological
products-alas, seldom fresh ones. In fact, nearly every philosopher of
mind is primarily indebted to folk psychology-the ordinary and intuitive
knowledge of self and others-and secondarily to the findings, genuine or
bogus, of previous generations, more often than not armchair psycholo-
gists. Three instances of this regrettable habit will suffice to make this
point.
Ryle's once famous The Concept of Mind (1949) relies exclusively on
radical behaviorism, then a novelty in Great Britain. Strawson's philoso-
phy of mind, in his influential Individuals (1959), boils down to the medie-
val thesis that a person is a compositum of body and mind-with no
precise indication as to either the nature of the components or their mode
of composition. And Popper's contribution to his famous joint book with
Eccles (1977) is a direct descendant of Descartes's interactionist mind-
body dualism, it does not analyze any of the key concepts involved, it
takes no notice of physiological psychology, and it defies the law of con-
servation of energy. Other philosophers have been taken in by Freud's
amusing stories and speculations, or even by Lacan's rhetoric. The list of
philosophers familiar with the contemporary psychological literature oc-
cupies perhaps one line.
In sum, psychology and philosophy interact vigorously, though usually
with a long time lag, in a clandestine fashion, and seldom to one another's
benefit. The same holds for other sciences, particularly mathematics,
physics, biology, and social science. The clearer we become about such
irregular interactions, the better should we be able to control them for the
good of the parties concerned. Such control should result, in particular, in
having science and philosophy march in step and exchange fertile in-
sights.
The present chapter is devoted to substantiating the claim that psychol-
ogy includes some philosophy, and to sketching the kind of philosophy we
deem suitable to promote psychological research and practice. Such a
philosophy will prove be the one centered in the general principles uti-
lized, more or less explicitly, in the more mature sciences.

1.1 Influence of Philosophy on Psychology


Philosophy enters psychology in two ways: through hypotheses concern-
ing the nature of mind and the proper manner of studying it, and through
general principles underlying scientific research in any field. Let us start
with the former, leaving the latter for section 104.
If mind is regarded as an immaterial entity-that is, if the idealist or
1.1. Influence of Philosophy on Psychology 5

spiritualist doctrine of mind is adopted-then mentalist psychology is


bound to be produced. The aim of such study is said to be the description
of states of mind, in particular of the stream of consciousness, as well as
the possible influence of mental states on bodily states. The bulk of classi-
cal psychology was of this kind: mentalistic and motivated by philosophi-
cal idealism.
Behaviorism evolved largely as a reaction against mentalism and in
close association with positivism, a variety of empiricist philosophy. It
denied the existence of mind (ontological behaviorism) or at least the
possibility of studying it scientifically (methodological behaviorism).
Moreover, it undertook to study overt behavior in a rigorous manner,
using the scientific (and in particular the experimental) method. However,
in common with mentalism, behaviorism paid no attention to the nervous
system: it focused on the natural environment (biological behaviorism) or
on the social one (social behaviorism). Consequently, although it did
attempt to explain behavior, it only succeeded in describing it.
Psychobiology differs from both mentalism and behaviorism without,
however, being totally alien to either. Indeed, it shares with the former
the belief in the existence of mental states, and with the latter the need to
do research in a scientific way. Psychobiology assumes that behavior is an
outcome of neural processes sometimes triggered by external stimuli,
whereas mental states are brain states of a very special kind. The latter
thesis, though strongly supported by contemporary physiological psy-
chology, originated in ancient Greece: it was the view of Alcmaeon,
adopted by Hippocrates. And the less precise thesis, that the mind is not a
separate stuff but a state of matter, is common to all materialist philoso-
phies. We shall come back to this in Section 1.2.
In sum, pmlosophy is a source of inspiration, good or bad but unavoid-
able, to psychology. See Table 1.1.
Philosophy has been more than a source of inspiration to psychology:
on occasion it has been an obstacle. For example, Kant and his influential
nineteenth-century followers grouped into the historico-cultural or hu-
manistic school decreed that psychology cannot be a natural science, that
it is a spiritual one (a Geisteswissenschaft) along with the social sciences.
(The family of the spiritual sciences, also called 'moral sciences', coin-
cides roughly with what the behaviorists call 'behavioral sciences. ')
The sciences of the spirit (or mind) were deemed to be nonexperimental
and nonmathematical, and they were placed in the humanities, because
their study required only books, and their teaching did not even call for
blackboards. The aim of such disciplines was said to be to describe and
understand empathically (i.e., uerstehen), not to explain (erkliiren) or
predict with the help of objective laws, since the spirit (Geist) was taken
to be immaterial and lawless. This philosophy is still much in evidence in
some contemporary schools, particularly humanistic psychology, psycho-
analysis, and to some extent Chomsky's psycholinguistics as well. All of
6 1. Why Philosophy of Psychology?

TABLE 1.1. Philosophy as a source of inspiration to psychology (1/1).


Idealism Positivism Materialism
mentalism behaviorism psychobiology
Behavior Byproduct of mind Response to external Response to both
stimuli external and inter-
nal stimuli
Mind Separate immaterial Nonexistent or at Collection of brain
entity least beyond the processes of a
reach of science special kind
Aim of I/J Description of mental Description, explana- Description, explana-
processes and their tion, prediction, tion, prediction,
bodily effects and modification of and modification of
behavior behavioral and
mental processes
Method of I/J Introspection, direct Observation, experiment, and mathematical
or indirect modeling, as well as statistical control
Status of I/J Branch of philosophy Branch of biology or Branch of biology and
or autonomous of social science social science
science
Motto I think, therefore I You behave, there- We exist, therefore
am. fore you are. we behave and
think.

these deal in immaterial minds and therefore shun experiment and avoid
biology even though they occasionally pay lip service to both.
The humanistic (or spiritualist, or historico-cultural, or historicist)
school has delayed the study of human beings, particularly for erecting a
barrier between humans and nature-or rather by importing that barrier
from Christian theology. The barrier has been crumbling from the moment
it was built. In fact, a number of flourishing scientific disciplines violate
the interdiction to study mind and society employing the scientific
method: physiological psychology (or psychobiology), experimental lin-
guistics, neurolinguistics, anthropology, and others bear witness to this.
However, this obituary would be incomplete and unfair if we neglected
to mention that the humanistic school was right in one important point,
namely in maintaining that the possession of a "spirit" (in contemporary
parlance, "highly evolved brain") puts humans in a very special category,
because it gives them the possibility of fashioning complex material and
conceptual artifacts, as well as a complex artificial environment com-
posed by an economy, a polity, and a culture. (In turn, this artificial
environment, that is, society, molds behavior and mentation.) This im-
plies that biology, though necessary, is insufficient to account for human
nature. To put it positively: Because human nature is not fully natural but
partly artificial (i.e., human-made), the study of humankind concerns not
only natural science but also social science. However, both kinds of study
are methodologically alike.
We must grant, then, that humankind possesses properties and satisfies
regularities (laws and rules) that single it out from the rest of nature. But
1.2. Philosophies of Mind 7

at the same time we can argue that such emergent properties and regulari-
ties do not free humans from the laws of biology and do not make them
unfit as a subject of scientific investigation. In other words, we may admit
the idealistic view of the singularity of human beings provided we conjoin
it with the following theses about those emergent features: (a) far from
being miraculous, they are the outcome of a long evolutionary process
involving only material factors, and (b) far from defying science, they can
be studied scientifically. Thesis (a) belongs in emergentist materialism
(though not in physicalism or vulgar materialism), and thesis (b) is part of
scientific realism. Because materialism is an ontological doctrine, and
realism an epistemological one, we can see that it is not philosophy as
such, but certain philosophies, which are inimical to the scientific study of
man and, in particular, to scientific psychology. Here, as elsewhere, one
grief cures another.

1.2 Philosophies of Mind


Psychology is so close to philosophy that every psychologist, no matter
how indifferent or even hostile he or she may feel toward philosophy,
cannot help holding some philosophy of mind or other. Whereas in excep-
tional cases this philosophy may be an outcome of reflections on scientific
findings, in most cases it is learned from teachers, colleagues, or publica-
tions. After all, no psychologist can escape tradition, which is composed
of a host of old, some even archaic, views on the so-called Big Questions,
among which that of the nature of mind stands out. (See Boring, 1950;
Hearst, 1979; Whetherick, 1979.)
Most philosophies of mind have been proposed by philosophers and
theologians over the past three millennia. Everyone of those philosophies
proposes its own solution to the mind-body problem, that is, to the ques-
tion: What is mind and how is it related to matter, in particular the body?
This question, once the exclusive property of theologians and philoso-
phers, is now being investigated by scientists as well. It is, then, alongside
a few other problems, such as "Which is the good society?", in the
intersection of science, philosophy, and ideology. Like others of its kind,
the problem can be handled scientifically, philosophically, or ideologi-
cally (in particular theologically). And, as in similar cases, every pro-
posed solution to the problem, and every argument about it, is likely to
elicit spirited reactions. As one distinguished psychologist put it, the mere
invitation to discuss the mind-body problem seems to activate mainly the
limbic system, even in otherwise sober scientists.
The various philosophies of mind can be grouped into two large fami-
lies: psychophysical monism and psychophysical dualism. Monism as-
serts that matter and mind are, in some sense, one; on the other hand
dualism holds that matter and mind are distinct kinds of stuff. However,
8 1. Why Philosophy of Psychology?

neither of these families of doctrines is homogeneous: each is composed


of at least five mutually incompatible views. We have summarized them in
Table 1.2.
The two families of proposed solutions to the mind-body conundrum
have a number of remarkable features. One of them is that the monist-
dualist division does not coincide with the classical idealism-materialism
dichotomy. In fact, both the monist and the dualist camps include ideal-
ists as well as materialists. For example, Plato and Hegel were idealists,
but whereas the former was a dualist the latter was a monist; and Darwin,
Vogt, Biichner, and Moleschott were materialists and epiphenomenalists
at the same time, for holding that the brain secretes thought much as the
liver secretes bile.
A second notable feature of the monist-dualist dichotomy is that it is
independent ,of epistemological questions. In particular, it does not coin-
cide with either the subjectivist-realist or the empiricist-rationalist divi-

TABLE 1.2. The ten major views on the mind-body problem.


lsychophysical monism Psychophysical dualism
MI Idealism, panpsychism, and phe- Dl Autonomism: cp and 1/1 are mutually
nomenalism: Everything is 1/1 independent (Wittgenstein).
(Berkeley, Fichte, Hegel, Fech-
ner, E. Mach, the later W. James,
A. N. Whitehead, Teilhard de
Chardin, B. Rensch).
M2 Neutral monism, or the double D2 Parallelism: cp and 1/1 are parallel or
aspect doctrine: cp and 1/1 are so synchronous (Leibniz, R. H.
many manifestations of a single Lotze, W. Wundt, J. H. Jackson,
unknowable neutral substance the young Freud, some Gestalt-
(Spinoza, at one time W. James ists).
and B. Russell, R. Carnap, M.
Schlick, and H. Feigl).
M3 Eliminative materialism: Nothing is D3 Epiphenomenalism: cp produces or
1/1 (J. B. Watson, B. F. Skinner, causes 1/1, which does not react
A. Turing). back upon cp (Hobbes, C. Vogt,
T. H. Huxley, C. D. Broad, A. J.
Ayer).
M4 Reductive or physicalist material- D4 Animism: 1/1 animates, controls,
ism: 1/1 states are cp states (Epi- causes or affects cp, which does
curus, Lucretius, Hobbes, La not react back upon 1/1 (Plato,
Mettrie, d'Holbach, I. P. Pavlov, Augustine, computationalist
K. S. Lashley, J. J. C. Smart, D. cognitive psychology, according
Armstrong, W. V. Quine). to which people are run by imma-
terial programs).
M5 Emergentist materialism: 1/1 is a very D5 Interactionism: cp and 1/1 interact,
special biofunction (Diderot, S. the brain being only the "material
Ram6n y Cajal, T. C. Schneirla, basis" of mind (Descartes, W.
D. Hebb, A. R. Luria, D. Bindra, McDougall, the mature Freud,
V. Mountcastle, J. Olds, H. W. Penfield, R. Sperry, J. C.
Jerison). Eccles, K. R. Popper, N.
Chomsky).

Note: <p stands for body (or the physical) and tjJ for mind (or the mental). Only a few well-known
supporters of each view are mentioned. Adapted from Bunge (1980).
1.2. Philosophies of Mind 9

sion. Thus, whereas Ayer and Quine are empiricists, the former is a
dualist and the latter a physicalist; on the other hand both Popper and
Smart are realists, but the former is a dualist whereas the latter is a
monist. (Recall Table 1.2.)
A third feature to be noted is that most of the philosophies of mind are
sketchy and consequently subject to much unilluminating and inconclu-
sive hermeneutic controversy. (In such an argument, commentators or
critics A and B disagree about what author C "really wanted to say." On
the other hand, in a scientific controversy people disagree on the worth of
a research plan, or the reliability of a method, or the truth of a datum or a
theory-and they are supposed to produce some evidence for their
views.) Let us recall three important cases offuzzy philosophies of mind.
The great Aristotle's view on the mind-body problem was so imprecise
that we have been unable to place it in Table 1.2. (The same holds for
Kant's.) On the one hand he criticized Plato's idealism and nativism,
formulated an empiricist view of learning, and stated clearly that the
human mind has no independent existence but is the "form" of the body.
But on the other hand he filled the body with "animal spirits" and he
admitted the existence of supernatural entities. This ambiguity gave rise
to a split among his numerous followers. There were monists and crypto-
materialists like Theophrastus and Averroes, as well as dualists like Saint
Thomas Aquinas and most of the other schoolmen.
Another case of ambiguity is that of Lenin (1947), who believed himself
to be a materialist but tripped when he stumbled on the mind-body prob-
lem. In fact, he took the materialist philosopher J. Dietzgen to task for
holding that thought is material, and held instead that the mental is the
opposite of the material. He argued that, if this opposition were denied,
there would be no contrast between idealism and materialism. (It would
seem that in this case Lenin sacrificed materialism at the altar of dialec-
tics.) His influence on present-day Marxist philosophy is such that psy-
chophysical dualism seems to be the most popular philosophy of mind in
the nations where Marxist-Leninist philosophy prevails-as elsewhere.
Thus the renowned Soviet historian of psychology Jarochewski (1975, p.
168) rejects as "vulgar materialism" the thesis that "identifies conscious-
ness with physiological processes in the brain."
A third interesting instance of fuzziness is Popper's three worlds doc-
trine (Eccles & Robinson, 1985; Popper & Eccles, 1977). According to it,
reality is composed ofthree "worlds": World 1 (physical), World 2 (sub-
jective experience), and World 3 (culture). World 1 is material, World 2
immaterial, and World 3 a mixed bag of material objects, such as books,
and immaterial ones such as the "contents" of books. Worlds 1 and 2
interact (at the so far unidentified "liaison brain" according to Eccles);
and World 3, a product of World 2, reacts back on World 2. At no place in
their writings on this doctrine do Popper or his coworkers or followers
bother to elucidate (e.g., define) any of the key concepts occurring in the
10 1. Why Philosophy of Psychology?

new trinity. In particular, they do not tell us (a) what kind of objects their
"worlds" are: sets, variable collections, aggregates, or systems; (b) what
a mental state is-except that it is not the state of some concrete thing;
and (c) what the mechanism of the mind-body interaction could be-save
the suggestion that it might be a case of telekinesis.
Interactionist dualism is then as woolly today as it was when Descartes
(1649) explained it. (Actually it is even fuzzier, because Descartes stuck
out his neck and conjectured that the pineal gland was the meeting place
of mind and body, whereas Eccles is still in search of his' 'liaison brain. ")
What is clear about the Popper-Eccles philosophy of mind is this. First, it
is half-baked because its key concepts-notably those of world, mind,
and interaction-are undefined and it does not contain any precise hy-
potheses about the nature of mind or its alleged interaction with the brain.
Second, it violates "a fundamental tenet of physics," namely the princi-
ple of conservation of energy (because it postulates that immaterial mind
can move matter). Third, it violates a tacit assumption of all experimental
science, namely that the mind cannot act directly upon matter-for, if it
could, no instrument reading would be worth anything. Fourth, it as-
sumes that mental states and processes are unlike any other states and
processes, in that they are not states of things or processes in things-
whence it perpetuates the ontological anomaly of classical psychology.
Fifth, it is inconsistent with the tacit presupposition underlying physiolog-
ical psychology, namely that mental states are brain states. Sixth, it is
inconsistent with evolutionary biology, which acknowledges only mate-
rial things. Seventh, the doctrine calls for a bit of parapsychology, namely
the conjecture that the mind is to the brain as the performer to the piano
keyboard (Eccles's metaphor). Eighth, although the doctrine fits in nicely
with mainstream Christian theology, it has been used to accuse material-
ists of dogmatism and even of confusing their science with their religion
(Eccles & Robinson, 1985, p. 36).
Most but not all of the philosophies of mind are sketchy and vague.
Emergentist materialism and its scientific companion, psychobiology, are
fairly elaborate, include some mathematical models, and enjoy strong
experimental support. (See Bindra, 1976; Bunge, 1980, 1981; Dimond,
1980; Edelman & Mountcastle, 1978; Hebb, 1949, 1980; Milner, 1970;
Thompson, 1975; Vttal, 1978). Moreover, unlike dualism and idealistic
monism, emergentist materialism does not postulate the existence of an
immaterial stuff, that is, one exempt from natural law as well as inaccessi-
ble to experimental manipulation. In short, unlike dualism and idealistic
monism, emergentist materialism keeps psychology witbin the fold of
science instead of encouraging it to go back to philosophy or theology.
But, unlike eliminative materialism and physicalist or vulgar materialism,
emergentist materialism admits the specificity of the mental and, conse-
quently, the need to investigate it using the methods of psychology in
addition to those of neurophysiology. (See chap. 13.)
1.2. Philosophies of Mind II

So much for the sketchiness and vagueness of most philosophies of


mind. A fourth feature of this family of doctrines is that most of its
members are stray: that is, they are seldom components of comprehen-
sive worldviews or ontological (metaphysical) systems. (This was cer-
tainly not the case of Aristotle or Leibniz, who were systematic thinkers,
but it is the case of nearly all contemporary philosophers, who are typi-
cally specialists rather than generalists.)
The isolation of a philosophy of mind from a comprehensive body of
philosophical hypotheses about the world has the disadvantage that it
gives speculation free rein. Every product of such wild speculation is
bound to remain vague and weak-vague because it employs certain basic
notions, such as those of thing, property of a thing, state of a thing,
process, matter, space, time, causation, and chance, without explicating
them: it is forced to borrow them from ordinary knowledge, which is
fuzzy, inconsistent, and largely fossilized knowledge. And the view is
bound to remain weak because it lacks the support of other branches of
ontology, in particular those that inquire into the most general features of
organisms and societies. In sum, a necessary condition for a philosophy
of mind to make full sense is that, far from being the outcome of an
isolated exercise, it be a chapter of a comprehensive, clear, and consis-
tent worldview or ontology. In turn, a necessary condition for such an
ontology to provide useful insights and to be appealing to scientists is
that, far from being alien to contemporary science, it should harmonize
with it. This leads us to a fifth feature of most philosophies of mind.
Most philosophies of mind are speculative and out of touch with psy-
chological research, so that they often read like medieval texts. For exam-
ple, Wittgenstein wrote, "One of the most dangerous of ideas for a philos-
opher is, oddly enough, that we think with our heads or in our heads"
(1967, p. 105). And he added: "No supposition seems to me more natural
than that there is no process in the brain correlated with associating or
with thinking; so that it would be impossible to read off thought-processes
from brain processes" (1967, p. 106). With his characteristic dogmatism,
Wittgenstein did not care to explain why the first idea is "dangerous" and
the second one "natural."
Second example: From an analysis ofthe (English grammar) of the verb
"to see," the analytic philosopher Ryle concludes that seeing cannot be
an experience or phenomenon, in particular a state or a process. "The
programme, therefore, of locating, inspecting and measuring the process
or state of seeing, and of correlating it with other states and processes, is a
hopeless programme . . . the product, almost, of inattention to gram-
mar" (1960, p. 104). If only Helmholtz and other scientists had paid more
attention to grammar at school, we would have been spared all of psy-
chophysics and physiological psychology! When out of touch with current
science, the philosophy of mind is a futile exercise in scholasticism rather
than a serious search for truth.
12 1. Why Philosophy of Psychology?

Psychologists retaliate by handling nonchalantly a host of concepts that


have a long and involved philosophical tradition, such as those of concept
and consciousness. Worse, they often subscribe unwittingly to philosoph-
ical hypotheses incompatible with their own work. One such hypothesis
is, precisely, psychophysical dualism. This view is so firmly entrenched in
ordinary thought and language, that even monists sometimes employ ex-
pressions that, strictly speaking, make sense only against a dualistic back-
ground.
Some expressions in common use that involve a dualistic philosophy of
mind are: 'the neurophysiological basis of mind,' 'neural correlates of
mental functions,' 'physiological equivalents of mental processes,' 'brain
systems subserving mental functions,' 'transformation of neural activity
into mental activity,' 'embodiments of mind,' and 'neural representation
(or coding) of mental processes.' Regrettably, the users of these expres-
sions do not elucidate the key concepts designated by the italicized
words, and they are rarely aware that the said expressions presuppose
psychophysical dualism. Even people who are normally fussy over quan-
titative accuracy and experimental controls frequently tolerate concep-
tual fuzziness. Consistency is rather rare.
Not the least of the virtues of psychoneural monism is that it avoids the
imprecisions and obscurities inherent in dualism. Let us take a closer look
at it.

1.3 The Identity Hypotheses


The (usually tacit) philosophy underlying psychobiological research is
materialism, according to which all real objects are material or concrete.
(See Bunge, 1981.) The materialist philosophy of mind boils down to the
so-called identity theory, which is really a hypothesis rather than a hy-
pothetico-deductive system, or theory proper. The identity hypothesis is
that all mental events are (identical to) brain events.
The identity hypothesis comes in two strengths. The strong or emer-
gentist identity hypothesis is that mental events are specific neural events
occurring in certain special subsystems of the brain, that cannot be ex-
plained solely with the resources of physics and chemistry. The weak or
leveling identity hypothesis is that mental events are just physico-chemi-
cal events occurring in the brain, on a par with the electrical signals
propagating along the neuron axon and with the binding of neurotransmit-
ters to the postsynaptic membrane-whence physics and chemistry
should suffice to account for them.
Clearly, the strong version implies the weak one: if mental events are
very special biological changes, then they are also physico-chemical
ones, though not only such. And both hypotheses are reductionistic, but
whereas the weak one is physicalist, the strong one is biologicist-and
1.3. The Identity Hypotheses I3

even so with qualifications, for it holds that the mental is a very special
kind of biological process, and moreover one strongly influenced by social
circumstances.
Many materialists object to the strong or emergentist identity hypothe-
sis because they mistrust the notion of emergence, believing that it is a
remnant of obscurantism. (Some epistemologists are to be blamed for this
resistance, for they define an emergent property of a whole as one that
cannot be explained in terms of the parts of the whole and the interactions
among them.) We shall dispell these doubts in sections 3.4 and 5.3. Suffice
it now to recall that emergence is nothing but qualitative novelty, and as
such is pervasive on all levels of reality. In particular, it accompanies
every chemical synthesis and every evolutionary novelty. Indeed, things
endowed with new (emergent) qualities result from assembly or substitu-
tion processes as well as from speciation ones. What a single cell may not
be able to do, a system of cells may achieve; and what an organism of a
given species may be unable to perform, its remote descendants may be
able to do. The leveler will not find mental processes in the single neuron
or in the invertebrate; the emergentist will seek it in large cell assemblies
of the brains of higher vertebrates. Hence the leveler, not the emergentist,
opens the door to obscurantism, which thrives on gaps in science.
The goal of psychobiologists, in particular physiological psychologists,
is to identify the neural systems that control behavior, as well as those
whose specific activity is mental (e.g., affective, perceptual, intellectual,
or volitive). Consistent psychobiologists do not look for the neural "cor-
relates," "equivalents," "subservers," "embodiments," or "represen-
tations" of mental processes, for all this is dualistic talk. Instead, they
intend to discover the neural systems that discharge behavioral or mental
functions, much as the legs do the walking and the digestive tract does the
digesting.
For example, in a psychobiological perspective, perceiving and imagin-
ing are not "represented" in the brain but are brain activities; thinking is
not "equivalent" to a brain process of some kind but is identical with it;
and there is no neural system that "subserves" planning or gets "trans-
formed" into it: planning is identical with the specific activity or function
of certain neural systems. And in all these expressions the word 'identity'
means the same as in mathematics: a = b if and only if a and bare
different names for the same item. (Further, if a = b, then b = a, and if
a = band b = c, then a = c.)
The differences between the two identity hypotheses, and between
these and their rivals, are best realized by spelling them out with the help
of the notation of elementary logic. Let M, N, and P designate the predi-
cates "mental," "neural," and "physical" respectively. Further, call C
the causal relation, x and y events, and t and u instants of time. Finally, let
'fix and 3y symbolize the quantifiers' 'for all x" and' 'for some y" respec-
tively; let:::;' stand for "if . . . then," and 1\ and V for "and" and "or"
14 1. Why Philosophy of Psychology?

respectively. With this notation, the ten best known philosophies of mind
boil down to the following formulas:
Strong (emergentist) identity: Mental events are neural.
(\:fx)(\:ft) [Mxt :::} (3y)(Nyt /\ y = x] [1.1]
Weak (leveling) identity: Mental events are physical.
(\:fx)(\:ft) [Mxt :::} (3y)(Pyt /\ y = x)] [1.2]
Parallelism: Every mental event is accompanied by a synchronous neu-
ral event.
(\:fx)(\:ft) [Mxt :::} (3y)(Nyt /\ y :;fx)] [1.3]
Epiphenomenalism: Mental events are caused by neural ones.

(\:fx)(\:ft) [Mxt :::} (3y)(3u)(u < t /\ Nyu /\ Cyx)] [1.4]

Animism: Mental events cause neural or physical ones.

(\:fx)(\:ft) [Mxt:::} (3y)(3z)(3u)(u > t /\ Nyu /\ pzu /\ (Cxy V Cxz)] [1.5]


Interactionism: Mental events cause or are caused by neural or physi-
cal ones.
(\:fx)(\:ft) [Mxt :::} (3y)(3z)(3u)(u < t)[(Nyu /\ pzu /\ (Cyx V Czx»
V (Nyt /\ pzt /\ (Cxy V Cxz»)]] [1.6]
Autonomism: The mental and the neural are unrelated.
(\:fx)(\:fy)(Mx /\ Ny :::}Axy), where A designates the empty relation. [1.7]

Idealism (spiritualism): All is mental.


(Yx)Mx [1.8]
Eliminative materialism: Nothing is mental.
(\:fx) l Mx, where l designates "not." [1.9]
Neutral monism: The mental and the physical are so many manifesta-
tions of an unknowable neutral substance.
(\:fx)(\:fy) [Mx /\ Ny :::} (3z) [Uz /\ (Azx V Azy)]], where U denotes the
neutral unknowable substance, and A is to be read "appears (or manifests
itself) as." [1.10]
The preceding formulas exhibit at a glance the virtues and shortcomings
of each view. The clearest of them all are the identity (both strong and
weak), parallelist, autonomist, idealist, and eliminative materialist views.
On the other hand, epiphenomenalism, animism, and interactionism are
vague because they do not hint at the mode of causation, or mechanism,
whereby the bodily or mental events in question would be generated. And
neutral monism (or the double aspect view) is even less clear, for involv-
1.3. The Identity Hypotheses 15

ing the vague concept of appearance or manifestation and, above all,


because it postulates that both matter and mind have an unknowable
source. Being unclear, neither epiphenomenalism, nor animism, nor inter-
actionism, nor neutral monism is in a position to stimulate psychological
research, let alone clarify philosophical issues. Let us then consider
briefly the six remaining views.
As sketched earlier, the two identity hypotheses are programmatic.
However, they are quite definite and both of them have proved to have a
tremendous heuristic power. The choice between them involves the ques-
tion of whether or not biology is reducible to physics and chemistry. We
have tackled this question elsewhere and have concluded that (a) living
things, though composed of physical and chemical items, have emergent
properties of their own, that is, properties that their components lack
(Bunge, 1979a); and consequently (b) biology, though based on physics
and chemistry, is not fully reducible to them: it has concepts, hypotheses,
and methods of its own (Bunge, 1985a). For this reason we reject the
weak or leveling hypothesis [1.2]. In other words, we deny that mental
events are characterizable in physical and chemical terms alone.
We adopt instead the strong or emergentist identity hypothesis, accord-
ing to which all mental events are biological events of a very special kind.
(We shall suggest in chapter 7 that the key to mentality is neural plasticity,
which starts at the level of the synapse.) The strong or emergentist iden-
tity hypothesis poses the challenging problem of identifying the neural
systems whose specific activities or functions are the mental processes of
the various kinds. This happens to be the program of psychobiology.
Indeed, the aim of this science is to flesh out the identity formula [1.1].
This is an extremely ambitious task, hence a tough and interesting one.
By taking this task in hand, psychologists become neuroscientists, and
the latter psychologists: the barrier between psychology and biology
crumbles. (See Lashley, 1941; Teuber, 1978.)
As for the remaining hypotheses concerning the nature of mind, paral-
lelism (formula [1.3]) is too facile a doctrine, for it does not suggest what
the relationship between the mental and the bodily might be. It only states
that, for every sequence of mental events, there is a parallel sequence of
brain events. But this is nearly universally acknowledged, and it is too
imprecise to inspire a research project. (More on parallelism in section
5.3.) The remaining alternatives to the identity hypotheses are just as
unsatisfactory as parallelism. Autonomism (Formula [1.7]) denies that
there is any relation between the mental and the physical, so it is a purely
negative doctrine, hence one incapable of fueling research into the mind-
body problem. Eliminative materialism (Formula [1.9]), by denying that
there are mental events, has exactly the same defect: it is barren. Worse,
it renders psychological research totally pointless. As for idealism (For-
mula [1.8]), it is certainly a positive doctrine, but one that happens to be
orthogonal to all of the factual (natural or social) sciences, everyone of
16 1. Why Philosophy of Psychology?

which studies nothing but material (concrete) things, be they protons,


cells, societies, or what have you. So, if we wish psychology to be a
science like other fields of scientific inquiry, as well as to harmonize and
interact with some of them, we must rule out idealism as a suitable foil.
This leaves us with the two identity hypotheses, (1.1) and (1.2), as the
only ones that satisfy the requirements of clarity, heuristic power, and
compatibility with contemporary science. However, the strong version,
namely (1.1), is preferable for being more definite and for suggesting the
union of psychology and neuroscience. In fact there is little doubt that,
although mental processes are biological processes that can be analyzed
into physical and chemical components, they have characteristics that are
not studied by physics or chemistry, or even by general biology. For
example, neither of these sciences seeks to understand typically psycho-
logical processes such as imagery or inference.
In addition to the just-mentioned virtues, the strong (emergentist) iden-
tity hypothesis is a charter member of a definite world view countenanced
by factual science. In fact, it is peculiar to materialism, the philosophy
according to which all the constituents of the real world (i.e., all exist-
ents), are material or concrete. (This does not entail denying the existence
of ideas, but denying their autonomous existence, that is, the existence of
ideas detached from brains.) The contrary of materialism is idealism, or
spiritualism, according to which all existents are immaterial.
Neither materialism nor idealism (or spiritualism) are for the faint-
hearted. The latter prefer diluted versions of the stronger stuff. These
watered-down doctrines are the (logical) contradictories of materialism
and idealism. The contradictory of materialism is immaterialism, which
holds that some real objects are immaterial (or ideal). And the contradic-
tory of idealism or spiritualism, which holds that some real objects are
material, will be called inidealism. There is no opposition between imma-
terialism and inidealism. Psychophysical dualism conjoins them.
The four theses and their mutual logical relations are shown in the
following square of oppositions. (To understand this square, recall that
the negation of "All x are F" is not "No x is an F" but "Not all x are F,"
which amounts to "Some x are not F.") See Figure 1.1.
Because workers in the factual (natural and social) sciences study ex-
clusively concrete things-though of course with the help of concepts-
they behave as materialists. (Those who affirm the existence of disem-
bodied ideas do not dare add that science condones them.) To be sure,
very few scientists realize this tacit commitment to materialism or care to
acknowledge it, and this for several reasons. First, very few people are
interested in laying bare their own presuppositions: this is a typically
foundational and philosophical task. Second, materialism has not exactly
made great strides in the course of this century, largely because it has
been mainly in the hands of amateurs. (See however Bunge, 1977a, 1979a,
and 1981 for an attempt to update materialism and free it from dogma-
1.4. Philosophical Presuppositions of Scientific Research 17

MATER IALISM .....----contrary - - - -.... IDEALISM (SPIRITUALISM)

'"
All real objects are material.

I
No real objects are material.

I
subaltern subaltern

I
IN IDEALISM ----compatible-----IMMATERIALISM
!
Some real objects are material. Some real objects are immaterial.

FIGURE 1.1. Materialism, idealism (spiritualism), and their denials.

tism.) Third, to declare oneself a materialist amounts to ringing the leper's


bell: convicted materialists are promptly isolated-or worse, joined by
undesirable company.
Still, whoever adopts either of the two identity hypotheses, if only as a
working conjecture, behaves as a materialist. Nor is this the only philo-
sophical hypothesis that underlies psychological research. We shall pres-
ently meet several other philosophical principles that guide scientific re-
search.

1.4 Philosophical Presuppositions


of Scientific Research
Philosophical psychology handles problems about behavior and mind
within the context of ordinary knowledge, in particular folk psychology,
with the help of exclusively philosophical tools. It is an armchair occupa-
tion, hence alien to experimental psychology, and a favorite with scholars
keenly interested in the so-called mystery of mind, though not sufficiently
interested to take the trouble to study the mainstream psychological liter-
ature.
It would seem that philosophical psychology is a thing of the past. This
may be so de jure, but is not so in fact. Indeed, philosophical psychology
survives not only in philosophy departments but also in the psychology
community, though marginally. In fact, the so-called humanistic psychol-
ogy, be it that of Frankl, Maslow, Rogers, Lacan, or someone else, is
nothing but a continuation of traditional philosophical psychology, which
keeps clear from both experiment and mathematical modeling. The only
difference is that humanistic psychologists, like psychoanalysts, see pa-
tients, whereas their philosophical counterparts see only books. Aside
from this, both groups share Kant's conviction that psychology cannot be
18 I. Why Philosophy of Psychology?

experimental and mathematical: that it is a branch of the humanities, not


of science.
To be sure, occasionally one finds more interesting insights on the
human mind in the writings of armchair psychologists, or even writers of
fiction, than in many a rigorous but unimaginative experiment. After all,
perceptiveness of the human condition and talent (particularly literary
talent) know no frontiers. However, neither scientific psychologists nor
science-oriented philosophers can have much patience with philosophical
psychology for, although it often approaches important and interesting
problems, it does so with disregard for the scientific approach as well as
for the findings, albeit still modest, of scientific research. Moreover, the
clinical psychologist faced with drug addiction, and the psychiatrist faced
with acute depression, cannot look upon the groundless logotherapic rites
with equanimity. They are bound to regard them as dangerous quackery.
When it comes to public health, particularly at the taxpayer's expense,
vigilance, not tolerance, is the watchword.
At first sight scientific psychology, unlike philosophical and humanistic
psychology, is totally divorced from philosophy. (In particular the behav-
iorists, though under the thumb of positivism, particularly operationism,
were proud of their alleged independence from philosophy.) However, a
methodological analysis of psychology will show otherwise. It will show
that scientific research, whether in psychology or in any other field of
knowledge, is not conducted in a philosophical vacuum but against a
complex philosophical background. Even the most modest piece of scien-
tific research presupposes (admits tacitly) a number of philosophical prin-
ciples. Three such principles will suffice for the moment: "Nothing comes
out of nothing or goes into nothingness," "We can get to know the world
even though only in part and gradually," and "Thou shalt not make up
thy data or fake thy calculations." Let us take a quick look at these and
other principles.
To check whether the preceding principles are in fact presupposed in
scientific research, imagine a project that did not admit them. Suppose an
experimenter is watching a caged monkey with the aim of finding out a
particular behavior pattern. Suddenly the monkey disappears from view.
Will the experimenter believe that this is a case of vanishing, or of tele-
transportation? More likely, he will conjecture that the monkey went into
hiding, or that he himself is the victim of a hallucination. In any event he
would not report the disappearence for fear that his colleagues will be-
lieve that he is not familiar with the old basic philosophical principle that
matter is uncreatable and indestructible.
Having solved the mystery of the missing monkey with the help of
a philosophical principle and some inquiries, our scientist resumes his
observations. He does so because he hopes to find out something new
about primate behavior. He thereby tacitly trusts the realist principle ac-
cording to which we can get to know things if we study them. What-
1.4. Philosophical Presuppositions of Scientific Research 19

ever his explicit philosophy, if any, our scientist behaves as a realist,


not as an idealist or a conventionalist. Finally, when reporting on his
observations, he will, let us hope, do so faithfully and being careful to
distinguish his raw data from the statistics constructed out of them, as
well as from his interpretation of the result: he will refer to the latter as
a hypothesis. By proceeding in this manner he will abide by one of
the methodologico-moral rules that govern the conduct of scientific
inquiry.
It might seem that all of the above goes without saying, and that learn-
ing such principles is part of anybody's scientific training. Quite so. Nev-
ertheless the point is that those principles, and many others of the kind,
are not studied in science but in philosophy. They are used and checked in
the practice of any science, but they often originate, and are sometimes
analyzed and systematized, in philosophy. (See Agassi, 1975; Bunge,
1977a, 1979a, 1983a, 1983b; Burtt, 1932.) We shall call a philosophical
background, general outlook, or worldview, the variable collection of all
such ontological, epistemological and moral principles. Table 1.3 lists a
few of them.
The philosophical principles inherent in scientific research are not al-
ways apparent-particularly to researchers with an antiphilosophical
bias. But they surface once in a while, particularly at critical junctures.
For instance, they surface when designing ambitious research projects
(such as mapping the mind on to the brain), and even when discussing
empirical findings that prove difficult to interpret; they also emerge when
building theories (hypothetico-deductive systems), and when evaluating
rival research projects as well as rival theories. In particular, referees
called upon to evaluate research projects cannot fail to brandish a whole
armamentarium of such general principles. Should some of the latter be
wrong, still a common occurrence in psychology, entire lines of research
would be thwarted. For example, a referee committed to the archaic
theologico-philosophical dogma that the mind is an independent immater-
ial entity might report unfavorably on any research project in psycho-
biology, whereas he or she may recommend some parapsychological re-
search-unless of course he or she is inconsistent.
The claim that scientific research involves the philosophical principles
listed in Table 1.3 and others besides is not descriptive but normative. It is
not being claimed that, as a matter of fact, all scientists abide by those
principles, even less that they do so consciously. Such claim would obvi-
ously be false. What is being claimed instead is that a philosophical analy-
sis of a piece of correct scientific research is bound to show that those
principles are in fact involved in such work-alas, most often unbe-
knownst to the workers themselves. The way to justify the claim was
suggested a while ago: Drop everyone of the philosophical principles one
at a time and see whether such omission is likely to cause us to make a
mistake or to disregard an interesting problem. The reader is invited to do
20 I. Why Philosophy of Psychology?

TABLE 1.3. A sample of philosophical principles involved, usually in a tacit


manner, in scientific research, both basic and applied, as well as in its planning
and evaluation.
Ontological (metaphysical) principles: On the world
01 The world exists on its own (i.e., whether or not there are inquirers).
02 The world is composed exclusively of things (concrete objects).
03 Forms are properties of things (not self-existing ideas).
04 Things group into systems: every thing is either a system or a component of
one.
05 Every thing, except the universe, interacts with other things in some respects
and is isolated from other things in other respects.
06 Every thing changes.
07 Nothing comes out of nothing and no thing reduces to nothingness.
08 Every thing abides by laws. (There are coincidences but not miracles.)
09 There are several kinds of law: causal, stochastic, and mixed; same-level (e.g.,
biological), cross-level (e.g., biosocial), etc.
010 There are several levels of organization: physical, chemical, biological, social,
and technological.
Oil All systems, except the universe, receive external inputs and are selective.
012 In every system there is spontaneous (uncaused) activity of some kind.
013 Every system has some properties (called 'emergent') that its components lack.
014 Every emergent property appears at some stage in the assembly of a system.
015 Every system belongs to some evolutionary lineage.
016 Every system, except the universe, originates by assembly.
017 The components of social systems are biological, chemical, and physical; those
of chemical systems are chemical or physical; and those of physical systems
are physical.
018 Every system, except the universe, is a subsystem of some other system.
019 The more complex a system, the more numerous the stages in the process of
its assembly.
020 The more complex a system, the more numerous its modes of breakdown.
Descriptive epistemological principles: On human knowledge of the world
EI We can get to know the world (reality), although only partially, imperfectly,
and gradually.
E2 Every cognitive act is a process in the nervous system of some animal.
E3 Humans can only know objects of two kinds: material entities (concrete things)
and conceptual objects (concepts, propositions, and theories).
E4 An animal can know a thing only if the two can be linked by signals that the
former can detect and decode.
E5 No inquiry starts from complete ignorance: We must know something before
we can formulate a problem and investigate it.
E6 Every cognitive operation is potentially subject to error, but every error is
corrigible.
E7 There are several ways of knowing: By perceiving, conceiving, and acting; and
these various modes combine in many an investigation.
E8 All human inquiry is done in society, and therefore in cooperation and
competition with others.
E9 Knowledge can be of particulars or of patterns.
EIO Every theory, when enriched with data and subsidiary hypotheses, can help
describe and predict, but only mechanismic theories can explain.
Normative epistemological principles: On the conduct of scientific inquiry
Ell Start your inquiry by choosing an open problem.
E12 Formulate your problem clearly: Unearth (or widen or restrict) its context,
presuppositions, and data.
E13 Do not mistake problems of being for problems of knowledge or conversely
(e.g .. do not try to define causality in terms of predictability, and do not
believe that facts alter when seen through alternative conceptual
frameworks).
1.5. Philosophies of Psychology 21

TABLE 1.3. Continued


Normative epistemological principles: On the conduct of scientific inquiry (cont.)
EI4 Do not let the available techniques dictate all your problems: If necessary try
new techniques or even whole new approaches.
EI5 Plan the investigation into your problem-but be ready to change your plan,
and even your problem, as often as necessary.
EI6 Whenever possible handle your problems scientifically (i.e., armed with
scientific knowledge and scientific methods, and aiming at a scientific or
technological goal).
EI7 Do not tolerate obscurity or fuzziness except at the beginning: Try and exactify
every key concept and proposition.
EI8 Do not commit yourself before checking: First get to know, then believe-and
doubt.
EI9 Revise periodically your most cherished beliefs: You are bound to find fault
with some of them-and with luck you may even start a conceptual
revolution.
E20 Check (for clarity, cogency, and effectiveness) all your methodological rules.
Moral precepts: On the right scientific conduct
MI Be truthful.
M2 Do not skirt research problems for fear of the powers that be.
M3 Regard all data, theories, and methods as fallible, and regard only research as
sacred.
M4 Correct everything corrigible, particularly your own errors.
M5 Do not disregard superstition and pseudoscience: expose and fight them.
M6 Do not hoard knowledge: Share it.
M7 Give credit where credit is due.
M8 Disregard arguments from authority and ad hominem.
M9 Cherish intellectual freedom and be prepared to fight for it. .
MIO Be modest: know your limitations, but do not be humble; do not bow before
authority or tradition.
MIl Do not use prestige gained in advancing knowledge to underwrite unjust
causes.
MI2 Be of service to coworkers, students, and the scientific community at large.
M13 Shun ideology in basic science but declare it in technology.
MI4 Refuse to use knowledge for purposes of destruction or oppression.
MI5 Do not boast of special (in particular paranormal) cognitive powers.
MI6 Try to justify all your claims.
MI7 Keep your independence of judgment and, if necessary, swim against the
stream.
MI8 Tolerate serious research on problems or with methods you dislike.
MI9 Be intolerant with regard to organized obscurantism.
M20 Keep constant moral watch on your own actions and even on your own moral
principles.

Note: For detailed discussions see Bunge (l967a. 1977a. 1979a. 1983a. 1983b. 1985a. 1985b).

this checking for himself and to look elsewhere (Bunge, 1977a, 1979a,
1983a, 1983b, 1985a, 1985b) for guidance.

1.5 Philosophies of Psychology


A philosophy of psychology is, of course, a philosophical study of psy-
chology. Because, as we saw in the beginning, philosophy is composed
basically of logic, semantics, epistemology, ontology, and ethics, a com-
prehensive philosophy of psychology should contain a logic, a semantics,
an epistemology, an ontology, and an ethics of psychology.
22 1. Why Philosophy of Psychology?

Now, at any given moment of its history, psychology involves a num-


ber of philosophical principles: Recall section 1.4. Besides, some of the
findings of psychological research, pure or applied, are bound to support
or undermine prior philosophical hypotheses or suggest new ones (e .g.,
that there is, or that there cannot be, mind without body, or knowledge
without learning). Consequently, a comprehensive philosophy of psychol-
ogy cannot avoid facing some of the logical, semantical, epistemological,
ontological, and moral principles involved in, or suggested by, psycholog-
ical research and practice. (For the denial that the philosophy of psychol-
ogy must tackle ontological issues, such as that of the nature of mind, see
Margolis, 1984.) To be sure, any particular study in the philosophy of
psychology will focus on one of these aspects, but the discipline as a
whole must handle them all.
Like other branches of philosophy, the philosophy of psychology is
anything but a unitary and firmly established field of knowledge. In fact
there are nearly as many philosophies as philosophers of psychology, and
all of them are rather sketchy and sectorial. None of them covers all of the
aspects of the discipline (logical, semantical, etc.), and most of them are
far removed from current psychological research, or they make hardly
any use of analytical tools, or they are not components of comprehensive
philosophical systems. In short, the existing philosophies of psychology
are not characterized by their unity of approach and theme. Nor is this the
exclusive fault of philosophers. Thus Robinson (1985), a psychologist,
defends old-fashioned psychophysical dualism with the help of obsolete
philosophical tools, and joins Eccles (Eccles and Robinson, 1985) in the
crusade against materialism.
There are several modes of philosophizing about psychology or, in-
deed, about anything. A philosophical discourse may be original or scho-
lastic, constructive or critical, exact or inexact, systematic or fragmen-
tary, and science-oriented or nonscientific (or even anti scientific ). Or it
may combine two or more such modes.
Thus, before introducing a new idea it is advisable to sketch the rele-
vant background knowledge, that is, to do a spot of expository and critical
work. (A piece of philosophy qualifies as scholastic only if it contains no
new ideas, particularly if it is apologetic.) An original discourse may
consist either in conceptual analysis or in theory construction, but it is not
necessarily devoid of critical remarks. Criticism may motivate construc-
tion, and in any event every new idea should be examined critically. (A
piece of philosophy is purely critical or destructive if it proposes no
alternati ves.)
When expounding a philosophical system one is justified in inteIjecting
clarifying comments or examples that, strictly speaking, do not belong to
the system. Also, the discourse may be exact in spots and inexact or
informal in others. (However, there is a difference between inexactness,
e.g., that of ordinary language, and obscurity, whether deliberate or a
1.5. Philosophies of Psychology 23

manifestation of a neurological disorder.) Finally, philosophical discourse


may be science-oriented in places and nonscientific (e.g., ideological) in
others. However, if the aim is to foster the advancemente of knowledge, it
should not be antiscientific.
Purity of mode, then, is not of paramount importance in philosophy.
What does matter is that the discourse be intelligible (perhaps with some
learning effort), interesting or relevant (i.e., dealing with important prob-
lems), true at least in part, and that it possess some heuristic power-that
is, suggest new hypotheses, experiments, or methods, or relate previ-
ously separate ideas. However, we submit that the mode of philosophiz-
ing most likely to lead to clarity, relevance, truth, depth, and heuristic
power, is that which combines criticism with exactness, and systemicity
with closeness to current research and practice. Let us try to justify this
claim.
The need for criticism is quite obvious, not only because criticism is a
component of all rational inquiry, but also because both philosophy and
psychology are still underdeveloped, despite their old age, partly because
they continue to harbor plenty of dogmas-for example, psychoneural
dualism and the belief that ordinary language suffices for philosophy,
psychology, or both. However, we should not exaggerate the value of
criticism at the expense of that of invention (of hypotheses) and discovery
(offacts). The function of criticism is to regulate research, not to replace
it. Criticism is to the advancement of knowledge as the thermostat is to
the furnace. Without a thermostat the furnace may go wild, but without
the furnace the thermostat will remain idle.
As for exactness, or the compliance with logical standards and the use
of mathematical tools, it may be dispensed with only in the early stages of
research. Thereafter it must be gradually increased for at least three rea-
sons. First, because we want to minimize misunderstandings and the
corresponding hermeneutic disputes. (If Freud had been an exact thinker
he might not have engendered more than 100 psychoanalytic schools.)
Second, exactness favors testability. (Compare, for example, "Y is a
power function of X" with "Y depends on X.") Third, depth, always
desirable, calls for exactness, because hypothetical constructs remain
suspect unless they are quantitated and related in an exact manner to
observable quantities.
The virtue of systemicity is equally obvious in dealing with complex
issues, such as those of psychological and philosophical research. In both
cases we are bound to resort to numerous hypotheses and definitions,
methods and data, that at first sight are mutually unrelated. Focusing on
only a few such components is likely to result in a distorted presentation
of the whole. A good example of the shortcomings of sectorial thinking is
faculty psychology, which ignored the interactions among the cognitive,
affective, and motor components of mental phenomena. (Much of con-
temporary cognitive psychology is guilty of the same quartering.) Another
24 1. Why Philosophy of Psychology?

example is the second Wittgenstein, whose books are all collections of


disjointed aphorisms and examples.
Finally, only a scientific attitude and proximity to current research can
produce a philosophical discourse attractive enough to psychologists to
have a chance of suggesting fruitful new scientific ideas or discouraging
wrongheaded research projects. (One of the reasons psychologists do not
think much of philosophers is that most of the latter deal only in folk
psychology.) However, proximity to the science of the day must be tem-
pered with a dose of skepticism, for otherwise the philosopher runs the
risk of being dragged by a fashionable current, which need not be the most
promising one. (Remember the days when famous psychologists identi-
fied mind with consciousness, or wrote it off altogether, or identified it
with a set of computer programs?) In short, even while openly consorting
with scientists, philosophers should keep their independence of judgment,
recalling that the best science can be wrong (e.g., for want of suitable
philosophical guidance). Philosophy should be neither the slave nor the
master of science, but its coworker. This cooperation should not hush up
mutual criticism and it should contribute to the development of both
parties.

1.6 Summing Up
Once upon a time psychology was a member of the philosophical family.
Toward the middle of the nineteenth century it suffered from the illusion
that it had become fully emancipated from philosophy. Now that its strug-
gle for independence is over, and that it is in the process of becoming a
mature science, it can afford to admit that, like any other science, psy-
chology is not disjoint from philosophy.
An examination of any ambitious psychological research project, or of
any psychological breakthrough, suggests that our science is shot through
and through with ontological, epistemological, and moral principles. In
particular, much research into mental phenomena presupposes some phi-
losophy of mind or other. And even more research of the same kind has
been avoided under the pressure of wrong philosophies of mind and of
science. Besides, some of the findings of psychological research ought to
be assimilated by philosophy for, after all, the problems of the nature of
mind, and of how best to study it, are of interest to both philosophy and
psychology.
The point, then, is not to renounce philosophy but to keep it under the
control of science, and to help it develop into a discipline capable of
actively advancing scientific knowledge.
CHAPTER 2

What Psychology Is About

Many psychologists and psychology watchers complain about the lack of


consensus concerning the very object or referent of their discipline. How-
ever, psychology is not unique in this regard. Thus, some biologists are
not sure whether it behooves them to study the chemistry ofbiomolecules
such as DNA. Many chemists count thermodynamics as their own but on
the other hand are apt to surrender to physicists when the latter claim that
the whole of chemistry is nothing but a chapter of physics. Even in phys-
ics, the oldest and most powerful of the factual sciences, there are some
spirited disputes about what it really is about. Thus, whereas most physi-
cists hold that physics happens to be the study of physical things, others
(the followers of the Copenhagen interpretation of quantum theory) deny
that there are autonomous things, and claim that physics studies what
appears to observers-that is, appearances. And a few go as far as to hold
that the quantum theory cannot be understood unless it includes the hu-
man mind-which, if true, would render physics indissolubly linked to
psychology. However, none of these controversies prevents the discus-
sants from going about their business: The uncertainties concerning the
proper objects of study certainly affect the way science is taught and
philosophized about, but it hardly touches mainstream research.
Things are quite different in psychology. Every view on the subject
matter or referent of psychology is likely to affect profoundly the kind of
problems to be tackled and the manner of investigating them. Thus, if
psychology is defined as the study of consciousness, everything else will
be left aside and introspection may be favored over all the other methods.
If, on the other hand, psychology is defined as the study of overt behav-
ior, then only observable movements will be studied, and everything else
will be ignored. One reason for the greater importance of the question of
subject matter in psychology than in any other science is that psychology
is still in the process of transition from the proto scientific stage to the
scientific one. Consequently the old tradition, formed in the womb of
classical philosophy, is more strong than sound, whereas the new tradi-
tion is still weak.
26 2. What Psychology Is About

Given these conditions it is not surprising that, whereas some students


of psychology still seek refuge in some school or other, there are those
who adopt a nihilistic or cynical view, saying that psychology is the
science that has at least two explanations for every phenomenon, and no
phenomena at all for most of its theories. But the present authors, along
with the great majority of psychologists, are neither dogmatists nor nihi-
lists: They believe that psychology has a fairly well defined class of refer-
ents-though by no means a narrow one-that can and should be studied
scientifically. They also believe that this class can be identified by analyz-
ing some of the key concepts and hypotheses of contemporary psychol-
ogy, such as the concept of learning and the hypothesis that learning is the
same as the strengthening of interneuronal connections.

2.1 Definitions of Psychology


As we all know, etymologically 'psychology' signifies the study of the
psyche, soul, spirit, or mind. This is how classical psychologists and
theologians conceived of the subject matter of psychology. So do most
philosophical psychologists, psychoanalysts, and humanistic psycholo-
gists as well. For example, Descartes's classic on the subject is titled Les
passions de l' arne, and Freud made frequent references to the Seele,
which in the standard English edition (1953-1965) got transformed into
mind-the lay sister of soul.
Some psychologists, in line with tradition and, more particularly, fol-
lowing James, Dewey, and even Piaget, refer to 'psychology' as "the
study of the functions of the mind." Taken literally this expression pre-
supposes that the mind is an entity or thing, for it is attributed functions,
that is, activities. Hence the expression, taken literally, presupposes
some version of psychophysical dualism (for which see section 1.2). In a
psychobiological context, where it is assumed that the mind is a collection
of brain functions (activities), the expression 'functions of the mind' is
equivalent to 'functions of a collection of brain functions.' Because the
latter expression makes no sense, the psychobiologist cannot accept the
definition of 'psychology' as "the study of the functions of the mind."
The radical behaviorists have reasons of their own for rejecting the
preceding definition. The first is that it leaves out precisely that which
interests them most, namely overt behavior. The second is that they do
not believe in the existence of the mind, or at least in the possibility of
studying it scientifically. Therefore they define 'psychology' as "the sci-
entific study of behavior." However, this does not solve our problem.
First, because, if the term 'behavior' is construed narrowly, namely as
observable bodily movement, then psychologists are prevented from
studying affect, cognition, and other important categories of phenom-
ena-which is a tacit invitation to pseudoscientists to fill the gap. And if
2.2. Referents of Psychology 27

'behavior' is construed in a broad manner, so as to include affect, cogni-


tion, and the rest, then the term 'behaviorism' loses its sting. (Parallel:
The democrat who tolerates friendly authoritarian regimes.) Second, psy-
chologists should be interested not only in overt behavior but also in its
motivation as well as in the neural mechanisms of both. Third, to a behav-
iorist psychology is only one of the "behavioral sciences" alongside an-
thropology, sociology, economics, politology, history, and linguistics.
But then what makes psychology so special? And what distinguishes it
from the study of the behavior of bacteria and amoebas, or even from that
of bodies in general (i.e., mechanics)? Any of the preceding reasons suf-
fices to rule out the classical behaviorist definition of "psychology"-
which of course is no reflection on the undeniable historical importance of
behaviorism.
We seem to have reached an impasse. We reject the definition of 'psy-
chology' as "the study of the functions of mind" while admitting that it
must study the mind (or the mental functions ofthe brain). And, although
we reject the behaviorist definition of 'psychology,' we do not deny that
our study must handle behavior-yet not of everything but only of ani-
mals. (The occasional reports on the psychic life of plants have proved to
be groundless: See e.g., Kmetz, 1978.) Moreover not every animal spe-
cies falls under the jurisdiction of psychology. For instance, psycholo-
gists, as such, are not interested in the behavior of organisms devoid of
nervous systems. As a matter of fact the vast majority of animal species
are studied by zoologists, not psychologists.
Psychologists are only interested in animals capable at least of perceiv-
ing and learning and, in particular, able to learn to modify their behavior
in an adaptive manner. And such learning is likely to take a nervous
system far more complex than, say, a neural net such as that of a sponge.
We shall stipulate then, in line with current mainstream views, that 'psy-
chology' is the scientific study of the behavior (and mind if any) of ani-
mals endowed with a nervous system that enables them to at least per-
ceive and learn. This definition excludes from psychology the
nonscientific study of behavior and mind, as well as the scientific study of
animals incapable of perceiving and learning. The latter are the concern of
zoologists and ethologists.

2.2 Referents of Psychology


If the preceding definition of "psychology" is accepted, the referents or
subjects of study of our science turn out to be all and only the animals
which, in normal circumstances, are capable of perceiving and learning.
The mention of normal circumstances is intended to account for learning
impairments due to genetic defects, injury, disease, sensory deprivation,
etc. However, it also behooves psychologists to study such impairments.
28 2. What Psychology Is About

Our definition leaves out of the purview of psychology all of the animals
that are normally incapable of learning. These are the animals lacking a
nervous system or possessing one that is genetically predetermined or
"prewired," as a consequence of which their behavior is rigid. Such
animals compose the vast majority of animal phyla. Their behavior is
usually studied by zoologists and ethologists.
There is some evidence that a few invertebrates, notably bees and
octopi, can learn. However, the attribution of a learning ability (as of any
other ability) depends critically on the definition of "learning." If mere
change of behavior in altered environmental circumstances, for example,
habituation ("adaptation"), counts as learning, then even worms and sea
slugs (Aplysia) qualify as subjects of psychological research-otherwise
they do not. (We shall return to this matter in sections 7.2 and 9.1.)
Because there is no consensus on the definition of "learning, " there can
be none on whether or not invertebrates can learn. Pending a resolution of
this conceptual conflict over the definition of "learning," we may leave
the study of invertebrate behavior to the zoologists and ethologists till
new notice.
As for vertebrates, there is no doubt that all of the higher vertebrates,
namely mammals and birds, can learn, and therefore qualify as subjects of
psychological research. However, other classes, particularly fish, am-
phibians, and reptiles, should be looked into in more detail before disqual-
ifying them. Yet, as in the case of invertebrates, we may leave them for
now to the ethologists and zoologists. In short, the referents of psychol-
ogy are mammals and birds. But, as cautioned previously, this refers only
to current mainstream psychology.
Our definition of "psychology" excludes animal societies from the
referents of our science. The reason is that only individuals of certain
species are capable of learning, and even fewer of being in mental states.
Societies do not learn, feel, perceive, or think. To attribute to societies
psychological properties or abilities is just as mistaken as attributing them
biological properties or functions.
This is not to say that psychologists must ignore society. On the con-
trary, social psychologists are supposed to investigate social behavior,
the social conditioning of learning and mental functions, and the (indirect)
social consequences of ideation (e.g., planning). (See chapter 10.) Still,
the focus of psychology, be it individual or social, is the individual in its
natural or social environment, not society. Societies are studied by social
scientists, not by psychologists. Likewise, geologists study rocks, not the
atmosphere, even though they are also interested in the action of atmo-
spheric processes, such as rain and wind, on rocks. Their central referent
is the lithosphere, not the atmosphere. In like manner, societies are,
alongside natural habitats, the peripheral referents of psychology: the
central referents of the latter are individual animals capable of learning.
2.2. Referents of Psychology 29

The preceding discussion is not as Byzantine as it might appear. In fact,


it disposes at one stroke of two branches of classical psychology: the
psychology of peoples (Volkerpsychologie) and the psychology of the
masses (Massenpsychologie). It is of course legitimate to study the psy-
chology of individuals belonging to different societies, e.g., preliterate
and literate, or agrarian and industrial-in short, to engage in cross-cul-
tural psychology-in order to find out the impact of social progress on
individual behavior and ideation. Likewise it is legitimate to study the
effects of peer group and mob pressure on the individual, as well as the
effects of leadership on social behavior. But to claim that social wholes,
such as peoples and masses, have a mind of their own, is sheer holistic
fantasy, because only individuals have nervous systems, and only some
nervous systems can be in mental states.
Another class of objects excluded from the reference class of psychol-
ogy is that of artifacts, even those endowed with artificial intelligence.
The reason for this exclusion is that they are not animals. This is the same
reason that ornithologists, as such, do not study aircraft: namely because
they are biologists, not engineers. Surely computers (endowed with pro-
grams and controlled by humans) mimick or replace certain animal func-
tions, but they work in an entirely different manner from that of an ani-
mal. Surely they do certain jobs that formerly only people could do, but
they do not perform them as people but as their surrogates or proxies. In
short, psychologists as such do not study machines, except to be able to
handle them, or to find out what animals are not. On the other hand,
artificial intelligence experts cannot help but study psychology, particu-
larly cognitive psychology, because what they want to mimic or replace is
natural intelligence, a prerogative of some animals. We shall come back to
this fashionable topic in section 5.4.
We have decided then that psychologists are not primarily social scien-
tists although they may have to take the social matrix into consideration.
We have also decided that they are not engineers either, although they
may use their knowledge of human psychology to help design computer
programs or robots. Psychologists study animals, in particular humans,
and on this count they are zoologists. But they are highly specialized
zoologists. Not that they confine their interests to a single class of ani-
mals, but that they specialize in learned behavior and mentation. Because
some behavior and some mentation are strongly conditioned by social
circumstances, psychology has some overlap with social science. This
intersection is composed of social psychology, social ethology, and bioso-
ciology. In short, our definition in section 2.1 entails that psychology is
primarily a biological science and secondarily a social one. Equivalently:
The central referents of psychology are animals capable of perceiving and
learning, whereas its peripheral referents are animal societies. We shall
return to this subject in chapter 13.
30 2. What Psychology Is About

2.3 The Fragmentation of Psychology


and How to Remedy It

Twentieth-century psychology looks like a huge mural on a great many


subjects painted in all colors either by an industrious schizophrenic or by
an army of workers belonging to hundreds of disjoint crafts and rival
schools. No pattern is to be seen. There is scientific psychology on the
one hand and a large variety of nonscientific psychologies on the other.
Within scientific psychology there are behaviorist and mentalist students,
as well as biological, social, and even engineering orientations. Besides,
there are the basic-applied, the animal-human, and the normal-abnor-
mal divisions. And, whereas some psychologists specialize in emotion,
other focus on cognition, language, mental retardation, or what have you.
(See e.g., Boring, 1950; Brunswik, 1955; Marx & Hillix, 1973.)
The various schools and "systems" of psychology are so many ap-
proaches to psychological problems, and they are usually supported by
different philosophies of mind. (Recall chapter 1.) The division into
schools is not as strident in basic research as it once was, it is no longer
associated with great names, and it is being deliberately downplayed in
academic teaching except in some underdeveloped countries. But it is still
there, as shown by the fact that one and the same problem is often ap-
proached in very different manners-that is, there is a multiplicity of
paradigms. Thus, learning is currently being studied by, among others,
ethologists, behaviorist psychologists, and physiological psychologists.
These different approaches are mutually incompatible more often than
mutually complementary. In any event, these various groups often ignore
one another, they employ different methods, and they arrive at mutually
inconsistent conclusions. Regrettable, but true.
To be sure, controversy in science is normal and healthy-as long as it
results in the demise of falsity and the birth of truth. However, factional-
ism in psychology has passed the limit, because some factions have be-
come barren and even pseudoscientific. What would we think of physics if
some of its leaders were to teach that bodies are set in motion by ghosts?
What of chemistry if well-known chemists were to proclaim that their
discipline is so special that it can and must be cultivated without regard to
physics? What of biology if some of its most prominent workers were to
claim that the study of animal toys is more rewarding than that of live
animals? And yet in present day psychology there are plenty and notori-
ous parallels to these ludicrous imaginary cases.
In addition to the fragmentation into schools there is the split into
different fields or problem systems. For example, there are experts in
vision and others in hearing, in memory, or in bilingualism, in personality
or in small groups, and so on and so forth. Some fragmentation into
2.4. Unification in Action 31

subfields is unavoidable given the huge amount of problems, as well as the


individual differences among researchers. It parallels what is occurring
nowadays in all scientific disciplines, and it is the price paid for studying
problems in depth-or so we are told. In fact, the consequence is some-
times shallowness rather than depth. For example, knowing as we do that
the limbic system-the "seat" of emotion and much else besides-has
multiple reciprocal connections with every region of the neocortex, it is
illusory to try to attain a full understanding of perception, learning, mem-
ory, will, and other processes in complete detachment from emotion.
Given that the current fragmentation of psychology into warring
schools and disjoint subfields hinders the advancement of our science,
what can be done to overcome it? Because the fragmentation into rival
schools derives from rival philosophies, it can only be overcome by
adopting a single underlying philosophy-preferably the one closest to
the "scientific spirit." And the fragmentation into disjoint subfields can
be overcome by recalling at all times that a single protagonist plays all of
the behavioral and mental roles, namely the nervous system.
Actually the two measures we have just proposed to forge the unity of
psychology are not mutually independent. In fact, adopting a philosophy
containing the psychophysical identity hypothesis (section 1.3) entails
that every item of psychological interest be viewed as being controlled by
the nervous system (the case of behavior) or as a particular function of
that system (the case of mental processes).
Note that we are not proposing that every item of psychological interest
be tackled exclusively by physiological psychologists: This would defeat
our purpose of promoting the unity of psychology. It would also deprive
physiological psychology of most of its problems, for the ultimate aim of
physiological psychology is to lay bare the mechanism of every psycho-
logical fact, regardless of the field where it was first studied. All we are
proposing is that, no matter what level of analysis or description be cho-
sen, it be kept in mind (or rather in brain) that the process happens to be
neural or under the control of some neural system, whence it should also
be tackled by physiological psychologists. In other words, we are propos-
ing that psychology be just as firmly based upon neuroscience as chemis-
try is based on physics, and biology on chemistry. Let us see what conse-
quences this approach can have for the actual conduct of psychological
research.

2.4 Unification in Action


Suppose a team of researchers convinced of the benefits of a unified
approach to psychological problems decides to study voluntary move-
ment in macaques. They are likely to begin by videotaping the overt
behavior of a monkey in the process of reaching for a peanut or some
32 2. What Psychology Is About

other stimulus likely to arouse the will. They will vary the context and the
setup (e.g., placing the peanut in a box while the monkey is looking and
allowing him to reach for it only after a few minutes). All this and more
has got to be described: It is the raw material to be processed, the data to
be explained.
If the researchers are curious they will want to know what the particu-
lar neuromuscular mechanisms of voluntary movement are, and how they
are altered by drugs or by surgery. This will involve the use of more or
less invasive techniques, starting with myography. Nor will this suffice.
They will also wish to identify the neural systems in the frontal lobes that
control those neuromoscular mechanisms. And this will require implant-
ing electrodes in the regions of the brain suspected of effecting such
control.
Having found out the "seat" of will, the psychologists will try to trace
the drives, perceptions, imageries, memories, and expectations that trig-
ger or interfere with the animal's decision to reach for the peanut, or to
refrain from doing so. All this will require further training, electrode
recording, and testing. Finally, our curious psychologists will want to
know how the presence of conspecifics of the same or different sex, age,
and social status alters the process and, in particular, which additional
neural systems are activated (stimulated or inhibited) in such circum-
stances.
In sum, the curious psychologist (or rather the cross-disciplinary team
of curious psychologists) will investigate voluntary movement, or any
other psychological process, on various levels and freely trespass the
frontiers between the various subfields of our science. He or she will
attempt to integrate these subfields because the borders between them are
quite artificial: They have not been erected by the subject of study but by
psychological tradition. Only such integration on the basis of neurophysi-
ology can yield a reasonably complete (pro tempore) picture (description)
and, in addition, a plausible explanation in terms of mechanisms. (We
shall come back to integration in section 13.2.)
Let us insist on the artificiality of the division of psychology into sub-
fields. So far all attempts to classify adequately the various kinds of
behavior and mentation have failed. To be sure, one can distinguish per-
ception from imagery, locomotion from problem solving, and so on. But
there is no clear criterion lfundamentum divisionis) allowing one to parti-
tion the gamut of psychological phenomena in a neat and orderly fashion.
At most there are more or less vague laundry lists. One reason for this
failure may be that all psychological phenomena are mixtures, that
is, they have a number of aspects or components, mainly affective,
behavioral, sensory, and cognitive. In some cases one of these
components prevails whereas the others are far less prominent. But
in other cases, such as those of sensorimotor activities, two or more
components are equally important. For instance, if I am expecting an
2.4. Unification in Action 33

important telephone call, when I hear the phone ring I may rush to it
charged with emotion while at the same time imagining the face of my
expected interlocutor and anticipating the message that I am about to
receive. Such a process is at the same time affective, sensorimotor,
and cognitive.
We may then distinguish behavioral, affective, sensory, and cognitive
aspects, but we may not be able to separate them in all cases. This being
so, there is no partition of psychological phenomena into behavioral,
affective, sensory, and cognitive. The same holds for other proposed
partitions. (A partition of a set is like that of a pie: it is neat, that is, any
two subsets are mutually disjoint.) Hence a person who claims to be
studying, say, a cognitive phenomenon such as inferring, must be under-
stood as saying that he or she focuses on the cognitive aspect of that
phenomenon, feigning that the remaining aspects do not exist. This is just
a useful fiction-useful, that is, until proved otherwise. (See section 9.4
for the nefarious detachment of cognitive psychology from the remaining
departments of the science of behavior and mind.)
The ultimate reason for the impossibility of drawing clear demarcation
lines between the various psychological phenomena, hence between the
corresponding subfields of our science, is this: All psychological phenom-
ena are processes occurring in, or controlled by, the nervous system. And
the latter, though unitary, is composed of a large number of subsystems
intimately coupled to one another as well as to other body systems, such
as the muscular, the endocrine, the immune, and the cardiovascular ones.
Likewise, it would be impossible to understand in detail the motion of a
car if only the intention of its driver, or the mechanical aspect, or the
thermodynamic or the electrical one, were taken into account. An ade-
quate understanding of the car-driver-road system calls for attending to
all those aspects.
Finally, note that the proposed unification of psychology on the basis of
neuroscience is not the only logically possible one. There is an alternative
and far more popular proposal, namely to disregard the nervous system
altogether and construe every psychological phenomenon as an instance
of information processing (or computing). We shall examine this proposal
in section 5.4. Suffice it here to say that we reject it emphatically for
several reasons, among them the following. First, because, by ignoring
the nervous system, it cuts the link between psychology and neuro-
science, and thus fails to explain psychological phenomena. All it does
is to redescribe them in information-language terms. Second, because
the attempted reduction of the wonderful, qualitative variety of be-
havioral and mental phenomena to computation impoverishes psychology.
Psychology is not about general-purpose information processors but
about animals endowed with a nervous system that is the outcome of a
long evolutionary process, that goes through a developmental process,
that learns and unlearns, and sometimes-in the case of the higher
34 2. What Psychology Is About

vertebrates-astonishes us by performing new actions or creating


new ideas.

2.5 Aims of Psychology


In discussing the aims of psychology we must start by distinguishing basic
research, on the one hand, from applied research and professional prac-
tice on the other. The aims of basic psychology are or ought to be the
same as those of any other basic science, namely to describe, explain, and
predict (or retrodict) the facts it studies.
Scientific psychology is of course more demanding than other kinds of
psychology. In scientific psychology the descriptions being sought are
objective, the explanations valid, and the predictions (or retrodictions)
correct. A description is said to be objective if it is an approximately true
statement of matters of fact rather than a work of fiction. An explanation
is valid if it is a valid argument involving only well-confirmed hypotheses
and well-attested data. (A valid argument is a deduction of a set of propo-
sitions from another set of propositions in accordance with the rules of
deductive logic. Nondeductive arguments, such as analogical and induc-
tive, are occasionally fruitful but they lack formal validity: There are no
universal rules of analogical or inductive inference.) Finally, a prediction
(or a retrodiction) will be correct if, in addition to being a valid argument
from confirmed hypotheses and firm data, it is borne out by observation or
experiment.
We shall not dwell on the above-mentioned typical operations of basic
science, because they are treated in works on epistemology, methodol-
ogy, and philosophy of science (e.g., Bunge, 1983a, 1983b). But we hasten
to point out that there is no consensus among psychologists on these
matters. In particular, there are entire schools of thought that deny the
possibility of objectivity, others that deny the existence or even the possi-
bility of well-confirmed general hypotheses (i.e., laws), and still others
that deny the need for explanation in psychology. Let us examine these
opinions.
The anti science movement and its philosophers claim that there can be
no objectivity in science, whence there is actually no difference between
science and nonscience. Some claim that this is so because scientists are
just as biased as laymen, and others because the scientists themselves
create the facts rather than finding them in the external world. An analysis
of scientific research proves both claims to be wrong. The first, because
even if an individual is biased, if one belongs to a scientific community
one's procedures and findings are subjected to the critical examination of
one's peers. For example, a researcher's results are not normally ac-
cepted unless they were obtained following certain norms and unless they
have been replicated by independent workers. This is not to say that bias,
2.5. Aims of Psychology 35

error, and even fraud are absent from science, but that they can be discov-
ered and corrected. Scientific knowledge is not perfect but perfectible.
As for the subjectivist thesis that facts are not out there but are the
creatures of scientists, it has a grain of truth. In fact, the experimenter can
cause events that do not normally occur in nature. For example, he or she
can train a pigeon to discriminate among certain drawings, or an ape to
use rudiments of the American Sign Language. But such facts occur in the
real world: They are not figments of the scientist's imagination. It is also
true that artifacts can occur in experimental work, but they can eventually
be discovered and corrected for (e.g., by altering the experimental de-
sign). Scientists may alter the world in small ways but they do not create
it: They are born into it and they intend to account for it. On the other
hand, the primary concern of technologists is with controlling and even
remodeling reality. Yet both technologists and scientists admit more or
less tacitly the real existence of the external world. If they did not, they
would not undertake to study it, and they would not test their hypotheses
and their designs. Whether they know it or not, they are scientific realists
(Bunge, 1983b, 1985c).
The opponents of scientific psychology, and even a few practitioners of
it, complain that mainstream psychology neglects the individual. Some of
them go as far as to claim that individuals are so unique that there can be
no psychological laws. The first charge is justified insofar as there is no
science of the (unique) individual (Aristotle). But it is unjustified in that,
except for the universe as a whole, every individual, whether atom or
society, is similar to some other individuals in a number of respects while
at the same time possessing idiosyncrasies. (Were it not for such similari-
ties all general concepts would be idle.) Such similarities make it possible
to categorize (i.e., to group) different things into species, and to hypothe-
size laws satisfied by every member of a given species. True, such gener-
alizations may be suggested by the examination of a handful of cases, as
was usually the case with Piaget. But the test of any generalization calls
for the examination of a representative sample of the species (population).
Like any other science, psychology studies individuals as well as spe-
cies. It studies individuals in the hope of finding general patterns, and it
uses the latter to account for idiosyncrasies. There is a constant concep-
tual flux between individuals and species and conversely. But, whereas in
basic research the flux from individuals to species is strongest, in applied
science and technology the inverse current is strongest. Indeed, whereas
in the former the study of the individual is a means in the search for
pattern, in applications the pattern is a means to understand and treat the
individual. For example, the finding (law) that all cases of mental disorder
X are caused by deficiency of chemical Y would allow the psychiatrist to
treat individual cases of X with doses of Y.
As for the controversy over explanation, the situation in psychology is
as follows. Traditional positivists, all the way from Comte to the Vienna
36 2. What Psychology Is About

Circle, decried explanation and exalted description. The early behavior-


ists adopted this stance and claimed that psychologists should only be
concerned with describing overt behavior and finding constant stimulus-
response relations. This deliberate restriction to description was not just a
case of vassalage to a philosophical school: It was an act of revolt against
the pseudoscientific explanations then prevalent in psychology, that is,
explanations in teleological and mentalistic terms. This situation has al-
tered radically with the consolidation of physiological psychology, which
is sometimes in a position to explain what makes the subject tick.
Whereas formerly psychologists attempted to account for behavior in
terms of final causes and ideas, neuropsychologists try to explain both
behavior and mentation in neurophysiological terms. In sum, scientific
explanation is now becoming possible in psychology. And it is necessary:
Without it there is neither understanding nor the motivation to conduct
research in order to understand. (More in section 13.3.)
The same laws and data that are used to explain psychological facts can
also be used to predict or retrodict them. Thus, knowing that a diet poor
in protein results in irreversible deficits in the development of the neocor-
tex, we can predict that all the children of a certain social group that
subsists almost exclusively on corn will grow up into adults suffering from
intellectual deficits. And conversely, having found a group of adults with
these characteristics, we may hazard the hypothesis that they were the
victims of a protein-poor diet during childhood-a case of retrodiction.
But, of course, both the prediction and the retrodiction must be checked.
The theoretical importance of prediction and retrodiction is obvious:
They allow us to test our hypotheses for truth. Their practical importance
is no less obvious: They allow us to cause certain events or to prevent
them from happening.
Prediction, explanation, and even description are best performed with
the help of theories (i.e., hypothetico-deductive systems). (A hy-
pothetico-deductive system is a set of propositions everyone of which is
either an initial hypothesis, or axiom, or a deductive consequence, imme-
diate or remote, of one or more axioms.) Any set of propositions sharing
referents (things described) and predicates (denoting properties or rela-
tions) can be organized as a theory. Such logical organization has a num-
ber of advantages: It fits the systemicity found in reality, it brings ideas
together, it increases the number of direct or indirect empirical supporters
for every idea in the system, it facilitates inference, and, last but not least,
it minimizes the number of ideas to be remembered.
Naturally, not all theories are of equal worth. The most powerful theo-
ries are those that combine maximal strength and generality with maximal
exactness, depth, and truth. Theory A is stronger than theory B if B
follows from A. Theory A is more general than theory B if the reference
class of A includes that of B. (That is, if A is about all the things B is about
and more.) A is more exact than B if A uses more mathematics than B
2.5. Aims of Psychology 37

does. A is deeper than B if A explains whatever B explains but not con-


versely. And A is truer than B if A makes more correct predictions or
retrodictions than B. (For details see Bunge, 1983b.)
For example, a theory accounting for habituation in all organisms is
more general than one that accounts for the same process in invertebrates
only. A mathematical model of a cell assembly is more exact than the
corresponding model couched in ordinary language. A neuropsychologi-
cal theory of problem solving (which is still to be built) will be deeper than
a phenomenological (or black box) theory about the same process. And a
theory about locomotion that involves not only external stimuli but also
internal states is likely to have greater predictive power.
However, we need theories of all possible powers. In particular, we
need special theories (i.e., conceptual models) alongside general ones,
because there are both special and general patterns. And in the beginning
we must settle for modest theories, to eventually be replaced with more
general and exact, as well as deeper and truer, theories. Only such potent
theories can help improve our current descriptions, explanations, and
predictions of behavioral and mental events. If in doubt, look at the
increasingly important role of theorizing in such disciplines as nuclear
physics, astronomy, chemistry, genetics, physiology, and sociology.
Regrettably, current psychology is extremely poor in theories proper
(hypothetico-deductive systems). It is even poorer in scientific theories
(theories both experimentally testable and in harmony with our back-
ground knowledge). There are three causes of this state of affairs. One is
the romantic prejudice against theories: Recall Goethe's dictum about
theories being grey, whereas the tree of life is green. This prejudice is
particularly intense with reference to theories in psychology and in social
science, often regarded as being unavoidably soft. A second cause is the
antitheoretical bias of the positivist philosop~y inherent in radical behav-
iorism: Remember Skinner's 1950 paper" Are theories of learning neces-
sary?" A third cause is probably an effect of these antitheoretical preju-
dices, namely the inadequate mathematical training that psychology
students ordinarily receive.
In any event, the theoretical poverty of current psychology is not only
an indicator of the underdevelopment of our discipline; it is also a serious
obstacle to experimental research. Indeed, if our ideas are few, scattered,
and neither very precise nor too deep, our experimental designs are un-
likely to be imaginative, and the interpretation of our experimental find-
ings is unlikely to be unambiguous. Whoever admits the accuracy of our
diagnosis must agree that theoretical research, and particularly the mathe-
matical modeling of neural systems, should be pushed far more vigorously
than heretofore-though, of course, in close contact with experimental
work.
So much, for the time being, for theory and its value for description,
explanation, and description. Sometimes a further condition is demanded
38 2. What Psychology Is About

from research in human psychology, namely that it be ecologically valid


(Neisser, 1976, Introduction). A hypothesis or a theory in human psychol-
ogy is said to have ecological validity only if it tackles interesting ("rele-
vant") problems about real people in ordinary situations. If on the other
hand it overlooks most phenomena and refers only to artificial (labora-
tory) situations, a theory is said to lack such validity. Ecological validity
has nothing to do with truth. A theory can be ecologically valid but false
(like psychoanalysis), or significantly true but ecologically invalid (like a
theory of maze learning); or, finally, it can be neither true nor ecologically
valid (like most information-processing models).
The ecological invalidity of much of scientific psychology is realized
upon reflecting that love, one of the most intense and admirable of all
human emotions, sung by poets and described and analyzed by novelists
and playwrights, has hardly attracted the attention of scientific psycholo-
gists. In a memorable paper titled "The Nature of Love," the distin-
guished primatologist Harlow (1958) berated psychologists for neglecting
the study of love, and proceeded to describe his own, now classical,
experiments on wire-and-cloth monkey mother surrogates. Since then
dozens of psychologists have studied the mother-infant bond of affection.
But adolescent and adult love, as well as friendship, have continued to be
overlooked by the vast majority of psychologists. Even those who are not
psychoanalytically oriented-and they are the vast majority-have pre-
ferred to focus on the sexual component of love, ignoring the other com-
ponents.
As for the aims of applied research and practice, they are efficacy
and efficiency. Efficacy is the ability to consistently (not occasionally)
produce the desired effects, in this case to help patients with some
psychological handicap. Efficiency is low-cost efficacy (i.e., low
input-output ratio). Behavior therapy is an example of a proce-
dure both effective and efficient in treating certain disorders, for
example, drug addiction and phobias. On the other hand, psycho-
analysis is a prime example of a technique both ineffective (it does
not work) and cost inefficient. We shall come back to these points in
chapter 12.
We shall wind up this section by listing the various conditions that an
item of our discipline-be it datum or prediction, hypothesis or theory,
method or experimental design-may satisfy or fail to satisfy. We stipu-
late that item X is

(1) ontologically valid if X concerns real animals (instead of disembodied


spirits or fictional characters) or means for studying them;
(2) methodologically valid if X is either
(a) a scrutable and justifiable procedure or design, or
(b) a datum or a prediction obtained or checked with the help of a
procedure or design complying with condition (a), or
2.6. Summing Up 39

(c) a hypothesis or theory testable with the help of items of types (a)
or (b);
(3) alethically valid if X is a datum, prediction, hypothesis, or theory,
found to be sufficiently true by using methodologically valid items
(i.e., satisfying the previous condition);
(4) ecologically valid if X is of interest to more than the particular investi-
gator(s) who handle X, and X is relevant to real life situations;
(5) practically valid if it is both effective and efficient.
For example, whereas the neuropsychology of learning is valid on all
five counts, parapsychology has only some methodological validity, psy-
choanalysis and humanistic psychology only some ecological validity,
and philosophical psychology has none whatsoever. (See section 5.5 for a
methodological criticism of pop psychology.)

2.6 Summing Up
To sum up, psychology is supposed to study the behavior and mental life,
if any, of animals capable of learning. Such study has three cognitive
aims-description, explanation, and prediction-and one practical goal,
namely the treatment of behavioral and mental disorders. Our next ques-
tion is how best to attain these goals.
II

Approach and Method


CHAPTER 3

Approaches to Behavior and Mind

Until the turn of the century mental functions were widely regarded as
mysterious. There was even a theologico-philosophical industry of mental
mystery mongering. Its motto was ignoramus et ignorabimus: We ignore
and will always ignore what mind is. The echoes of this obscurantist
attitude can still be heard today.
Scientists dislike mysteries and miracles. They are after problems and
laws. So, as psychology developed into a science, the would-be mystery
of mind was gradually transformed into a system of more or less clearly
stated problems. Moreover, some of these problems have been solved at
least to a first approximation (e.g., those of motor control, of classical and
operant conditioning, and ofthe treatment of phobias and drug addiction).
Actually, even a few much bigger problems may be regarded as having
been solved once and for all (e.g., those of the very nature of mind, of the
existence of racial memories, and of the possibility of paranormal phe-
nomena such as precognition).
However, it would be foolish to deny that most psychological problems
remain unsolved, or have been solved only to a first approximation. For
example, we have only sketchy ideas about the mechanisms of vision,
thinking, and consciousness. In other sciences a few chapters have al-
ready been written in all essentials. For instance, nobody expects any
breakthroughs in trigonometry or in wave optics. Not so in psychology:
Here nearly everything remains to be done. It still is, and will remain so
for a long time, the land whose moving frontiers are in sight, so that even
beginners and serious amateurs can make contributions to it.
What has been accomplished in psychology over the past hundred
years or so is to be credited to some of their practitioners having adopted
the right approach and rejected a sterilizing tradition. This is the tradition
of philosophical psychology, divorced from biology, alien to both experi-
ment and mathematics, skeptical of the possibility of finding general pat-
terns of behavior and mentation, and sure that, mind being immaterial, it
could not possibly be studied scientifically. Let us examine the approach
44 3. Approaches to Behavior and Mind

that has made the progress of psychology possible, and contrast it to some
of its alternatives.

3.1 Approach
In this section we shall analyze the concept of an approach. (For details
see Bunge, 1983a). To put it roughly and metaphorically, an approach is a
way of looking at things (e.g., people) or ideas (e.g., conjectures), and
therefore also of handling problems concerning them. This loose charac-
terization will shortly be replaced with a formal definition.
We shall distinguish eight broad types of approach to the study and
handling of things or ideas: the vulgar, empirical, doctrinaire, and human-
istic; and the mathematical, scientific, applied, and technological.
The vulgar approach rests on ordinary knowledge, tackles both basic
and practical problems, is mainly interested in practical results, and em-
ploys exclusively procedures from daily life, in particular routines and
trial and error. The empirical approach rests on both ordinary knowledge
and the knowledge gained in the practice of some art or craft, tackles only
practical problems, is interested exclusively in practical results, and em-
ploys procedures from both daily life and artisanal practice. The doctri-
naire approach rests on some rigid doctrinal body (e.g., an ideology or a
pseudoscience), tackles both basic and practical problems, is primarily
interested in practical problems (including the defense of the doctrine),
and resorts to authority, criticism, and argument. The humanistic ap-
proach is based on the body of knowledge concerning human culture,
handles cognitive problems concerning intellectual and artistic prob-
lems, aims at understanding its referents, and uses primarily heuristic
methods.
The mathematical approach is characterized by a formal basis (logic
and mathematics), formal problems, the aim of finding patterns and con-
structing theories, and conceptual methods, in particular the method of
formal proof. The approach of basic science rests on a fund of mathemati-
cal and experimental knowledge, as well as on a scientific worldview, it
deals with basic problems, aims ultimately at understanding and forecast-
ing facts with the help of laws and data, and employs scientific methods,
in particular the scientific method. The approach of applied science
shares the basis and methods of basic science, but is restricted to special
basic problems, and aims at supplying part of the cognitive basis of tech-
nology. Finally, the technological approach is like that of applied science,
but its basis includes also the fund of technological knowledge, its prob-
lems are practical, and its aim is the control of natural systems as well as
the design of artificial ones.
In general, an approach (Sl) may be defined as a body B of background
knowledge together with a collection P of problems (problematics), a set
3.2. Atomism. Holism. and Systemism 45

A of aims, and a collection M of methods (methodics) , that is, as the


quadruple
S'l = (B, P, A, M) [3.1]
Every component of this quadruple is to be taken at a given time, and so it
is a time-dependent collection rather than a fixed set.
An approach may be construed as an ordered set because we actually
proceed in an ordered manner when handling things or ideas even while
groping in the dark. In fact, B comes first because problems do not emerge
in a vacuum but in a body of antecedent knowledge, namely as gaps in it,
and every attempt to solve them uses some items in B. P comes immedi-
ately thereafter, for it is what we want to address on the basis of B. The
way to handle a problem depends on our aim, which may be cognitive or
practical (i.e., one of knowing or of doing). Hence A must be placed third.
And once the problem and the aim have been chosen, we pick or invent a
method to handle the former: Hence M comes last.
Moral: A well-written paper will start by presenting some background,
and will go on to pose the problem(s) to be addressed, to declare the
goal(s) aimed at, and the methodes) employed in the research.

3.2 Atomism, Holism, and Systemism


Every approach is based on a body of antecedent or background knowl-
edge, and every such body includes a general philosophical framework, or
worldview, which is often tacit rather than explicit. For example, as we
saw in section 1.4, the scientific approach presupposes a number of prin-
ciples concerning the nature of things (the ontology of science), the ways
of getting to know something about them (the epistemology of science),
and the right conduct of the investigator (the morality of science).
In the present section we shall restrict our examination of the back-
ground knowledge to the general philosophical framework or world-
view-or, rather, the numerous families of such worldviews. Even a
cursory examination of it reveals three rival approaches, everyone of
which can be coupled with some of the eight approaches discussed in the
previous section. These three approaches, which are the object of consid-
erable attention (but little analysis) on the part of psychologists and
neuroscientists, are atomism, holism, and systemism. Let us proceed to
sketch and evaluate them. (For details see Bunge, 1977a, 1977b, 1977c,
1979a, 1983a, 1985b.)
The atomistic (or individualistic, or analytic) approach rests on an at-
omistic ontology, according to which the world is an aggregate of units of
a few kinds, and a reductionistic epistemology, according to which the
knowledge of the composition of a whole is both necessary and sufficient
to get to know the whole. The goals of atomism are the same as those of
46 3. Approaches to Behavior and Mind

science, and the atomistic methodics boils down to analysis into compo-
nents (or the top-down method). Examples: associationist psychology
and faculty psychology.
The holistic (or synthetic) approach rests on a holistic or organismic
ontology, according to which the world is an organic whole that may be
decomposed into large partial wholes that are not further decomposable.
This ontology comes together with an intuitionistic epistemology accord-
ing to which such ultimate wholes must be accepted and grasped as such
(on their own level) rather than analyzed and tampered with. The goal of
holism is to emphasize and conserve wholeness and emergence (the quali-
tative novelties that accompany the formation of some wholes); and its
method (or rather nonmethodical procedure) is often intuition rather than
reason or experiment. Examples: the view that the brain is an undifferen-
tiated (unstructured) whole, and Gestalt psychology.
Finally, the systemic (or system-theoretic) approach rests on a systemic
ontology according to which the world is a system composed of subsys-
tems belonging to different levels, and an epistemology that recommends
combining reason with experience in order to understand the formation
and dismantling of systems in terms of their components, the interactions
among these, and the environment. The aims of systemism, like those of
science and technology, are description, understanding, prediction, and
control. Its methodics includes analysis as well as synthesis (in both cases
conceptual and empirical), generalizing and systematizing (in particular
mathematical modeling), and empirical testing (of hypotheses, theories,
and methods). Examples: the view that the brain is a system composed of
mutually interacting subsystems, and the hypothesis that every mental
process has affective and cognitive components, as well as sensorimotor,
visceral, and endocrine concomitants.
Because every approach is in part characterized by its own problem-
atics (section 3.1), each of the just mentioned approaches can handle only
certain problems. Thus the atomistic approach can tackle only questions
regarding individual behavior: Because it does not admit the existence of
wholes with emergent properties, it sees no point in looking for patterns
of global behavior (i.e., for laws of systems as wholes or units, such as the
so-called molar laws of perception). Likewise, the problem system of
holism is limited. It is hardly interested in finding out, say, the details of
the long chain of events triggered by a visual stimulus and ending up in the
experiencing of an image. By contrast, systemism retains the positive
aspects of atomism and holism; it studies both the whole and its parts, and
it admits the occurrence of emergence, or qualitative novelty, as well as
the possibility of explaining it. Therefore, of the three approaches sys-
temism is the one that best fits in with the scientific approach. Actually
the latter includes the former.
We shall now introduce a completely general model of concrete sys-
tems of any kind, living or nonliving. Call Cf6(s,t) the composition, ~(s,t)
3.2. Atomism, Holism, and Systemism 47

the environment, and :f(s,t) the structure of a system s at a time t. We


define the composition of a thing as the collection of its parts; the environ-
ment as the collection of things, other than the given system, that act
upon the latter or that can be influenced by the system; and the structure
as the collection of relations among the components of the system (inter-
nal structure) as well as among these and items in the environment (exter-
nal structure). See Figure 3.1. (For details see Bunge, 1979a.) Our quali-
tative model of an arbitrary material system s at time t is the ordered triple
m(s,t) = <~(s,t), ~(s,t), :f(s,t». [3.2]
This model is generally suitable provided the context indicates clearly
which levels of analysis have been chosen-for instance, if it has been
explicitly indicated that the level of composition is to be the cellular one,
whereas that ofthe environment is to be that of macrophysical entities. In
biological matters one may be interested in a number of levels of analysis:
atomic, molecular, cellular, and several others. Therefore one should
indicate explicitly the level of analysis-call it A -at which the composi-
tion is to be taken. We do this by writing: ~A(S,t), where ~A(S,t) is the
intersection of ~(s,t) with A. What holds for the composition holds also
for the environment: We indicate the level of its analysis B by writing:
~B(S,t). These two levels of analysis (which may coincide) uniquely deter-
mine that of the structure, which we call C; that is, we call :fds,t) the
structure of s at t on level C. This yields a somewhat more precise model
of a system, namely
[3.3]
We can now appreciate somewhat better the advantages of the systemic
approach. For one thing, only the systemic approach does justice to all

~~~~~~~
~EJ@EJ@88
(a) (b) (c) (d) (e) (f) (9)

FIGURE 3.1. Differences among systems resulting from changes in their composi-
tion, environment, or structure. (a) Different composition, same environment and
structure; (b) different structure, same environment and structure; (c) different
environment, same composition and structure; (d) different composition and
structure, same environment; (e) different composition and environment, same
structure; (j) different structure and environment, same composition; (g) different
composition, environment, and structure.
48 3. Approaches to Behavior and Mind

three aspects or coordinates: composition, environment, and structure.


On the other hand, atomism focuses on composition while neglecting
environment and structure; environmentalism pays exclusive attention to
the second coordinate of the above triple; and structuralism ignores both
composition and environment.
The systemic approach has two important additional virtues. First, it
prepares us to study systems on all the necessary levels-unlike atomism,
which drags us to the lowest level, and holism, which pulls us up to the
highest. Second, as a consequence of this attention to the multiplicity of
levels, as well as to the context or environment of the object of study, the
systemist student will tend to cross some of the artificial frontiers between
fields of inquiry: He or she will tend to adopt an interdisciplinary ap-
proach. And this approach happens to be indispensable to the understand-
ing of highly complex things immersed in extremely complex and rapidly
varying environments, such as ice crystals, clouds, neurons, whole ani-
mals, or societies.
Let us next apply the various distinctions we have made in this section
to psychology.

3.3 Nonscientific Approaches to Psychology


Psychology is probably the only field of learning where all eleven ap-
proaches discussed in the previous sections are still being tried. In partic-
ular, in no other field of knowledge dealing with matters of fact does one
still find vulgar, empirical, or doctrinaire approaches. Can anyone imag-
ine what a vulgar physics, a purely empirical chemistry, or a doctrinaire
biology might look like?
The vulgar approach to behavior, affect, and cognition yields what has
been called folk psychology. It has recently become fashionable to ridi-
cule folk psychology, placing it on the same footing as folk physics. This
analogy is inadequate, because we have no inside information about at-
oms, clouds, or stars. On the other hand, we do possess a large body of
inside information-admittedly often wrong-about ourselves. We also
have plenty of opportunities to check our ordinary knowledge conjectures
about ourselves and others. Whereas physicists have no use whatsoever
for folk physics, folk psychology is often a point of departure of psycho-
logical research. In fact, a major goal of the latter is to enlarge, deepen,
correct, and explain folk psychology, including some of the insights found
in great works of art. For this reason folk psychology is unlikely to ever
disappear. Instead, it is likely to be gradually corrected and enriched with
some of the findings of scientific psychology.
The empirical approach-not to be mistaken for the experimental
one-has resulted in classical clinical psychology and psychiatry. These
are essentially collections of case studies, empirical generalizations, and
3.3. Nonscientific Approaches to Psychology 49

untested conjectures unrelated to either neuroscience or experimental


psychology. Classical clinical psychology and psychiatry are slowly being
replaced with the corresponding scientific disciplines, which are finding
the biological and social roots of behavioral and mental disorders as well
as ways to cope with them.
The doctrinaire approach has ensued in the two hundred or so existing
schools of verbal psychotherapy, none of which takes serious notice of
scientific psychology. The new schools that keep popping up every year
are not the result of research but of speculation and controversy.
Finally, the humanistic approach to psychology yields, of course, hu-
manistic psychology. (See e.g., Welch, Tate, & Richards, 1978.) Actually
the humanistic approach to psychology is a combination of the vulgar,
empirical, and doctrinaire approaches. Whatever truth and efficacy may
be found in humanistic psychology derives from folk and empirical psy-
chology. The current popularity of this kind of psychology seems to be
due to its large measure of ecological validity (for which see section 2.5),
as well as to the fact that it makes only modest intellectual demands on its
consumers. But the current vogue of humanistic psychology is likely to
fade out. for scientific psychology, whether basic or applied, is finally
tackling problems of interest to everyone, and it is gradually yielding
some important results. The humanistic approach will eventually be re-
stricted to its proper field of application, namely the humanities (e.g.,
philosophy, art history, and literary criticism).
Most philosophers of mind have restricted their attention to folk psy-
chology, doctrinaire psychology (in particular psychoanalysis), and hu-
manistic psychology. In particular, they often claim (e.g., Davidson,
1970) that, unlike the physical, the mental is lawless. This claim is in line
with some idealistic philosophies, particularly those derived from Kant.
The thesis is not arbitrary: It is suggested by the large individual differ-
ences found in the behavior and mentation of human beings. No two
individuals behave in exactly the same manner, and even one and the
same individual behaves differently at different times. But all sciences are
in the same predicament, because no two things are strictly identical, and
no one individual remains identical to itself. Yet all sciences attempt to
find similarities and constant patterns beneath individual differences, as
well as laws of change. Why should psychology be different in this re-
gard?
The only legitimate controversy with regard to psychological laws con-
cerns their scope. Whereas psychophysicists, behaviorists, and physio-
logical psychologists have looked for completely general (cross-specific)
laws, workers in other disciplines, particularly ethology, comparative
psychology, and cognitive psychology, have emphasized specific differ-
ences. However, far from denying the existence oflaws, all these workers
have searched for species-specific patterns. (See, e.g., Bitterman, 1984.)
No wonder, since the search for law is part of the very definition of the
50 3. Approaches to Behavior and Mind

concept of scientific research. (See, e.g., Bunge, 1967a, 1983b.) The mo-
ment a domain offacts is declared lawless, it is placed beyond the reach of
science and within the province of mythology or religion.
Not all philosophers have remained content with adopting some scraps
of vulgar, empirical, doctrinaire, or humanistic psychology. Some ofthem
have tried their hand at armchair psychologizing, in recent years particu-
larly of the information-processing kind. The outcome has a been a num-
ber of philosophical psychologies, most of them tacitly or explicitly dual-
istic. (See e.g., Dennett, 1978; Haugeland, 1981; Hofstadter & Dennett,
1981; Pears, 1975; Popper & Eccles, 1977; Searle, 1983.) There is of
course nothing wrong with speCUlating about matters of fact: Without
speculation we would have neither science nor technology. But specula-
tion, to be fruitful, must be disciplined or sound, not wild; it must be
testable (at least in principle) and it must be compatible with the bulk of
our scientific background knowledge (Bunge, 1983c). Thus it is unprofit-
able to speculate on worlds other than our own and with which we could
never exchange any signals. It is equally unprofitable to speculate
on minds that could be neither studied nor modified by experimental
means.
Regrettably, most philosophical speculations and many philosophical
arguments about behavior and mind are wild. Thus one well-known objec-
tion to the psychoneural identity hypothesis is that there can be no such
identity, for genuine (or "necessary") identities hold "in all possible
worlds"-whatever these may be. And it might be the case that, in a
world other than ours, minds have nothing to do with brains, just as heat
might not be the same as random atomic or molecular motion (Kripke,
1971). Those psychophysical dualists who have realized that postulating
interactions between mind and brain would involve a violation of the
principle of conservation of energy have persuaded themselves that this
should be no cause for worry because, after all, that principle might tum
out to be false (e.g., Popper & Eccles, 1977). Still others (e.g., Broad,
1949), anxious to defend parapsychology and realizing that telepathy,
precognition, telekinesis, and the like would violate all the ontological
principles underlying modem science (section 1.4), have said frankly: So
much the worse for those principles.
The authors of such wild speculations expect to be taken seriously.
Occasionally, when they are not, they get angry and abusive (e.g., Met-
zinger, 1985). But why should one take them seriously at a time when
behavior and mind are being studied scientifically with some success?
Philosophical psychology need not be taken more seriously than philo-
sophical bacteriology or philosophical metallurgy if there were such.
Speculative metaphysics lost the right to exist the moment modem sci-
ence became established three centuries ago. Now is the time for scien-
tific metaphysics (Bunge, 1971, 1977a).
3.4. Toward a Scientific Psychology 51

3.4 Toward a Scientific Psychology


At least three different approaches have been tried to turn psychology
from a branch of philosophy into a science: the mentalist, the behaviorist,
and the biological ones. They have been tried precisely in the indicated
order, all three are alive in the current psychological literature, and every
one of them has made some contribution to our understanding of behavior
and mind. It will therefore be interesting to see what their virtues and
shortcomings are. (For details see chapters 5 to 9. For an alternative
evaluation see Marx & Hillix, 1973.)
The mentalist approach was an outgrowth of the idealistic philosophies
of mind. It studies mental phenomena in themselves, without any refer-
ence to biology. Although it usually comes together with some idealistic
philosophy of mind, it can be pursued without making any commitment as
to the nature of mind. In fact, one may study, say, emotion, perception,
or inference without inquiring whether these occur in an immaterial mind
or in a material brain. A commitment one way or other becomes unavoid-
able only if one attempts to explain the findings of a piece of research-in
particular if one wishes to explain them as being identical to neural pro-
cesses.
The behaviorist approach emerged in part as a reaction against mental-
ism and, in particular, against the abuse of introspection and speculation
on the part of mentalist psychologists. Behaviorists reject the definition of
psychology as the study of the mental, and like to be known as scientific
students of animal and human behavior-although, quaintly enough,
they pay no attention to zoology. Although positivism was a powerful
motivation for behaviorism (section I. I), it is possible to study ani-
mal behavior in the objective way behaviorists have taught us without
subscribing to positivistic strictures, in particular while claiming that
behavior, far from explaining anything, is one of the things that we
want explained. Moreover, because behaviorists are not particularly
interested in the mind-body problem, they need not take sides in this
controversy.
Finally, the biological approach to psychology is a sort of extension of
both the mentalist and the behaviorist strategies, since it tackles both
mental and behavioral processes. It is also deeper than its predecessors,
because it attempts to explain the findings of the students of mind and
behavior. (Actually the biological approach is also capable of explaining
why some of those findings were spurious-e.g., why we cannot experi-
ence arbitrarily long uninterrupted "streams of consciousness," or why
there can be no strict stimulus-response regularities that skip over inter-
nal states of the organism.)
Let us now take a closer look at the three approaches, recalling that
every approach (.71) is a body B of background knowledge together with a
52 3. Approaches to Behavior and Mind

collection P of problems (problematics), a set A of aims, and a collection


M of methods (methodics), or s!l = (B, P, A, M) in short (section 3.1).

Mentalism

Background
Mentalism, which is very much in evidence in contemporary cognitive
psychology, can be scientific or nonscientific. If the former, its B will not
contain the views that minds are disembodies entities, or that they can
interact with bodies, for these are at best untestable dogmas without any
empirical support. Only nonscientific mentalism contains these wild spec-
ulations; and, not being scientific, it makes no use ofthe scientific method
to test them. Scientific mentalists need not take sides in the mind-body
controversy: one may just declare that he or she is only interested in
discovering, describing, and explaining mental events. The scientific men-
talist's B is made up of a number of the ontological, epistemological, and
moral principles underlying any mature science (section 1.3). An impor-
tant exception is Principle 02, "The world is composed exclusively of
things (concrete objects)." In fact, if the mentalist psychologist is com-
mitted to some version of psychophysical dualism, he or she will explic-
itly reject that principle. But whether he or she will be able to do scientific
research on disembodied minds, is another matter. Actually, if the men-
talist psychologist proceeds scientifically he or she will handle exclusively
concrete things, even though he or she may adopt dualism while theoriz-
ing. The Wiirzburg school was a good example of such double talk (see
Marx & Hillix, 1973). (Double standards are not infrequent in science:
Consistency is hard to come by.) As for the morality of mentalist investi-
gators, in principle it can be as strict as any. However, being methodolog-
ically naive, they are more exposed to self-deception and dogmatism than
others. Finally, mentalism has no specific scientific background to speak
of; in particular, it makes no use of mathematics (except perhaps some
rudimentary statistics) and biology. In this regard, that is, in its isolation
from other fields of knowledge, it resembles pseudoscience and ideology.
In short, the background of mentalism is extremely small; it is literally
nearly groundless.

Problematics
Mentalists make a point of showing that, unlike others, they tackle most if
not all of the traditional problematics of psychology. This is certainly the
main virtue of mentalism. Regrettably, this problematics is rather narrow:
It disregards most of the problems about behavior and about the neural
"correlates" of mental processes.
3.4. Toward a Scientific Psychology 53

Aims
The declared aim of mentalism is to describe and understand human
mentality. The most progressive among the mentalists, in particular the
members of the Gestalt school, have searched for patterns (laws). In fact,
they have found a handful of generalities of a qualitative kind. But, be-
cause they disregard both overt behavior and the nervous system, they
cannot find any precise quantitative regularities.

Methodics
Nonscientific mentalism is typically speculative, metaphorical, dogmatic,
nonexperimental, and nonmathematical. On the other hand, scientific
mentalism, as practiced by Wundt and his successors, is sober and partly
experimental. However, its main method is introspection, which is hardly
an instance of the experimental method. And practically the only piece of
apparatus needed to do experimental mentalistic psychology is a stop-
watch. Such modest equipment looks medieval by comparison with the
sophisticated electronic paraphernalia found in a behavioral or neuro-
physiological laboratory .

Behaviorism

Background
The general outlook or world view of behaviorism is thoroughly naturalis-
tic; in particular, it denies the existence of an immaterial mind. But the
outlook is narrow because it discounts nonbehavioral phenomena such as
emotion, imagination, and consciousness. The epistemology of behavior-
ism is realistic, for it endeavors to account for an aspect of reality, the
existence of which it admits the moment it demands that research be
objective. However, this realism is rather primitive because it shuns hy-
pothetical constructs such as desire and intention. Behaviorism can make
do with a primitive epistemology because it avoids deep hypotheses and
theories (i.e., constructs that do not represent immediately observable
features). Moreover, it deals only with molar events, such as the organ-
ism's response to the nth presentation of a stimulus of a certain kind.
Finally, it makes no reference to mental states or, ifit does, it attempts to
handle them exclusively by means of intervening variables-intervening,
that is, between stimulus and response. On the other hand, the morality of
behavioristic basic research is strict. We should be grateful to the found-
ers of behaviorism for having introduced such a rigorous code of conduct
into psychology, where illusion and deception (usually unwitting) were
not uncommon. Finally, the specific background of behaviorism is ra-
ther narrow; although it makes some use of mathematics (in particular
54 3. Approaches to Behavior and Mind

of the probability calculus), it makes none of biology. Its main link with
science is by way of output, not input: It consists in the contributions
it has made to the scientific description of molar animal and human
behavior.

Problema tics
The problematics of behaviorism is complementary to that of mentalism;
it is exclusively interested in behavior, and uninterested in mind. There
would be nothing wrong with this restriction if behavior could be satisfac-
torily understood without making hypotheses about neural mechanisms,
motivation, expectation, and all that. The curious scientist cannot re-
main satisfied with finding out that a rat is willing to undergo a mild
electric shock in exchange for the possibility of exploring its surround-
ings. But asking whys is forbidden in the land of behaviorism: Only de-
scriptions of phenomena and their relations are permitted there. At most,
halfway explanations are allowed, such as "Animal A produced response
R on cue S because A was conditioned to associate S with R." The
curious scientist wishes to find out the mechanism of such conditioning;
this will lead him or her to inquire into the corresponding neural mecha-
nism-which is out of bounds for the behaviorist.

Aims
The aims of behaviorism are to describe, predict, and control animal and
human behavior. The description is supposed to include general (cross-
specific) laws of behavior, in particular oflearning. Explanation is written
off for either or both of two reasons. First, because it is not regarded as
possible or even desirable. Second, because any correct explanation of
overt behavior is to be sought in the neuromuscular apparatus and, in the
case of higher vertebrates, also in the brain mechanisms that control that
apparatus. Trying to understand behavior solely on the strength of obser-
vations of behavior is like trying to understand television by looking at the
screen and abstaining from theorizing about electromagnetic waves and
electrons.

Methodics
The methodics of behaviorism is as scientific and as narrow as its aims. It
employs observation, measurement, experiment, and statistics, all of
which is fine. But it is narrow because it rejects theorizing or restricts it to
the construction of models involving only stimuli, responses, and inter-
vening variables. These models are superficial because they are of the
black box type, such as classical thermodynamics. Moreover, they are in
line with the Aristotelian conception of change, according to which the
cause (stimulus) suffices to produce, and thus explain, the effect (re-
3.4. Toward a Scientific Psychology 55

sponse) regardless of the inner structure and state of the system. A nar-
row background suggests a narrow problematics and a narrow range of
aims, which in turn calls for a narrow methodics. The harvest is meager
compared to the enormous effort invested in the design and execution of
experiments over the better part of our century.

Psychobiology

Background
Psychobiology adopts the full scientific world view summarized in section
1.3, plus the identity hypothesis that mental processes are brain pro-
cesses. Because it makes use of mathematics-albeit on a rather mo-
dest scale for the time being-we may count mathematics too in its
background. And, of course, it is based on biology, in particular neuro-
science, which in turn presupposes chemistry and physics. In short,
psychobiology has the broadest basis of all three approaches. For this
reason it is the one most firmly entrenched in the system of scientific
knowledge.

Problematics
The problematics of psychobiology is the full range of behavioral and
mental facts. It excludes no problem of this kind that can be handled
scientifically, not even the problems of the nature of consciousness and
free will. Hence the problematics of psychobiology includes that of be-
haviorism and a large part of that of mentalism. It drops some of the
problems of mentalistic psychology, such as where the mind goes in deep
sleep or in coma, or at death. But on the other hand, it adds the entire
problematics of developmental and evolutionary biology, which mental-
ism had ignored. In particular, it asks at which stage in individual devel-
opment consciousness starts, and poses problems about the origins of
cerebrallateralization, language, and rationality.

Aims
The aims of psychobiology are those of behaviorism and more. Indeed,
in addition to describing behavior, psychobiologists attempt to explain
it in neurobiological terms. However, this task has only begun. The
ultimate aim of psychobiology must be the building of theories, both
broad (general) and narrow (specific) ones capable of explaining and
predicting behavioral and mental facts in biological terms. We need not
just descriptive theories, but theories capable of explaining behavior and
subjective experience as processes involving the nervous system and
possibly other systems as well-preferably theories formulated mathe-
matically.
56 3. Approaches to Behavior and Mind

Methodics
Unlike mentalism, which is short on measurement, and unlike behavior-
ism, which is short on theory, psychobiology makes full use of the scien-
tific method: problem-hypothesis (or, even better, theory)-logical pro-
cessing-empirical operation-inference-evaluation of hypothesis (or
theory)-new problem-and so on. Unlike mentalism, which records (in-
trospectively or by questioning) mental phenomena, and behaviorism,
which pays no attention to them, psychobiology is in a position to monitor
and alter mental processes in a direct manner because it identifies them
with processes in the brain. It is then in a position to make full use of the
experimental method, which can nowadays be implemented by a number
of sophisticated and accurate techniques. In fact, psychobiology employs
not only the methodics of neuroscience but also the methods invented by
the psychophysicists and behaviorists, and even introspection. (The latter
is useless for testing purposes but it is indispensable as a source of infor-
mation and even insight.)
In conclusion, we note a progressive movement toward the constitution
of psychology as a full-fledged science: from mentalism to behaviorism to
psychobiology. This movement has been accompanied by a shift in the
underlying philosophies, namely from idealism to positivism to natural-
ism. The mentalist and the behaviorist approaches proved to be deficient
by having excessively narrow backgrounds, particularly in borrowing
very little from other fields of knowledge. The strongest point of mental-
ism is its problematics, that of behaviorism its methodics. The aim of
mentalism is grandiose but unattainable with introspection and armchair
speculation alone. On the other hand, the aim of behaviorism is far too
modest-whence its meager achievements relative to its research effort.
The biological approach to behavior and mind shares the virtues but not
the shortcomings of its predecessors. It has the broadest background, it
handles the vastest problematics, it has the most ambitious aims, and it
makes full use of the scientific method. For these reasons the inception of
psychobiology must be counted among the great scientific revolutions of
this century. And, like every other scientific revolution, far from erasing
its predecessors it has incorporated whatever valid elements they contrib-
uted. (See Bunge 1983b, chap. 13, sect. 3, for the concepts of epistemic
evolution and revolution.)

3.5 Scientific Psychology


The scientific approach to behavior and mentation has resulted in the
gradual development of psychology as a science on the same footing as
other scientific disciplines, only less advanced than some. Being a sci-
3.5. Scientific Psychology 57

ence, scientific psychology shares a number offeatures with its sisters. In


particular, it shares the scientific world view or attitude, the scientific
method, and the general aims of science: description, explanation, and
prediction. Besides, like every genuine science, scientific psychology in-
teracts vigorously with its neighbors. But, of course, psychology has its
peculiarities. For example, it is the science of behavior and mind, and the
one that looks for ways of objectifying (i.e., finding objective indicators)
of mental events. But it does so in close cooperation with other sciences,
particularly biology and, to a lesser extent, social science. Psychology is
thus a very special discipline but, far from being autonomous, it is a
member of the tightly knit system of the sciences.
Scientific psychology may be characterized as the following 10-tuple:
tjJ = (C, S, D, G, F, B, P, K, A, M), [3.4]
where, at any given moment,

1) C, the research community, is the part of the psychological commu-


nity and is composed of persons who have received scientific train-
ing, hold strong information links among themselves, and initiate or
continue a tradition of scientific research;
( 2) S is the society (complete with its culture, economy, and polity) that
hosts C and encourages or at least tolerates the activities of the
components of C;
( 3) D, the domain or universe of discourse of tjJ, is the collection of
behavioral and mental states and changes of state (events) of animals
capable of perceiving and learning;
( 4) G, the general outlook or philosophical background of tjJ, is com-
posed of the ontological, epistemological and moral principles that
guide the scientific study of D (for which recall section 3.1);
( 5) F, the formal background of tjJ, is the collection of logical and mathe-
matical theories that are or can be used by members of C in studying
the Ds;
( 6) B, the specific background of tjJ, is the collection of items of knowl-
edge obtained in other fields of scientific inquiry, mainly biology and
social science, and utilizable by the Cs in studying Ds;
( 7) P, the problematics of tjJ, is the collection of problems (actual or
potential) that can be investigated by members of C;
( 8) K, the fund of knowledge of tjJ, is the collection if items of knowledge
utilized by C and obtained by it at previous times;
( 9) A is the set of aims or goals of the members of C with regard to their
study of Ds-namely the description, explanation, and prediction of
behavioral and mental states and events;
(10) M, the methodics (often misnamed 'methodology') oftjJ, is the collec-
tion of methods utilizable by members of C in the study of Ds-in
particular the general scientific method and the experimental method.
58 3. Approaches to Behavior and Mind

In addition, scientific psychology satisfies two basic conditions:


(a) I/J has strong permanent links with other scientific disciplines, in par-
ticular mathematics, biology (especially neuroscience), and social sci-
ence (especially anthropology and sociology);
(b) the membership of everyone of the last eight components of I/J
changes, however slowly, as a result of inquiry in I/J as well as in
related fields.
The first three components of the lO-tuple [3.4] constitute what may be
called the material framework, and the last seven the conceptual frame-
work, of psychology. The former may be so called because both the
research community C and its host society S are concrete (material,
though not physical) systems, and the domain D of facts of central interest
to psychologists is a collection of states of, and changes of states in,
material things, namely animals of certain kinds. On the other hand, the
remaining components of [3.4] are conceptual: They are ideas-though
not, of course, freely floating ones. The moment a person starts research
work in scientific psychology, he or she becomes a member of C and is
expected to make good use of the scientific tradition entrusted to him or
her, as well as to contribute to the enrichment of the problematics, the
fund of knowledge, or the methodics of his or her science.
The community (C) is one of scientific researchers-of seekers and
doubters-not of believers. This must be emphasized in view of the exis-
tence of a number of schools of nonscientific psychology composed of
believers in, commentators on, or appliers of bodies of untested or refuted
conjectures, such as those of Freud and Lacan. As for the host society S,
it must be mentioned because every society stimulates or inhibits certain
kinds of psychology. We have all heard about governments that have
discouraged psychological research-for instance, by way of drastic cut-
backs in research budgets-and of others that have tried to put it under
the control of the dominant ideology.
The domain (D) of I/J includes mental phenomena occurring in animals
but not disembodied minds; postulation of the latter is proper for theology
and parapsychology, not for scientific psychology. This point ties in with
the next component, namely the general outlook or philosophy G. As we
saw in section 1.4, this includes some philosophy of mind-the more
explicit the better because the more easily controllable. Now, of the three
philosophies of mind that have exerted the strongest influence on psychol-
ogy-namely idealism, positivism, and naturalism (or materialism)-the
third is the more congenial with the scientific world view and the one that
has fueled the biological approach to behavior and mind, which is the
more comprehensive and promising (section 3.4). For these two reasons
we submit that the philosophy of mind that best serves the interests of the
advancement of psychology is the one that postulates the identity of
mental states and brain states.
3.5. Scientific Psychology 59

Theformal background (F) of present day psychology is rather modest,


and part of it, namely logic, is tacit rather than explicit. But so was the
formal background of physics before Newton. We should set no limits on
F, because we do not know which kinds of mathematical tools future
psychologists may find useful. Let us only recall that mathematics is a
formal science, hence not married to any domain of facts, thus portable
from one field of knowledge to another (Bunge, 1985a, ch. O. There is no
"mathematics of psychology" any more than there is a "mathematics of
biology"; at most, there are branches of mathematics known to some
psychologists (or biologists) and therefore used by them to formulate
hypotheses or theories. In principle all of mathematics is utilizable by
psychology. Hence the formal background F should include, just as a
precautionary measure, the full range of mathematical theories.
The specific background (B) of psychology has been expanding quickly
in the course of our century. Psychologists need to know more and more
about biology, and even chemistry and physics; and some of them are
forced to borrow from social science, particularly from anthropology and
sociology.
The problema tics (P) of psychology also has been expanding rapidly in
recent times. Psychologists study animals and people, just as zoologists
do, but they address specific problems about their subjects (e.g., how they
learn, or fail to learn, or adapt to new circumstances). And, like neuro-
scientists, psychologists study nervous systems, though not in all animal
species, and always with an eye toward explaining behavior and mind in
neural terms.
The fund of knowledge (K) of psychology is still modest and it includes
items of folk psychology as well as an unknown number of hypotheses
that are bound to prove untestable or false. The field is difficult, the
scientific approach to it is young, the experimental controls are not al-
ways easy to set up, the differences between individuals are often great
and, last but not least, the dead weight of pre scientific philosophy is still a
burden. However, K is growing.
The aims (A) of basic psychology are the same as those of any other
basic science. On the other hand, those of applied psychology, in particu-
lar clinical psychology, psychiatry, and educational psychology, are prac-
tical rather than cognitive. (More on this in chapter 12.)
The central piece of the methodics M of psychology is, of course, the
scientific method and, in particular, its application to empirical research,
namely the experimental method. Note that we distinguish between the
two, because theoretical psychology includes only conceptual proce-
dures; it is up to the experimental psychologist to put the hypotheses and
theories framed by the theorist to the test.
Unlike its nonscientific, though popular, cousin, scientific psychology
interacts strongly with other branches of science. It shares this feature
with all the other genuine sciences. On the other hand, the folk and
60 3. Approaches to Behavior and Mind

pseudoscientific psychologies are characteristically marginal to the sys-


tem of science. In science the only valuable independence is the indepen-
dence of judgment: A totally autonomous discipline, one that neither
borrows nor lends, is at best harmless wild speculation, at worst danger-
ous quackery.
Finally, unlike a body of beliefs, scientific psychology is always in flux.
And, unlike an ideology, which evolves if at all as a result of infighting or
external pressures, scientific psychology evolves as a result of research
within itself or in adjoining fields, particularly neuroscience, social sci-
ence, and mathematical statistics.
Our characterization of scientific psychology embraces its various as-
pects and consequently the various ways in which it can be seen. It
accounts for the social aspect: Psychological research is not done in isola-
tion but in a community C embedded in a society S. It covers the tradition
from which every researcher starts: the general outlook G, the formal
background F, the specific background B, and the items of knowledge that
he or she can take from the fund of knowledge K and the methodics M of
his or her time. Our characterization covers also the view of psychology
as a body of knowledge, namely K. And it includes the view of science as
an activity, namely that of tackling the problematics P with the aims A,
the methodics M, and the background knowledge formed by G, F, B, and
K-as well as staying in close touch with other sciences.
It would have been mistaken to attempt to define the concept of scien-
tific psychology (or any other science) by a single trait, as has been the
tradition in philosophy. (Remember the simplistic characterizations of
science in terms of either induction, or refutability, or use of mathemat-
ics, or freedom from controversy, or some other single property.) In
particular, an indication of the domain or subject matter D is insufficient,
because any collection of items may be studied in various manners, scien-
tific and nonscientific. Nor would an indication of the fund of knowledge
be sufficient, for a body of knowledge may be taken on faith rather than as
a springboard for further research-not to speak of the impossibility of
listing all the achievements of psychology to date. Nor, finally, is the use
of the scientific method a guarantee that a research project be scientific,
for it may bear on the ghostly, or it may be based on a totally wrong
general, formal, or specific background. We had to contrive a fairly com-
plex definition of the concept of scientific psychology because it is in fact
a very complex object.
So much for basic scientific psychology. The applied branches of psy-
chology will be examined in chapter 12. Suffice it to anticipate here that
clinical psychology (in particular neuropsychology), psychiatry, educa-
tional psychology, and the like share nearly all the features of basic psy-
chology, except that they focus on human beings and have practical as
well as cognitive aims. Moreover, applied psychologists are called upon
to make many more value judgments than are their colleagues in basic
3.6. Summing Up 61

research, and they are committed to moral guidelines, deriving from the
Hippocratic oath, that do not apply to animal research. However, these
differences do not prevent applied psychology from being scientific or
from interacting vigorously with basic research.

3.6 Summing Up
One and the same item of psychological interest may be approached in a
number of ways, some of which are mutually compatible. The crucial
distinction between approaches is that between the scientific and the
nonscientific ones. But it is not the only one; we must also reckon with the
holistic, the atomistic, and the systemist approaches.
This century has seen a progression from mentalism to behaviorism to
psychobiology. This movement has been accompanied by a shift in the
underlying philosophies-idealism, positivism, and naturalism (material-
ism) respectively. It has also been accompanied by a shift from holism to
atomism to systemism. The upshot has been an increase in methodologi-
cal rigor as well as a broadening in background, problematics, and aims,
all of which has enhanced the scientific status of psychology.
These changes have not consisted in total replacements, ruptures epis-
temoiogiques, or scientific revolutions a la Bachelard (1938), Kuhn
(1962), or Feyerabend (1975). Instead, they have been phases in an evolu-
tionary process that has conserved some features of the research enter-
prise while altering others. There is no science without some tradition.
The point is not to destroy tradition but to promote its evolution.
CHAPTER 4

Methodology

Methodology is the normative branch of epistemology or theory of knowl-


edge. It does not study how animals and people actually go about solving
problems; this is a concern of cognitive psychology. Instead, methodol-
ogy studies the best research strategies and tactics, that is, those more
likely to attain truth and depth. For example, methodology studies,
among other things, the scientific method, whereas descriptive epistemol-
ogy and cognitive psychology study the trial-and-error procedure. But of
course methodology is not restricted to examining the scientific method as
it is used in all the sciences: It also studies the specific methodics of each
research field-scientific, technological, or humanistic. In short, method-
ology studies all the regular or standardized procedures for gaining genu-
ine knowledge.
N or is this the sole task of methodology. It also studies a number of
general concepts that scientists, technologists, and humanists employ
daily, albeit in an intuitive fashion (i.e., without pausing to analyze them).
These are, among others, the concepts of hypothesis, law, and theory; of
definition, axiom, and theorem; of observation, measurement, and experi-
ment; of description, explanation, and prediction, as well as their numer-
ous kin. Methodology studies them in general, and leaves to the specialist
the study of particular hypotheses, definitions, explanations, measure-
ments, and so on. But, of course, a realistic methodology will draw inspi-
ration from such particulars and it will check the rules it comes up with
against scientific, technological, or humanistic practice. It should thus be
always on the move instead of standing still and becoming dogmatic.
Why should psychologists care about methodology? Because they, per-
haps more than anyone else, encounter tough methodological problems in
the course of their research or practice. By handling these problems,
sometimes in collaboration with neuroscientists and at other times jointly
with social scientists, psychologists have in recent years considerably
enlarged their panoply of special methods (techniques).
Like every progress, any advancement in techniques is double-edged.
On the one hand, it increases the quantity and quality of data. On the
4.1. Method 63

other hand, the sophistication of some of the new techniques calls for the
formation of specialists in a single technique. This development is not
healthy because the point of having multiple access to a domain of facts is
to look at it from different angles and to check the results of one proce-
dure with the help of another. As it is, the specialist in mental testing
tends to overlook the results of the EEG expert, who in turn may ignore
those of the psychopharmacologist, and so on. Some methodological re-
flection would teach researchers that they should not allow any technique
to dictate their ideas, for techniques are supposed to be means, not goals.
Some of the methodological problems faced by psychologists are em-
pirical, such as how to obtain information from a given region of the brain;
others are conceptual, such as how to compare two rival theories.
Whether empirical or conceptual, every problem about method tackled by
psychologists is bound to share some features with problems in other
sciences. Consequently both the methodics and the methodology of psy-
chology have a general part-common to all of the factual sciences-and
a specific one, peculiar to psychology. For example, the problems of
characterizing measurements and theories in general belong to general
methodology, whereas the questions of defining concepts denoting mental
abilities, and of devising objective indicators of the latter, belong in the
special methodology of psychology.
In this chapter we shall only sample the rich general methodics of
science and, in particular, that of psychology. (For the former see Bunge,
1967a, 1967b, 1983a, 1983b; for the latter see Bredenkamp & Feger, 1983;
Sarris, 1986.)

4.1 Method
A method is a prescription for doing something, that can be formulated in
an explicit manner. It is a rule, or set of rules, for proceeding in an orderly
fashion toward a goal. A method can then be formalized as an ordered n-
tuple every member of which describes one step of the procedure: First
do this, then that, and so on. (Contemplation, intuition, and guessing are
procedures but not methodical ones because they are not rule-directed.)
Introspection, or self-observation, is a good example of a procedure
that passes for a method without being one. It has occasionally been
claimed that introspection does not even exist for, strictly speaking, it is
impossible to turn one's gaze inward. But this is sophistry. There is no
denying that we can register and examine some of our own mental pro-
cesses: What else is consciousness? However, such monitoring and re-
flection are haphazard, not methodical, even though they can be some-
what disciplined. (Guessing, seducing, and other activities can be
educated, but there are no methods for carrying them out successfully.) In
short, introspection exists even though there is no such thing as the intro-
64 4. Methodology

spective method. Moreover, introspection is an indispensable component


of psychological research; without it not even the simplest psychophysi-
cal experiment would be possible. We shall come back to it in section 4.2
and chapter 5.
A technique is a method for accomplishing something very special, be it
of cognitive or of practical value. We shall stipulate that a technique, or
special method, is scientific provided it is countenanced by a body of
scientific knowledge. More precisely, a technique will be said to be scien-
tific if, and only if, (a) it aims at an attainable goal, (b) it is reasonably
effective (i.e., it does help attain the goal in a large percentage of cases),
(c) it is intersubjective (i.e., yields roughly the same results for all compe-
tent users), (d) it can be checked or controlled by alternative methods,
and (e) there are well-confirmed hypotheses or theories that explain how
and why it works.
A method that complies with only the first three or four preceding
conditions will be said to be semiscientific, and one that fails to satisfy
them altogether, nonscientific. There is some hope for semi scientific
methods, none for the nonscientific ones. In fact, it may be possible to
perfect a semi scientific method to the point of turning it into a scientific
one. This has been the way of many a rule of thumb in science and
technology.
Free association, as used in the psychoanalytic consulting office, is a
good example of a nonscientific technique, for it yields different results
for different therapists, its results are not comparable to those of alterna-
tive methods, and, although psychoanalytic theory does explain why the
technique should work, it happens not to be a scientific theory. (See
section 5.5.)
On the other hand, the Rorschach or ink-blot test is a good example of a
semiscientijic technique. It rests on the correct empirical datum that psy-
chotics and normal people tend to interpret the inkblots in different ways.
But no well-confirmed theory explains why the test should work, and the
claim that the test can be used to trace personality profiles is just commer-
cial hype. The memorization of nonsense syllables, widely used to study
verbal learning and memory, is in a slightly better position. In fact, its aim
is attainable, it is reasonably effective, it is intersubjective, and it can be
checked by alternative techniques. Still, it is only semi scientific because it
rests on the false hypothesis that memorizing nonsense syllables, such as
sep and pes, does not depend on previous knowledge, so that the "blank
slate" condition is being reproduced. This hypothesis is false because to
memorize such syllables, or anything else, we are likely to associate them
with familiar words, images, or situations (e.g., 'sep' with 'septic' and
'pes' with 'pest'). Like all learning, verbal learning is a growth process
rooted in previous experience, not one of a haphazard piling up unrelated
items.
4.1. Method 65

The invention of new techniques, as well as the novel combination of


old ones, are important aspects of research in any discipline, particularly
in a young one such as psychology. The learning of techniques is also an
important part of the training of researchers. However, one should not
forget that methods are means, not goals, and that training exclusively in
techniques forms technicians, not scientists. Not that there is anything
wrong with being a good technician, for without his or her contribution no
research work would be possible nowadays. However, it must be realized
that technicians are not the same as scientists. Left to themselves, techni-
cians are likely to engage in routine work or in tinkering. On the other
hand, as members of research teams they put their skills and ingenuity in
the service of original research, their tinkering may lead to relevant im-
provements, and the difficulties they meet may be transformed into inter-
esting scientific or technological problems.
Some techniques have evolved into general methods, that is, methods
utilizable in several research fields. The method of successive approxima-
tions by iteration, microscopy, and statistics are well-known examples of
such broadening of scope. But, of course, the method with the widest
scope is the scientific method. Yet, curiously enough, although every
researcher uses it, everyone seems to conceive of it in his or her own way,
and a few even claim that it does not exist. The most popular version
equates it with the so-called inductive method, which would be the se-
quence: data-induction (compression of the data into an empirical gener-
alization)-prediction-test (in cases other than those included in the data
base).
Although induction works well in elementary cases, it does not cover
the most interesting one, namely when the generalization (hypothesis or
theory) is noninductive because it contains concepts that fail to occur in
the data. Well-known cases of this kind are theoretical mechanics and
cognitive psychology. In fact, the basic equations of classical mechanics
include magnitudes, such as mass, force, and internal stress, that are not
directly measurable. Likewise, the hypotheses of cognitive psychology,
whether classical or physiological, do not contain any behavioral vari-
ables. Given the narrow scope and depth of induction, we cannot equate
the inductive method with the scientific method. (See Popper, 1959 and
Bunge, 1967a, 1967b, for further criticisms of inductivism.)
We take the scientific method to be the following ordered sequence of
cognitive operations (Bunge, 1983a, chap. 7, sect. 2.2):

1. Identify a problem (whether gap or dent in some body of knowl-


edge)-if possible an important bit of ignorance. If the problem is not
clearly stated, go to the next step, otherwise to Step 3.
2. State the problem clearly, if possible in mathematical terms or in
terms of measurement operations.
66 4. Methodology

3. Search for information, methods, or instruments likely to be relevant


to the problem-e.g., empirical data, theories, methods of computa-
tion or measurement, measuring instruments, and so forth. That is,
scan what is known to see whether it can help solve the problem.
4. Try to solve the problem with the help of the means collected in the
previous step. Should this attempt fail, go to the next step; if not, to
Step 6.
5. Invent new ideas (hypotheses, theories, or techniques), produce new
empirical data, or design new experiments or new artifacts that prom-
ise to solve the problem.
6. Obtain a solution (exact or approximate) of the problem with the help
of the available conceptual or material means.
7. Derive the consequences ofthe tentative solution thus obtained. Ifthe
solution candidate is a hypothesis or a theory, compute predictions or
retrodictions; if new data, examine the effect they may have on exist-
ing ideas; if new experiments or artifacts, assess their possible uses
and misuses.
8. Check the proposed solution. If the solution candidate is a hypothesis
or a theory, see how its predictions fare; if new data, try to replicate
them using alternative means; if new techniques or new artifacts, see
how they work in practice. If the outcome is unsatisfactory, go to the
next step, otherwise to Step 10.
9. Correct the defective solution by going over the entire procedure or
using alternative assumptions or methods.
10. Examine the impact of the solution upon the body of background
knowledge, and state some of the new problems it gives rise to.
The most crucial steps are the first (problem "finding") and the fifth
(invention or discovery). The most exhilarating experience is that of find-
ing that the solution is correct (Step 8) or that it has a significant impact on
the body of antecedent knowledge (Step 10). Table 4.1 exhibits schemati-
cally the scientific treatment of three typical problem types in psychology:
one experimental, one theoretical, and one practical.
It is, of course, perfectly possible to conduct first-rate scientific re-
search without having an explicit knowledge of the scientific method, just
as Moliere's Monsieur Jourdain had been speaking prose all his life with-
out knowing it. But a knowledge of methodology is necessary, though
insufficient, to evaluate the scientific credentials of theories and practices.
This is particularly so in the case of young sciences, where new ideas
have often to fight old superstitions.

4.2 Observation
The oldest, and still most basic, of all data-gathering procedures is obser-
vation. Observation can be casual or methodical. If the latter, it can be
scientific or nonscientific. A scientific observation is one conducted in an
TABLE 4.1. Three typical problem types in psychology.
Step Empirical problem: measure Theoretical problem: explain Practical problem: treat patients
I What is the value of X? Why does X have the value x? How can the value of X be altered?
2 Which is the measured value of X to within Which premises entail that the value of X is What kind of treatment is capable of alter-
error e? x? ing the values of X?
3 Does experimental arrangement Y help to Do theory Y, subsidiary hypotheses h, and Is treatment Yeffective in altering the
measure X with error less than e? data d imply that the value of X is x? values of X?
N
"""
4 Perform measurement of X with means Y. If Compute the value of X with the help of Y, Use treatment Y. Should there be no im-
the result is implausible, go to Step 5, h, and d. If the result is inadequate, go to provement, go to Step 5, otherwise to 7. ocr"
otherwise to 7. Step 5, otherwise to 7. CJ)
(1)
5 Design new technique Y'. Invent new theory Y' or new subsidiary Design new treatment Y'. :2
hypotheses h'. ~
6 Use Y' to measure X. Compute the value of X with the help of Y' Employ treatment Y' in a pilot study. O·
:::I
and h'.
7 What does the outcome of Step 6 imply or suggest?
8 Evaluate the new results. If they are unsatisfactory go to Step 9, otherwise to 10.
9 Look for systematic errors, and correct Look for possible sources of error, and Look for flaws in the design or the test of
them. correct them. Y', and correct them.
10 How does the new result affect knowledge or practice, and what new problems does it pose?

Note: X stands for a behavioral or mental trait.

~
68 4. Methodology

attempt to solve a definite basic or practical problem with the help of


scientific hypotheses and techniques. This excludes self-observation (in-
trospection) but not questioning, provided the answers of the subjects can
be checked for truth or at least are not taken at face value. (Veracity need
not exclude error. Thus, we may ask a subject to estimate an angle to the
best of his ability, and must record his or her answer even if it is wide of
the mark. What matters here is the subject's veracity and our accuracy.)
Psychologists make scientific observations in the laboratory and in the
clinic, in the field and the farm-the latter when they act as applied
ethologists. They make increasing use oflaboratory results obtained using
sophisticated techniques such as studies of regional blood flow in the
brain, CT (computed tomography), and PET (positron emission tomo-
graphy), not to speak of post mortem pathology reports. (See e.g., Swash
& Kennard, 1985.) They also make use of quantitative observations (i.e.,
measurements)-of which more in section 4.3. And they may not only
observe overt responses to stimuli, such as flashes of light, but also some
of the physiological processes mediating between stimulus and response.
However, if such observations involve the deliberate application of stim-
uli and a comparison with a randomly selected control group, they are
experiments, and the latter will be examined in section 4.4.
Psychological observations in the laboratory and in the field are very
much like observations in other sciences, so they do not deserve a special
methodological study. On the other hand, clinical observations have pe-
culiarities found only in psychology and medicine; they are in-depth case
studies, in the sense that they consist in thorough studies of individuals
who are sometimes unique, and they are conducted over a period of
months or years. And yet they are incomplete because they start only
when the patient is admitted to the clinic, and they rarely follow up on the
patient after he or she leaves. Let us inquire into the consequences of
these two features of clinical observation.
Traditional psychology, like traditional medicine, was based on a few
case studies. Although this method yielded some remarkable results when
employed by insightful researchers, such as Piaget, it also harbored wild
speculation, such as Freud's, as well as trust in the predictive power of
individual observation-which trust proved to be utterly unwarranted.
(See, e.g., Paunonen & Jackson, 1985.) By contrast, contemporary psy-
chology and medicine admit almost exclusively group studies. The advan-
tages of the latter are well known: They spare us our being misled by
exceptions, they allow us to discover central trends, and they are the only
ones capable of SUbjecting hypotheses to rigorous tests.
However, the group-study method has shortcomings of its own. In
particular, it cannot be used when the population is small (e.g., extremely
talented people, aphasics of a rare type, and the amnesic). Besides, it
erases certain traits through lumping into broad categories or through
averaging. (Example: A given trend line is y = x in group A, and y = -x in
4.2. Observation 69

group B. The result of averaging over the two groups is y = 0.) For these
reasons the case study or idiographic method is now coming back, though
with a vengeance. In fact, it is experimental rather than exclusively obser-
vational, quantitative rather than qualitative, and it uses finer categories,
sometimes exemplified by a single subject (Shallice, 1979).
The second feature of clinical studies that we pointed out is incomplete-
ness. Normally the patient is a newcomer to the clinic. The psychologist
(or neurologist) does not have a reliable record of the patient's behavioral
and mental abilities before the onset of the disorder ("pathology").
Hence he or she cannot evaluate the patient's behavioral or mental deficit
except by comparison with normal subjects of the same category (age
group, sex, educational background, etc.) and, consequently, may be led
to believe that the patient has a pronounced deficit of a given kind, with-
out being able to ascertain whether the patient had the same deficit,
though perhaps in a less pronounced manner, long before he or she was
seen at the clinic.
Even a knowledge of the initial and final state of the patient is insuffi-
cient, because it does not tell us anything about the time pattern of the
disease or treatment for different groups of patients. Such patterns must
be known because different patients respond differently to apparently
identical insults as well as to one and the same treatment. For example,
drugs have different effects on different patients, if only because no two
subjects have exactly the same metabolism or come to the clinic in ex-
actly the same state. For these reasons it is desirable to combine case
studies not only with statistical techniques but also with time series analy-
ses (Keeser & Bullinger, 1984).
Having emphasized the limitations of observation, and particularly of
the spotty observations conducted in the clinic or the hospital ward, let us
now note the indispensable role of studies of behavior in natural settings
such as the home, the school, the workplace, the club, or the street
corner. In such natural settings the subject is less likely to wear a
"mask," is not particularly anxious, and faces real-life problems rather
than the problems invented by the experimenter. In such situations it is
possible to watch the way the subject behaves most of the time: when
alone or in his or her relations with relatives, friends, coworkers, supervi-
sors, strangers, and so on. The problematics tackled by field studies is
thus richer and less arbitrary than the one accessible in the laboratory, or
even the clinic or the hospital ward. See Table 4.2.
The problem with field studies is that they are hard to conduct in a
scientific manner, because the student can hardly measure, much less
control, the relevant variables. The use of hidden cameras and the collab-
oration of confederates, well-known techniques in social psychology,
may be necessary but is insufficient. Unless some behavioral and physio-
logical variables are controlled in artificial settings, the results will be
ambiguous. To be of scientific value, naturalistic studies should corne at
70 4. Methodology

TABLE 4.2. Natural versus artificial settings and studies.


Field studies Clinical studies Hospital Laboratory
Setting Natural Artificial Artificial Artificial
Physical stimuli Uncontrolled Partially con- Partially con- Controlled
trolled trolled
Observed Observed Observed Measured
Social stimuli Uncontrolled Mostly ignored Mostly ignored Controlled
Observed Measured
Disturbances Maximal Moderate Moderate Minimal
Tasks Natural and Artificial and Artificial and Artificial and
plenty few few numerous
Overt behavior Observed Observed Observed Measured
Subjective Conjectured Conjectured Conjectured Conjectured
experience
Kinds of re- All kinds Concerning Concerning All kinds
search prob- pathologies pathologies
lem
Kinds of hy- All kinds Concerning Concerning All kinds
pothesis diagnosis, diagnosis,
treatment, treatment,
and prognosis and prognosis
Testability Very weak Weak Weak Strong
Outcomes Problems, data, Problems, data, Problems, data, Problems, data,
and weak and weak and weak and hypothe-
hypotheses hypotheses hypotheses ses of all
strengths

the beginning and at the end of research cycles that include experimental
stages. At the start, to supply problems and data; at the end, to test the
hypotheses conceived in an attempt to solve the problems, and to provide
new data, particularly those gathered in the laboratory or the clinic to
check the hypotheses. In short, field and clinical studies supplement
rather than rival experimental research. (For the characteristics and prob-
lematics of field research see Tunnell, 1977; Patry, 1982.)
So much for the observation of others. What about self-observation, or
introspection, a major tool of classical psychology? Behaviorists and
neobehaviorists have vigorously criticized the use of introspection in psy-
chology. Some of them have gone to the extreme of denying its existence,
arguing that a person cannot observe that which is doing the observing.
This would be true only if there were no parallel thought processes-but
there are. The truth is that introspection provides neither the best nor the
worst access to the mind. It is as indispensable as it is imperfect. Let us
explain.
Among the valid criticisms of introspection we find the following. First,
introspection is a procedure but not a rule-directed one: It is not a method
proper in the sense of section 4.1. Second, introspective data are unrelia-
ble. For example, reports on motives for certain deeds are just as suspect
as memories of episodes in the distant past. Third, many psychologically
relevant data are unavailable to introspection for not being conscious.
4.3. Measurement 71

Fourth, reporting on one's own sUbjective experiences may interfere with


the latter. (Hence "think-aloud" protocols, though necessary, are not
quite reliable.) Fifth, language is not a faithful mirror of mind: "Language
is only as fine a tool as the discriminations it contains" (Osgood, 1953, p.
647)-which are not many.
However, other criticisms of introspection are injustified. For example,
it is not quite true that all the data it yields are unverifiable. Sometimes
ways are found (e.g., electrophysiological measurements) to check intro-
spective reports. It is also false that the external observation of behavior
and of neurophysiological processes can yield all the data supplied by
introspection. Unless the subjects tell us what they are feeling, perceiv-
ing, or thinking, we will not be able to design objective (behavioral or
physiological) indicators of such subjective processes. In this regard the
physician who practices internal medicine and the psychologist are in the
same boat: Both rely on introspection as well as on overt symptoms and
objective tests. In short, introspection is indispensable, but it must be
checked and supplemented with objective tools.

4.3 Measurement
Measurement is quantitative observation, or the observation of quantita-
tive properties such as frequencies and concentrations. Therefore a me-
thodical treatment of measurement must start with quantitation, or the
formation of quantitative concepts (such as that of distance) representing
quantitative properties (such as that of separation).
Most properties come in degrees or intensities, objective, such as volt-
age, or subjective, such as pain intensity. Call S the collection of such
degrees, and assume that it is simply ordered by a relation:::;). That is, if x
and yare in S, then either x :::;) y or y :::;) x, and both statements hold just in
case x - y. For many purposes this comparative concept is insufficient,
and we are required to form a quantitative one. We achieve this by map-
ping S on to a set T of numbers in such a way that (a) each degree
(member of S) is assigned a single number in T, and (b) the order in S is
preserved in its numerical image T. In short, the quantitation of S consists
in introducing a mapping or function M from S to T, or M: S ~ T for
short, where T is included in the real line, and such that, for any x and yin
S, x ;:$ y if, and only if M(x)s M(y). Any function M with these charac-
teristics is called a magnitude.
(In general, it is possible to form different magnitudes representing one
and the same property: Think of the different temperature scales. More-
over, in general, the domain S of M is not made up of single items but of n-
tuples, such as (a, b, c, d), where a and b name physical objects, c a
reference frame, and d a distance unit. In other words, in general, the
domain S of M is the Cartesian product of certain sets. For example, the
72 4. Methodology

distance function in relativistic physics is of the type D : P x P x F X U d


---?!R+, where P is the collection of all possible physical objects, F that of
reference frames, and Ud that of all possible distance units, whereas !R+ is
the set of positive real numbers.)
Magnitudes may be classed into extensive and intensive, according to
whether they are additive, or nearly so, or not at all. More precisely, a
magnitude M is said to be extensive if its value for an arbitrary composite
object x y equals at most the sum of its values for the components, that
0

is, if M(x 0 y) ::.; M(x) + M(y)-otherwise M is called intensive. Here x 0 y


stands for the object composed of the objects x and y; the composite
object may but need not be a system. Length and weight are extensive,
age and intelligence are not.
If the equality sign holds in the preceding, the magnitude is called
additive, otherwise subadditive. Volumes and populations are additive,
whereas entropies and prices are subadditive. The most important magni-
tudes are intensive, for they generate the corresponding extensive magni-
tudes. For example, the total mass (or charge) of a body equals the vol-
ume integral of the mass (or charge) density, which is an intensive
magnitude. Whether a magnitude is intensive or extensive is not a matter
of convention but law: It depends upon the law(s) in which it occurs. For
this reason it is unwise to go much further into the matter of quantitation
in general. There is an additional reason for caution in this domain,
namely that the invention of a magnitude is not a rule-directed operation,
even though it is reasonable to suspect that it fits objective psychobiologi-
cal laws-alas, so far unknown ones.
Genuine quantitation obeys an empirical condition as well as a mathe-
matical one. It must be accompanied by an indication that there are
known or thinkable, direct or indirect ways of effectively assigning (mea-
suring) some values of the function (magnitude) in question. Otherwise
the quantitation in question may be judged phoney. Herbart's mathemati-
cal psychology was empty for that reason. (See Miller, 1964.) By the way,
much of the mathematical literature on psychological measurement is
phony for the same reason. Although it is mathematically rigorous, and
even sophisticated, it is restricted to the simple case of additive magni-
tudes and, worse, it is totally irrelevant to measurement proper. The
reason for this is that it rests on the confusion between measurement, an
empirical operation, and measure, degree, or intensity (i.e., that which
measurement operations try to determine) (Bunge, 1973a). We shall there-
fore keep clear of that comedy of errors.
Once we have formed a quantitative concept (magnitude) suspected of
faithfully representing a property of interest, we may face the problem of
measuring the latter, that is, of finding (some) values of the function in
question. If the property happens to be directly observable, as is the case
with some behavioral and physiological variables, the measurement can
be fairly direct and therefore of little methodological interest. But in most
4.3. Measurement 73

cases the property of interest happens to be inaccessible to direct obser-


vation. Think of field intensities, atomic masses, astronomical distances,
geological ages, mental abilities, or international risks. In all such cases
we must hit on adequate objectifiers or indicators, that is, observable
properties that are lawfully linked to the unobservable ones that we wish
to pin down. For example, the concentration of noradrenaline in blood is
used as a stress indicator, and rapid eye movement during sleep as a
dream indicator (whereas daydreaming is indicated by reduced ocular
motility).
The unobservable-observable links used to be called operational defi-
nitions (Bridgman, 1927). Actually they are not definitions (a kind of
convention) but fallible hypotheses. If scientific, these hypotheses are
testable, whence they had better be called indicator hypotheses. An un-
ambiguous indicator hypothesis maps observable values on to unobserva-
ble ones, in such a way that measuring the former we can infer the latter
via some formula. In self-explanatory terms, U = flO). For better or for
worse, most indicators are ambiguous and therefore fallible (i.e., they are
one-to-many relations rather than functions). For example, a slowdown in
speech delivery may indicate mere tiredness or serious neurological trou-
ble. Fortunately, the ambiguity inherent in an isolated indicator hypothe-
sis may be removed by using two or more indicators simultaneously. In
other words, unobservables are best ferreted out with the help of whole
batteries of mutually compatible indicator hypotheses. Preferably, every
one of these is a member of a well-confirmed theory rather than a stray
empirical conjecture. The presence of such high-grade indicator hypothe-
ses is in turn an indicator of a high level of advancement in the research
field in question. Still, in the beginning we may have to settle for empirical
indicators.
So far we have dealt with two conceptual preliminaries to measure-
ment, namely quantitation and the finding of objective indicators. The
next step is the design of a measurement technique, which, if scientific,
will use well confirmed scientific theories. Each property calls for its own
measurement technique-preferably a whole family of techniques, so that
we can use some of them to check the results obtained with the help of the
others.
From a philosophical viewpoint, the interesting difference between
techniques is that between invasive (or obtrusive) and noninvasive (or
unobtrusive). A measurement technique is said to be invasive in case it
alters in a significant way the state of the object of measurement, other-
wise it is noninvasive. A sodium amy tal injection is a (mildly) invasive
procedure, whereas a question about a trivial matter is noninvasive. Psy-
chologists, like all other scientists, employ techniques of both kinds. And
when they resort to invasive techniques they attempt to minimize their
effects or, at least, to determine by independent means the magnitude of
the disturbance they have introduced. For example, although all subjects
74 4. Methodology

change their behavior when entering a psychological laboratory, they


change it the least if they are not constantly aware that they are under
observation; for this reason hidden cameras and semitransparent mirrors
are often used.
It is often claimed (e.g., by Valentine, 1982) that the use of invasive
measurement techniques taints research with artifact. Accordingly, nei-
ther experimental psychology nor atomic physics could possibly be objec-
tive, hence scientific. This is not true in physics: (a) most calculations in
theoretical physics refer to things, such as atoms and photons, that are
not being subjected to any experimental manipulations; (b) the effects of
the interference on the part of a measuring instrument, when they exist,
can be calculated, at least in principle, if we know how the instrument
works; and (c) all good experimental designs keep the observer at arm's
length, precisely in order to maximize objectivity. The situation in experi-
mental psychology is in principle the same, only more difficult in practice
due to the dearth of theories explaining how indicators and instruments
work. Only a perverse philosophy of science could suggest that measure-
ment, a major guarantor of objectivity, could make objectivity impossible.
In conclusion, once we have built an adequate magnitude representing
the property of interest, devised a faithful indicator, and designed a suit-
able measurement technique, we may proceed to performing a run of
measurements. The outcome of the latter will be a collection of rational
numbers (fractions) read off, say, a dial. The entire process is represented
schematically in Figure 4.1.
The living-brain imaging techniques are among the most interesting
measurement techniques of all. Electroencephalography (EEG) was his-
torically the first. It indicates wakefulness and sleep, and a few gross
anatomical features such as hemispheric asymmetry, but it records only
superficial mass electrical potentials: it has no depth and it has poor
resolution. (One point on the cortex is imaged as a 2.S-cm radius circle
on the scalp.) Consequently EEG recordings are poor indicators of
mental activity. Worse, they have fostered the belief that the brain is

Degrees of property Ouantitation Real numbers and units


(e.g. length) ....-------------1.., (e.g. V2 cm)

FACTS Property-indicator Inclusion CONSTRUCTS


relation

Marks on a dial '-----------I.~. Rational numbers and units


Scaling (e.g. 1,414 cm)

FIGURE 4.1. Measurement as a one-to-one correspondence between the degrees


of a property and instrumental readings. Redrawn from Bunge (l983b).
4.3. Measurement 75

an unstructured whole rather than a system with many distinct


components.
On the other hand, the tomographic techniques (CT scan and PET scan)
have a far better (though still somewhat coarse) resolution, reach every-
where in the brain, show which regions are more active, and allow scien-
tists to make in vivo measurements of the distribution and rates of a few
important chemical reactions, in particular glucose utilization. They have
thus allowed them to image a listening, speaking, or thinking brain-a
sensational victory for physiological cognitive psychology. The philo-
sophical interest of these and other brain-imaging techniques is that they
have jeopardized the dogma of the privacy of mind. Minds are becoming
just as public, though of equally difficult access, as atoms and govern-
ments.
However, psychological measurement has still a long way to go, as is
obvious from the spirited controversies surrounding intelligence testing.
We shall take up this problem in section 9.4. Suffice it to note here that
mainstream mental testing relies so much on language that it is mostly
inapplicable to animals. To be sure, some advances have recently been
made in animal mental testing. However, the techniques evolved so far
are so primitive that there exists no reliable objective ranking of animal
species with regard to intelligence. Worse, our knowledge in this field is
so meager and so contaminated by ideology that, whereas some students
continue to uphold Descartes's thesis that animals are automata deprived
of initiative and creativity, others are inclined to believe all of George
Romanes's anecdotes on the mental feats of pets of various species. Be-
lief in the gradual character of biological evolution is still so strong that a
recent study concludes that "there is no sense to the notion that there are
'lower' and 'higher' vertebrates, at least where intelligence is concerned"
(Macphail, 1982 p. 331).
Finally, a caution concerning the value of observation and measuring
instruments and techniques. Many psychologists and neuroscientists be-
lieve that "scientific progress at any point awaits the discovery of instru-
ments and techniques" (Boring, 1942, p. 609). In particular, it has been
said that "the gains in brain are mainly in the stain" (Bloom, 1975). Even
a cursory examination of the history of science suffices to refute this
thesis. The Newtonian, Darwinian, Marxian, Maxwellian, and Einstein-
ian revolutions did not result from improvements in instrumentation or in
technique: They were conceptual revolutions.
Whereas some advances are made possible by the invention of new
instruments or techniques, others consist in the invention of new hypoth-
eses or theories. After all, no instrument and no technique, however
powerful, lays bare any laws; only hard thinking can hypothesize laws.
And laws are what we most want in science.
Moreover, instruments and techniques are unlikely to bring about any
breakthroughs unless handled with some good ideas in mind. For exam-
76 4. Methodology

pIe, Golgi invented a new tissue-staining technique, but it was Ramon y


Cajal who exploited it to the full because he looked for structure in what
had hitherto been believed to be like a bowl of porridge. And Penfield
found some astonishing facts when applying electrical stimulation to
choice points in the cortex of waking patients, but it was Hebb who made
the most out of these findings by reinventing the hypothesis that percepts,
images, and thoughts are activities of cell assemblies.
We wind up this section with a caveat: "Under certain conditions,
subjective judgments may be just as reliable as some unreliable objective
modern electrical apparatus" (von Bekesy 1965a, p. IS1). After all, our
perceptual systems are extremely sophisticated systems comprising parts
of highly evolved brains capable of supplementing and correcting (as well
as of distorting) the signals received by the sensors. And the processes
going on in such living systems, though subjective, are real and, in princi-
ple, they can be monitored just as objectively as any other natural pro-
cesses (e.g., by electrophysiological means). (We shall return to the prob-
lems of subjectivity and objectivity in section 5.3.)

4.4 Experiment
We all know what an experiment is, namely an observation or measure-
ment involving a controlled change in some of the features of the object
being studied. We also know that the goal of an experiment can be to find
new facts or to test a hypothesis-or both. However, the evidence sug-
gests that many psychologists make blunders in the design or in the inter-
pretation of experiments, because they overlook the conceptual basis of
every experiment. A few examples will suffice to confirm this claim.
Tversky and Kahneman (1971) found that a large percentage of psy-
chologists commit the gambler's fallacy, believing, for example, that the
regular sequence HTHTHT of coin tossings is more probable than
HHHHHH. Others believe that any sample, regardless of size and mode
of obtainment, will do. Still others believe that computer simulations are
experiments, whence computers could replace laboratories. And many
attempts to teach monkeys and apes to perform certain cognitive tasks
failed for not posing the question to the animal "in a way that was mean-
ingful or worth the animal's effort [. . . J so that it might be missing the
point of the experiment" (Weiskrantz, 1977, p. 432). For example, it had
been assumed that, unlike human infants, monkeys are incapable of
cross-modal matching. But when offered chow prepared into different
shapes, some tasty and others adulterated with quinine and sand, the
monkeys that had tried the morsels in the dark, using only their sense of
touch, were able to recognize them in the light (Cowey & Weiskrantz,
1975). Given the importance of the assumptions underlying any experi-
ment, it will be useful to examine this matter.
4.4. Experiment 77

CONCEPTUAL
CONTROLS

~
Hypotheses .
Methods . EXPERIMENTAL
New data
Data .. SET·UP

t
EXPERIMENTAL
CONTROLS

FIGURE 4.2. A well-designed experiment has conceptual (in particular statistical)


controls as well as experimental ones (e.g., of voltage). And in the design and
interpretation of the experiment not only data and methods but also hypotheses
(philosophical, statistical and scientific) take part. (From Bunge, 1983b.)

The design of any experiment, as well as the interpretation of its


results, presupposes a number of hypotheses. Some of these are generic
(i.e., shared by all experiments), whereas others are specific (i.e., charac-
teristic of a given type of experiment). See Figure 4.2. The generic hy-
potheses are of two kinds: philosophical and statistical. The specific hy-
potheses are scientific: they refer to specific features of the experimental
setup. We shall mention only some of the philosophical presuppositions,
for they are seldom if ever dug up. Here is a sample (taken from Bunge,
1983b):
(1) Reality: The members of the experimental and control groups, as well
as the measuring instruments, exist really (autonomously), although
some of the hypothesized objects may be imaginary. (If all of the
things involved in an experiment were figments of our imagination,
then imaginary experiments, e.g., computer simulations, would suf-
fice.)
(2) Lawfulness: All the objects involved in the experiment behave law-
fully, even though we may be unable to predict, hence to control,
some of the effects of the stimulus applied to the objects of the experi-
ment. (There would be no point in performing experiments if things
were to respond erratically to our "questions".)
(3) Causality: All the things involved in the experiment satisfy some form
of the causal principle, however weak, for example, "Every event is
the effect (perhaps with some probability) of some other event." (Oth-
erwise no deliberate production of an effect and no effective control of
variables would be possible.)
78 4. Methodology

(4) Randomness: All the variables involved in the experiment are subject
to some random fluctuation, both intrinsic and due to external pertur-
bations. (Otherwise we would be unable to explain the statistical scat-
ter of many results.)
(5) Insulation: Objects other than the object of experiment, the experi-
menter, and his or her experimental means, can be neutralized or at
least monitored during the experiment. (Otherwise no significant
changes could be attributed exclusively to changes in the control vari-
ables.)
(6) Disturbances or artifacts: It is always possible to correct to some
extent, either empirically or theoretically, for the "artifacts," distur-
bances, or contaminations caused by the experimental procedures. (If
such partial corrections were impossible we could not legitimately
claim that the thing for us-as it appears to us-resembles closely the
thing in itself, such as it is when not subjected to experiment.)
(7) No psi: It is always possible to design experiments in a manner such
that the experimenter's mental processes exert no direct influence on
the outcome of the experiment. That is, the experimenter can be
shielded or uncoupled from the experimental setup, so that his or her
bodily and in particular brain processes do not alter the experimental
results. (Otherwise the outcome of the experiment could be produced
at the experimenter's whim, and the experimenter would be testing
nothing but his or her own mental, for instance, psychokinetic, abili-
ties.)
(8) Explicability: It is always possible to justify (explain), at least in out-
line, how the experimental setup works, (i.e., what it does). In other
words, it is possible to form a conceptual model of the experimental
device using well-confirmed hypotheses and data. (Otherwise we
would be unable to draw any conclusions.)

A few comments are in order. First, according to (1) a scientific experi-


ment presupposes a realistic epistemology. Second, at first sight the law-
fulness principle (2) is refuted by the so called zeroth (or Harvard) law of
experimental psychology: "Any well-trained animal, on controlled stimu-
lation, will respond as it damn well pleases." No such thing: The response
will look arbitrary only if the context, the previous history, and the inter-
nal state of the animal are disregarded. Third, the existence of probabilis-
tic laws does not refute causality (3) but only restricts its scope. In fact,
those laws are typically of the form' 'The probability (or the rate of change
of the probability) that cause c will produce effect e equals p" (see Bunge,
1979b). Fourth, the randomness referred to in (4) is the one inherent in
both the object and its environment; it has nothing to do with another
rationale for using statistical methods, namely the existence of individual
differences ("variability"). Fifth, the insulation condition (5) must be
satisfied at least partially, but it can never be fully enforced, because
4.4. Experiment 79

every thing interacts with some other things. Sixth, the condition (6)
regarding environmental disturbances and experimental "artifacts"
makes it possible to retain objectivity even while altering the object of
experimentation. Seventh, the no-psi condition (7) suggests that believers
in psychokinesis, if consistent, should have no faith in their own experi-
ments. Finally, the condition (8) of explicability is a requirement of ration-
ality: Scientists should pay no heed to blind manipulation. In other words,
one must know a lot before proceeding to design and perform an experi-
ment. Consequently experiment cannot be the ultimate source of knowl-
edge.
The last point is related to the problem of choice of experimental ani-
mal. Experimenters have an understandable tendency to choose animal
preparations that are easily accessible and comparatively easy to work
on, such as the neurons of Aplysia for being large and specialized, squid
axons for being long and tough, rats because they are aplenty and compar-
atively easy to train, Japanese macaques for being sedate and coopera-
tive, and college students for being abundant and easy to communicate
with.
However, the choice of experimental animal should also be guided by
some idea of what kind of thing one would like to find, and what kind of
means might do the trick. Thus, if one wishes to investigate cognitive
processes one may start working on people because of the advantage of
language, but if one wishes to perform electrophysiological measurements
below the scalp, and a fortiori if one wishes to make lesions on normal
animals, one will have to switch to nonhuman primates for ethical rea-
sons. However, it is unintelligent and immoral to tamper with animal
welfare by applying invasive stimuli in a haphazard way. Blind manipula-
tion is alchemy, not science. Science is empirical but not empiricist: Ideas
play in it as important a role as experiences.
Experiment is neither the alpha nor the omega of science; it lies right in
the middle of it, as a synthesis of experience and reason. Experiment is
most useful when designed with the help of some items of scientific
knowledge and when yielding data that will either motivate the invention
of new ideas, or that can be fed into existing theories for the purpose of
explanation, prediction, or testing. However, one can often learn even
from experiments that fail to yield any conclusive results; at least one can
learn that there was a flaw in the design or in the execution. Unsuccessful
experiments are thus more valuable than no experiments at all-that is,
provided they are analyzed with the intention of improving on their design
or execution.
However, there are antiexperimental schools in psychology, notably
psychoanalysis and humanistic psychology. The antiexperimentalist of-
fers several reasons for this attitude: that you cannot measure the imma-
terial soul, that there are no laws of behavior or mentation to be found
(e.g., because no two individuals are identical, or because no experimen-
80 4. Methodology

tal situation can be exactly replicated), or that you must not manipulate
people. The real motive, though, is a deep mistrust of the scientific atti-
tude combined with a bookish attitude towards learning. In any event,
what little solid knowledge of behavior and mentation we do have has
been obtained with the help of experiment. (See e.g., Estes, 1979; Par-
ducci & Sarris, 1984.)
Let us finally dispel a dangerous mistake that has become popular
among those who take literally the brain-computer analogy. We refer to
the belief that computer programs are experiments, whence computer
simulation can replace authentic laboratory experiment on real animals or
people. Thus Newell and Simon (1981, p. 36) have claimed, "Every new
machine that is built is an experiment. . . . Each new program that is
built is an experiment." They do not say just that one can tryout, or
"experiment" with, every new release of the computer industry, but that
every artifact of this kind, whether hardware or software, is an experi-
ment proper. This is a serious mistake because, whereas genuine experi-
ments yield more data than they consume, computer programs are insatia-
ble data guzzlers; and, whereas the former allow us to test hypotheses,
computer programs use both explicit and tacit hypotheses. The mistake is
dangerous because it is a tacit invitation to replace laboratories with
computers, experimentalists with programmers, and the scientific method
with apriorism.
An example of the danger of exaggerating the power of the computer is
the collection of computer models of cerebral lesions and mental ill-
nesses. Take a mathematical model of a neural system characterized by a
connectivity matrix C, and set equal to zero all the matrix elements Cmn
for a fixed m. The resulting matrix will represent the neural system from
which the neuron (or the subsystem) labeled m has been eliminated (e.g.,
by surgical procedures). One can then hope to discover the corresponding
mental or behavioral deficit. This method works only if one has got a
mathematical model, involving the connectivity matrix, that has success-
fully passed some experimental tests. Otherwise the exercise proves
nothing. Unfortunately this caution is not always being observed.
For example, Wood (1982) found that, by simulating lesions in the
Anderson et al. (1977) distributed (nonlocalized) model of memory, cate-
gorization, and other functions, only a quantitative impairment results.
That is, the effect of the lesion is what Lashley's ill-fated "mass action"
(or "equipotentiality") pseudo-law would have predicted: Only the
amount of nervous tissue, not its location, would matter. But given the
nature ofthe mathematical model, this result was to be expected. Indeed,
a tacit assumption of the model is that every neuron in the system is like
every other neuron. That is, no specialization was assumed to begin with,
so that the removal of anyone neuron produced a deficit that, on the
average, was equivalent to that of any other neuron. The computer simu-
lation could not possibly have contradicted any of the hypotheses incor-
4.5. Inference 81

porated in it. Hence it could teach us nothing that we did not know before.
On the other hand, actual ablation experiments do teach a lot-provided
one is careful in "drawing conclusions." But this is a matter for the next
section.

4.5 Inference
Once the experimental data are in, they are supposed to be cleansed and
organized with the help of mathematical statistics. The result will be the
elimination of some outlying data (for being suspected of resulting from
systematic errors in the design or the execution of the experiment), as
well as the correlation or the aggregation of the remaining ones. Having
performed this data reduction job we are confronted with the far trickier
problem of "drawing conclusions" from the results of that processing.
The double quotes around the expression 'drawing conclusions' was
intended to suggest that, strictly speaking, no (logical) conclusions can be
drawn except for the trivial one that some things, when in such and such
circumstances, behave in such and such a manner. On the other hand, the
data may confirm or infirm (and occasionally refute altogether) previously
known but not tested (or at least not well-confirmed) hypotheses or theo-
ries. The data may also suggest new hypotheses-to those prepared to
"see" the underlying pattern. We will not make a methodical study of
such "inferences"-again the wrong word, because there are no infer-
ence rules for performing such leaps. Instead, we shall deal with a sample
of common pitfalls in experimental psychology.
One much debated problem concerns the legitimacy of extrapolating
animal findings to humans. On the whole, the behaviorists have taken it
for granted that all the findings concerning rats, dogs, cats, and even
pigeons, can be extrapolated to humans without further ado. In a way
they were justified in this belief because of the similarity, in many basic
regards, between the nervous systems of all the higher vertebrates. This
similarity, in other words, is the objective basis of the behaviorist goal of
finding cross-species (i.e., non-species-specific) behavior patterns. The
strategy paid off handsomely: A lot was learned from the use of animal
"models" (a further misnomer).
However, there are limits to such extrapolations. For one thing, hu-
mans possess certain neuronal systems (e.g., those that "mediate" ab-
stract thinking and language) that are absent or only very rudimentary in
other animals. For another, some of the neuronal systems common to
man and other higher vertebrates, such as the olfactory bulb, are either far
more or far less evolved in humans than in our relatives. Third, humans
live in an environment that is largely man-made, namely human society,
complete with institutions and artifacts. For these reasons only some
results of animal experimentation can be extrapolated to humans, namely
those that do not involve institutions or artifacts. However, this condition
82 4. Methodology

can only be ascertained by performing similar experiments on humans.


Hence any result from animal experimentation should be treated as a
hypothesis that might apply also to humans.
Speaking of hypotheses, we must warn against the common mistake
that all hypotheses are nothing but inductive syntheses, that is, data pack-
ages and generalizations. There certainly are hypotheses of this kind
(e.g., those obtained by curve-fitting methods). Such hypotheses are in-
dispensable but they have three limitations: one logical, one semantical,
the third epistemological. The logical limitation of such hypotheses is that
they have a very low deductive power-so low in fact that they are the
propositions that ought to occur as theorems in good theories. The seman-
tical limitation is that they contain nothing but down-to-earth concepts,
namely concepts that stand for directly observable variables, which are
often indicators of deeper properties designated by more abstract predi-
cates. The epistemological limitation of inductive syntheses is that they
contain no ideas that were not contained in the data from which they were
inferred. Thus a curve joining points on the Xx Y plane relates the known
concepts X and Y. As C. S. Peirce recognized a century ago, the one
procedure that introduces radically new ideas is what he called "abduc-
tion," and others call "the method of hypothesis." However, inventing
hypotheses is not the only way to advance knowledge, and there is no
method for inventing high-powered hypotheses, that is, conjectures in-
volving concepts, such as those of synaptic strength, evolution, and intel-
ligence, that are not accessible to ordinary observation.
Our third problem is the common complaint that the psychological
literature abounds with mutually inconsistent reports of experimental or
clinical findings. For example, whereas one group of experimenters re-
ports that ablation in brain region A caused behavioral or cognitive deficit
B, another group swears that no such impairment results. (Thus, the early
intelligence tests on patients who had suffered extensive excisions in the
frontal lobe failed to exhibit any deficits. Several decades later severe
deficits were found using different tests, involving initiative and plan-
ning.) Such inconsistencies are sometimes due to faulty techniques, but at
other times they are due to individual differences among the subjects, or
to the difficulty in finding the exact borders of the brain subsystems con-
cerned.
However, at other times the problem results from conceptual vague-
ness. For example, if different workers hold different tacit (and equally
woolly) ideas about consciousness, it is likely that, whereas some ofthem
will declare that the destruction of neural system X affects consciousness,
others will deny it (particularly if they mistake consciousness for alertness
or sensitivity to external stimuli). In such cases it is mandatory to clarify
the key ideas before proceeding to repeating the experiment. Regrettably,
most experimenters have little faith in conceptual clarification; they tend
to believe the empiricist myth that facts speak for themselves.
4.5. Inference 83

D~D
LIGHTS SAMPLE (G)
PRESS
SAMPLE KEY
DELAY
(NO LIGHTS)
PRESS
ONE KEY:
IF G, REWARD
D LIGHTS

FIGURE 4.3. Delayed matching to sample (DMS) task. The panel shows four
buttons that turn off lights of different colors: Green, Blue, and Red.

The frequency of inconsistent results may be reduced not only by clari-


fying ideas but also by trying different animal strains, by trying different
experimental techniques, or by combining two or more of them. The latter
is the rule in physiological psychology. Consider, for example, the de-
layed matching to sample (DMS) task. A typical arrangement aimed at
testing for recall of sensory information is shown in Figure 4.3. The exper-
imental animal faces a panel with four response buttons. A trial begins
with presentation of a sample, in this case a green light (G). The animal
presses the button, thereby turning the light off and acknowledging the
stimulus. A delay of a few seconds follows. Then three different response
colors, among them the sample one (G), appear. The correct choice of the
sample color is rewarded with a squirt of fruit juice. This psychological
experiment is more rewarding if supplemented by implanting electrodes in
the cells suspected of being activated by "set," stimulus presentation,
recall, button-pressing movement, reward, or what have you. (See e.g.,
Fuster, 1984a.)
Such combination of methods is indispensable to "map the mind onto

the brain," that is, to localize behavioral and mental functions. But the
interpretation of such psychophysiological experiments is tricky because
of the interconnectedness of the various subsystems of the brain. To
disentangle them the so-called principle of double dissociation (Teuber,
1959) is widely used. Let BI and B2 be two brain "structures" (subsys-
tems), and FI and F2 two behavioral or mental functions. We can as\.,.:r-
tain that FI is a specific function of B I, and F2 one of B 2, if and only if the
two following conditions are jointly met: (a) the destruction of BI causes
the loss of FI but not that of F 2, whereas (b) the destruction of B2 causes
the loss of F2 but not of Fl. In symbols, lBI :? IF/\F2, whence lBI :?
IF,,which amounts to FI :? B I . Likewise, lB2 :? F/\ lF2, whence lB2
:? lF2, which is equivalent to F2:? B 2. In other words, BI is necessary for
Fh and B2 for F2 • For example, the destruction of the Broca area causes
the loss of speech but not of the understanding of speech; on the other
hand, the destruction of the Wernicke area causes the loss of understand-
ing of speech but not of speech delivery.
84 4. Methodology

The invention of new experimental techniques will be fruitful as long as


it is accompanied by clear ideas and by the realization that every such
method determines largely what will be found-and missed. The early
behaviorists, by studying "mazes with rats in them," forced the animals
to behave in certain ways and were unable to find out anything about
spontaneity and creativity. The physiologists who use exclusively the
EEG succeed in recording the total brain activity and tend to view the
brain as an unstructured whole. At the other extreme, Eccles, by investi-
gating single neurons, could not find the mind. And Lashley, using abla-
tions, was unable to localize any engrams: he unwittingly destroyed them.
Laboratory studies are indispensable but they do not eliminate the need
for either field or clinical investigations. Take for instance the study of
animal sexuality. The advantages of the experimental approach are obvi-
ous: All the parameters suspected of being relevant can be measured, and
some of them can even be varied almost at will. But the restriction to
laboratory animals is likely to obscure an important aspect of mating
behavior, namely strategies of choosing a mate and the accompanying
sexual competition-an important factor in animal evolution. Indeed,
caged animals do not have much choice, so that observation on them
alone is likely to suggest the popular (null) hypothesis of random mating.
However, ethologists have been accumulating evidence against this hy-
pothesis, by combining studies in the wild with the experimental method.
They have found, for example, that the female of the African widowbird
prefers the males with the longest tails. They have established this fact by
cutting the tail of some males and gluing additional feathers to that of
others, and watching the behavior of nearby females (Andersson, 1982).
Speaking of the null hypothesis, its peculiar methodological status de-
serves more thinking than it is usually accorded. A null hypothesis is of
the form "There is no correlation between A and B." But this is only a
particular case of a negative hypothesis, of the form "There is no X" (or
"There are no objects with property P"). Now, negative hypotheses can
be confirmed only in trivial cases, namely when thorough inspection is
possible. For example, we can confirm conclusively that there are no
elephants between the covers of this book. But in the open situations
characteristic of scientific research, negative hypotheses can only be re-
futed, namely by exhibiting at least one item of the kind whose existence
the hypothesis denies. Thus, it cannot be proved conclusively, at least on
the strength of empirical evidence alone, that there is no homunculus
(little man in the head), or no telepathy, or no afterlife. But then it is an
accepted norm of rational argument that the onus probandi for any posi-
tive hypothesis H lies on those who advocate H. In the absence of posi-
tive empirical evidence for H, the skeptical scientist must adopt the null
hypothesis at least for the time being.
Finally, some thoughts about the role of mathematical statistics in psy-
chology. There is no gainsaying the usefulness of statistics before and
4.6. Summing Up 85

after experiment, that is, in designing it as well as in processing its out-


come. But one should not exaggerate the power of mathematical statis-
tics, and one should mistrust the blind application of statistical recipes.
After all, statistics can only help weed out dubious data and establish
correlations (i.e., refute null hypotheses) and trends. It is not a method for
finding laws proper, let alone for building hypothetico-deductive systems
(theories). As a matter offact there is no method for constructing nontri-
vial (in particular counterintuitive) hypotheses out of data. (Interpolation
methods allow one to find only polynomials with adjustable parameters
and devoid of explanatory power; those methods are just data-packaging
techniques. )
Take for instance the case of two variables, X and Y, that exhibit a
strong (positive or negative) and "consistent" (i.e., lasting) correlation.
There are four possible substantive hypotheses underlying this statistical
hypothesis: (a) changes in X cause changes in Y; (b) changes in Y cause
changes in X; (c) a feedback mechanism-changes in X cause changes in
Y, and conversely; and (d) both X and Y depend on a third variable Z. A
priori one of the four substantive hypotheses explains the observed statis-
tical hypothesis. Only a detailed empirical and theoretical examination of
all four substantive hypotheses will allow us to choose among them.
Without mathematical statistics there is nothing-but raw datum, edu-
cated guess, or wild speculation. But as long as mathematical statistics
remains the main formal tool of psychology, the science will not advance
beyond its present pre-Newtonian or proto scientific stage. Statistical so-
phistication cannot compensate for theoretical indigence or experimental
sloppiness. What psychology needs most at present is substantive and, if
possible, deep theories unveiling the mechanisms of behavior and mind,
much as Newtonian mechanics unraveled those of motion. The more
powerful the theories in a research field, the less use is there for sophisti-
cated statistical techniques (Meehl, 1978). For example, few physicists
have heard of tests of significance, let alone of Bayesian inference. This is
due to the fact that they rarely remain content with formulating statistical
hypotheses of the form' 'X and Y covary." They regard these as program-
matic hypotheses to be worked out and eventually replaced with substan-
tive hypotheses. They learned long ago that it pays far more handsomely
to invest in conceptual refinement than in data processing. For this reason
they find it hard to believe that the scientist's brain is a data processor-
an axiom of information-processing psychology.

4.6 Summing Up
Psychological research raises a far richer methodological problematics
than that treated in the standard literature on the subject, which is all too
often limited to statistics or to repeating uncritically the opinions of phi-
losophers who have never stepped into a laboratory.
86 4. Methodology

Some of the trickiest and most neglected of those problems are of a


conceptual nature: They derive from the use of vague concepts (such as
that of consciousness) or of tacit hypotheses (such as that language is a
faithful mirror of the mind). If specific, the presuppositions are bound to
surface eventually. (For example, the "conclusions" derived from the
study of aphasics have often presupposed that the inability to name an
object indicates incapacity to conceive of it. Further studies revealed the
mistake.) But other hypotheses underlying experiment are of a philosoph-
ical nature, hence normally accepted uncritically-or ignored altogether.
Another often neglected subject is that of the nature of objective (be-
havioral or physiological) indicators or objectifiers, which used to be
called "operational definitions." This neglect has had two negative conse-
quences: one is the dearth of objective indicators, another is the accep-
tance of phony ones. (For example, many psychologists continue to be-
lieve that the interpretation of inkblots enables one to trace personality
profiles, whereas actually they only serve to detect psychotic conditions.)
Methodology can only be ignored at serious risk. But, of course, it is no
substitute for imaginative theorizing-an operation that abides by no
known methodological rules.
This concludes our study of general issues. Our next task will be to
examine the three main strands of modern psychology: mentalism (inclu-
sive of information-processing psychology), behaviorism (classical and
neo-), and psychobiology (or biopsychology).
III

Brainless Psychology
CHAPTER 5

Mentalism

We shall call mentalism the collection of psychological schools that treat


mental processes separately from behavioral and neural ones. The strat-
egy is often underpinned by some immaterialist philosophy of mind. (Re-
call sections 1.1 and 1.2.) Any such philosophy postulates that mind is
immaterial, though possibly linked to matter, so that knowledge of the
latter is irrelevant to psychology.
In some form or other, with or without explicit commitment to the
thesis of the immateriality of mind, mentalism was the dominant school of
psychological thought until the beginning of this century. After a long
exile from academic psychology, it has made a vigorous comeback in the
field of cognitive psychology in recent years. In fact, information-process-
ing psychology is the latest version of mentalism.
The strength of mentalism can be explained as follows. First, it is
commonsensical, at least to anyone innocent of physiological psychology.
Second, it enjoys the support of most theologies and philosophies. Third,
but not least, mentalist psychologists have studied real facts, namely
those of subjective experience, which many antimentalists have either
denied or held to be inaccessible to the scientific method.
We shall take it for granted that mentalist psychology has made impor-
tant contributions to our knowledge of ourselves-though none to animal
psychology. It is not our task to examine those contributions. Instead, we
shall tackle some of the philosophical and methodological problems raised
by mentalism.
Mentalism is most consistent, and at the same time most vulnerable,
when it espouses explicitly an immaterialist philosophy of mind. So far
the immaterialist views of mind have gone through three phases. In the
earliest, the mind was taken to be not only immaterial but also immortal
and a divine gift, and it was called variously psyche, anima, or soul. In the
second phase the soul was secularized, but still conceived of as an imma-
terial thing. In the third and most recent phase, the mind is conceived of
as a collection of programs. But in all three cases the mind was or is
regarded as having an existence separate from the body, so that a knowl-
90 5. Mentalism

edge of the latter was held to be irrelevant to any efforts to account for the
former.
We shall study some philosophical and methodological aspects of the
most interesting versions of mentalism: classical psychology, the Gestalt
school, and information-processing psychology. Finally we shall take a
cursory look at pop psychology.

5.1 Subjective Experience


Feeling hot and being in a merry mood, hearing and listening, remember-
ing and expecting, imagining and experiencing illusions, inferring and
planning, and many others are subjective experiences. These constitute,
of course, the subject of study of mentalist psychology.
The mere existence of subjective experience raises two main philosoph-
ical problems: one ontological, the other methodological. The first is this:
Are they real and, if so, do they happen in the mind or somewhere else?
The methodological problem: Is it possible to study subjectivity in an
objective manner, the way science demands?
Subjective phenomena are sometimes regarded as unreal, just because
they are not immediately accessible to subjects other than those who
experience them. They are reputed to be "only in the mind." Some ofthe
most striking among such experiences would seem to confirm this view:
optical illusions, hallucinations, and afterimages. In all these cases the
subject sees something that is not out there. For example, in the so-called
wall-paper effect, one sees a physically nonexisting contour: See Figure
5.1. Experiences of this kind suggest that mind is immaterial-until one
takes a closer look at them.
There is no doubt that the contour in Figure 5.1 is in here, not out there.
It is then real, though not autonomously so; far from existing in the
external world of the subject, it exists only in his or her inner world. This
poses a problem only if "reality" is defined as the totality of whatever
exists independently of any subjects. But this definition leaves us, sub-

FIGURE 5.1. The wallpaper effect. After staring for a few moments at the drawing
one sees a triangle.
5.1. Subjective Experience 91

jects, out of reality-not a comforting thought. And it makes subjective


experience unreal, and therefore a subject for art rather than science.
Since we find these consequences unpalatable we had better give up the
equation of "reality" and "external world": We must frame a definition
of "reality" that includes the inner world as well.
There are three ways in which the world of subjective experience can
be included in the world, or reality. One is to make the external world part
of the inner world, that is, to adopt some form of subjectivism (e.g.,
solipsism). A second way is to conceive of the inner world as immaterial,
and add it to the external world; this is a dualistic solution. The third way
is to equate "reality" with "materiality," and to construe the inner world
as a collection of processes in an organism; this is the materialist solution.
The first solution (subjectivism) is unacceptable to scientists because it
renders science impossible. The second solution makes physics, chemis-
try, and biology possible, but forces us to view mind as something unnatu-
ral, hence psychology as a nonnatural science separate from the other
sciences. Let us then examine the third solution.
According to the third view, the triangle in Figure 5.1 does not exist in
the world external to me but it is real because it is a process in my own
brain. And the latter can become an object of study for someone else,
hence it exists in the world external to him. Note that there is no such
thing as the external world: Every external world is a world external to
some subject. The world (or reality) may be regarded as the (physical)
sum of all the external worlds-with the proviso that the world existed
long before any cognitive subjects emerged. See Figure 5.2.
An equivalent way of formulating the same idea is this. The triangle in
Figure 5.1 is a biological, in particular psychological, object, not a physi-
cal one. This thesis squares with emergent materialism, which recognizes
several levels of reality or materiality-not degrees of existence, though.
(By the same token it is incompatible with physicalism or eliminative
materialism, which equates "real" with "physicaL") What we have just
said about illusions can be generalized to all subjective experiences: They
are perfectly real-even though they do not occur in the world external to
the subject who experiences them-because they are processes in some
real brain.

w,

FIGURE 5.2. The world (W) as the (physical) sum of the external worlds WI and W2
of subjects SI and ,\'2 respectively.
92 5. Mentalism

The third, or emergent materialist, view clears the way to a fully scien-
tific approach to subjectivity. In fact it shows that an account of subjec-
tive experience need not be itself subjective and therefore untestable
with the help of scientific methods. Thus, when explaining someone's
doing B because he or she wants something, we refer to a complex
process in a real animal, which starts with the subjective or mental
experience of wanting that something, and ends up with performing
behavior B.
Moreover, at least in principle we can test our account by monitoring
physiological or behavioral indicators of such neural and muscular activ-
ity. For example, we can conjecture that an animal is about to perform a
voluntary movement the moment the readiness potential in the parietal,
precentral, and vertex areas starts to exceed its normal value. And we can
affirm that a human subject is dreaming if the subject is asleep and his or
her eyes move in ajerky way or the individual's EEG shows beta waves.
In short, the identity hypothesis makes it possible to study subjectivity in
an objective manner.
Our view is in apparent contradiction with the realist philosophy of
physics, which demands that the cognitive subject be kept out of the
picture because, by definition, the task of physics is to study physical
things, not thinking ones. (See Bunge, 1973b, 1985b.) Actually there is
complementation rather than contradiction. Indeed, psychologists are
supposed to study animal subjects as objects (i.e., things in the world
external to themselves). For example, they will study personal prefer-
ences and subjective values (utilities) in an objective manner. The fact
that subjects a and b assign different values to item c is as objective as
rain. Differences in valuation only prove that not all valuation is univer-
sal. It does not prove that the study of valuation is necessarily subjective.
Although every scientific finding is the work of a person, or a team of
persons, if scientific it is impersonal in the sense that it can be reproduced
by anyone who bothers to master the necessary knowledge and adopt the
scientific attitude.
Although the identity hypothesis suggests the most fully scientific ap-
proach to subjectivity, we owe many important advances to mentalist
psychologists. Thus Fechner, Wundt, Kohler, at times Piaget, and many
other distinguished psychologists were dualists: To them mental phenom-
ena happen in the mind, much as light can propagate in a vacuum. This
belief makes it possible to study mind in its own right, independently from
any explicit reference to its neural "basis," "substrate," or "corre-
late"-though of course not independently from the minding animal.
True, but the mentalist strategy is crippling, for it isolates psychology
from biology and therefore overlooks the intimate couplings between
mental processes and other bodily processes.
Quite aside from the restricted fertility of psychophysical dualism, is it
true or false? Does the hypothesis of the immateriality of the mental have
5.1. Subjective Experience 93

any empirical support, and is it consistent with our background scientific


knowledge as well as with the scientific world view? Let us see.
At first blush there is plenty of empirical support forthe hypothesis that
mind is immaterial. Let us recall just two hackneyed pieces of evidence:
voluntary movement and the fact that we can think quite well without
having a clue as to what neural systems do the thinking, much less how
they do it. True, I can move my little finger at will: I first will it, then do it.
But this does not prove dualism right, for psychobiology has a straight-
forward (though so far only sketchy) explanation for this fact: Some-
times the neural systems that do the thinking activate those that
control bodily movement. Everything remains in the family of neural
systems.
As for the objection that we can think without becoming aware of any
physiological processes in our brains, this is true but it proves nothing.
We also live without being aware of the processes of protein synthesis,
combustion, and cell division that occur uninterruptedly in our bodies. In
fact, all of the routine bodily functions go on without our normally being
aware of them. Body awareness occurs, if at all, when something goes
wrong with the routine. And when the dashboard does signal trouble
(e.g., when we feel pain, thirst, or tiredness) we explain it as a brain event
triggered by another bodily event, such as, for example, a water deficit.
It is pointless to look for experimental evidence in favor of the immate-
riality of mind, because all experiments bear on material things, such as
light rays or brains. In fact, by definition every experiment involves the
controlled application of a stimulus to a concrete thing, and the monitor-
ing of the corresponding response. (Recall section 4.4.) If the stimulus
cannot be controlled within bounds, it is no part of a scientific experi-
ment. (For this reason parapsychological experiments are not quite exper-
imental: See section 5.5.) In short, the hypothesis of the immateriality of
mind cannot be subjected to experimental tests. This should suffice to
eliminate it from science. But there are further reasons as well.
A first additional reason is that the hypothesis is quite unnecessary.
Indeed, the identity hypothesis accounts for all that dualism "explains"
and more, and it does so in a scientific fashion. Second, dualism renders
physiological psychology pointless; hence, if adopted, we would be de-
prived of a valuable body of knowledge. Third, the hypothesis that there
are changes, namely mental processes, that are not changes in concrete
things, is just as out of tune with the scientific world view as the theses
that there can be smiles without faces, life above and beyond living things,
or politics without people. Fourth, psychological dualism is not a solution
to the mind-body problem, but a statement of it. A solution to the prob-
lem can only be a hypothesis relating the two terms (e.g., equating mental
processes with brain processes of a special kind). The identity hypothesis
is just such a solution. (See further arguments against dualism in Bunge,
1980; Hebb, 1980, 1982; Pavlov, 1955; and Smith Churchland, 1986.)
94 5. Mentalism

Sometimes dualism, beaten in the ontological field, takes refuge in


methodology by asserting that there will always be imponderables (i.e.,
unmeasurable qualities), that must remain the preserve of subjectivity and
conceptual sloppiness. An apparently obvious example would be the
quality of musical performance. However, it has become possible to rank
violinists on the basis of measurements performed in a laboratory, and in
a way that matches the sUbjective or intuitive judgment of musical ex-
perts. This is done by investigating the feedback loop existing between the
performer and his or her instrument-that is, the way the performer
regulates the tension of the strings, which determines the pitch (von Be-
kesy, 1968b). To the imaginative scientist unmeasured qualities are chal-
lenges: Some of today's imponderabilia are bound to be tomorrow's pon-
derabilia, and some of today's qualitative theories are bound to be
replaced by quantitative models.

5.2 Classical Psychology


Mainstream psychology between Descartes and World War I may be
called classical psychology. It was characterized by the following fea-
tures: (a) it espoused more or less explicitly the thesis of the immateriality
of mind, and consequently (b) it was nonphysiological, and (c) it focused
on mental events in humans; (d) it split the mind into separate faculties,
and (e) it made much of introspection.
Some of the big names in classical psychology were Aristotle (faculties
of the mind), Descartes (interactionist dualism), Locke (empiricism),
Fechner (psychophysics), Wundt (trained introspection), Ebbinghaus
(nonsense syllables), and Thorndike (animal learning). On the other hand,
the work of Flourens, Broca, Wernicke, Helmholtz, Sechenov, and
Pavlov belongs squarely in physiological psychology. At the time, this
was regarded as part of physiology and irrelevant, or even inimical, to
psychology-and occasionally as subversive. (Sechenov's Reflexes of the
Brain (1863), was censored for undermining public morals: see Boring,
1950.)
The first three features of classical psychology noted earlier go hand in
hand: If mind is immaterial, and possibly a divine gift, then there is no
point in doing physiology or in studying animals. (Classical psychology
took up the study of animal behavior only toward its end.) Only evolution-
ary psychology and the naturalistic (in particular materialist) outlook that
began to prevail in the scientific community from about 1850 on stimu-
lated animal studies, by suggesting that some animals might experience
mental processes because, after all, evolution is continuous in some re-
gards. We have dealt already with the dogma of the immateriality of mind
and its negative effects (section 5.1). Let us now examine the last two
features of classical psychology listed previously: the splitting into facul-
ties, and introspection.
5.2. Classical Psychology 95

Classical psychology postulated that the mind is composed of distinct


mental faculties (capacities, "powers," or "mental organs"), such as
memory, perception, intelligence, volition, and language. This is the weak
version of faculty psychology. The strong version adds the postulate that
each faculty functions separately from the others (e.g., that the language
and the number faculties are mutually independent). The methodological
consequence of the strong doctrine is clear: Each mental faculty can and
should be studied independently from the others. Shorter: Divide and
conquer. Factor analysis implemented this methodological rule and its
creator, C. Spearman, was an eloquent advocate of faculty psychology
even when it was generally thought to be far gone.
By the turn of the century the weak version of faculty psychology had
fallen into utter disrepute, mainly for failing to explain mental life. It was
accused of labeling mental processes instead of accounting for them. (In
fact, to say of someone that she thinks because she possesses an intellect
is not very illuminating. See section 13.3 for pseudoexplanation in psy-
chology.) And the strong version of faculty psychology was harshly criti-
cized by the Gestalt school, which rejected the parcellation of the mind
and, in particular, the detachment of perception from thinking.
Faculty psychology was thought to be dead everywhere except in the
field of mental testing, until Chomsky and his followers revived it in the
late 1950s. In particular, Chomsky (e.g., 1975) characterized linguistics
as the branch of human psychology that studies the "language facul-
ty," an allegedly independent capacity or mental organ that "creates"
particular (nonuniversal) grammars, which in turn "generate" sen-
tences. (Notice the reification of properties, such as mental capacities,
as well as of constructs, such as grammars. See Bunge, 1984, for
criticisms. )
However, the most complete, elaborate, and interesting statement of
the strong version of faculty psychology is due to Fodor (1983). His
theory of the modularity of mind has been well received by cognitive
scientists of the information-processing (or computerized) persuasion.
This is understandable because it renders explicit and systematizes some
of their pet assumptions, notably the thesis of the autonomy of cogni-
tion-which entails that of the autonomy of cognitive psychology. (More
on this school in sections 5.4. and 9.4.)
Fodor has revived and worked out the classical hypothesis that mental
life is the exercise of a small band of distinct and mutually independent
faculties, which he calls 'modules.' Each module is assumed to operate
("compute") in a fixed manner peculiar to it. (A computation is defined as
a certain transformation of representations, but the latter are left unde-
fined.) This modus operandi is supposed to be autonomous in two ways:
independent of the remaining modules as well as of the inputs it receives.
Moreover, such "mechanism" is conjectured to be constant from birth,
that is, genetically determined and hard-wired. The mind is thus likened
96 5. Mentalism

to a toy built with ready-made and essentially unalterable Meccano or


Lego pieces-or silicon chips.
There are a number of problems with this theory. For one thing, it is
extremely imprecise for lack of definitions. (Fodor himself (1983, p. 37)
tells us that he is not in the business of defining his terms.) In particular, it
is not quite clear whether a module is a neural system, a psychological
process, or both. However, judging from Fodor's earlier work (e.g., 1975
and 1981) one may surmise that his modules are immaterial, although they
can be "embodied" now in the flesh, now in the machine. Second, al-
though occasionally he pays lip service to neuroscience, he makes no use
of it. Moreover, he assures us that "there is nothing to know about the
neurophysiology of thought" (1983, p. 119)-which implies that thought
is not a brain process, and that the whole of physiological cognitive psy-
chology is wrongheaded (or should we say 'wrong-minded'?). Third, be-
cause the cognitive modules are supposed to be rigid (nonplastic), the
theory is at odds with the mounting evidence in favor of neural plasticity,
as well as with the whole of developmental and evolutionary psychology.
Fourth, because the modules are assumed to operate independently of the
stimuli they receive, they are not modified by experience; the mind can
acquire or lose information but it cannot learn or unlearn to do so. A
fortiori, the innateness (or hard-wiring) hypothesis renders creativity illu-
sory and forces upon us the gloomy prospect that human knowledge is
intrinsically and radically bounded (Fodor, 1983, p. 120). Fifth, virtually
the only data Fodor cites in support of his theory come from nonbiological
studies of language-and even so he admits, pace Chomsky, that speech
and perception act in unison. So, at best his is a theory about the exercise
(though not the acquisition or the deterioration) of language skills. Sixth,
because the nervous system plays no essential role in his theory, the latter
does not involve any mechanisms proper (i.e., no processes in a material
thing). Consequently, it explains nothing. (For explanation, genuine and
bogus, see section 13.3.) Seventh, the theory includes no law statements;
in particular, it specifies no "computation" laws. Therefore it can make
no predictions. Eighth, because the theory is disjoint from neuroscience
and uninterested in social science, it is of no possible use in clinical
psychology or in psychiatry. The strong theory of the modular mind is, in
conclusion, theoretically shaky and shallow, empirically weak, and prac-
tically barren.
There is, though, some empirical support for the weak version of the
modularity doctrine. In fact, some cognitive capacities, notably speech
and spatial representation, are fairly localized in the brain-not in the
immaterial mind. Hence a malfunction of one of the corresponding (neu-
ral) systems may not affect the others. (However, language is not local-
ized in a single compact center. Rather, it "resides" in a number of brain
components-which is not surprising given that it has syntactic, seman-
tic, and pragmatic aspects.) On the other hand, memory and learning do
5.2. Classical Psychology 97

not seem to be localized in any particulary subsystem of the brain. We


shall return to the matter of functional localization in section 7.5.
But the strong version of modularity is certainly false. In fact, any
moderately complex cognitive task involves a number of intertwined
"faculties" or, rather, neural systems. Think of vision, which involves
the saccadic movement of the eyeballs, as well as previous learning and
expectations. Or think of the production of a drawing or a sentence: both
processes involve not only cognitive mechanisms but sensorimotor ones
as well.
We submit that the various mental "faculties" are intimately linked to
each other, and we explain this coupling in terms of the neural connec-
tions existing between their corresponding neural systems. Whenever a
connection ofthis type is severed (e.g., surgically), a severe pathological
condition known as disconnection syndrome results. For example, the
Kliiver-Bucy syndrome, which involves loss of ability to make correct
choices under visual control, results in monkeys from the disconnec-
tion between the visual and the limbic systems (Geschwind, 1974). In
sum, the mind is not fully modular because every mental process is a
brain process, and the various subsystems of the brain are normally
interconnected. A fully modular mind, were it to exist, would be disas-
trously sick. There is localization, to be sure, but there is also inter-
dependence.
The methodological consequence is clear. The various mental capaci-
ties cannot be adequately studied in isolation from one another or from
motor behavior. It is not that the mind (or the brain) is a homogeneous
("equipotential") block. Instead, it is a system of interacting compo-
nents, which can and must be distinguished but not put asunder. The
correct methodological slogan is not "Divide and conquer" but "Distin-
guish without detaching: analyze and synthesize." For example, in order
to understand how people learn to play the piano it is not enough to appeal
to a "musical faculty": One must investigate auditory imagery, hearing,
score reading, fingering, and their mutual relations-not to speak of mem-
ory, attention, and reflexes.
Faculty psychology left us with a laundry list of mental capacities. It
made no effort to explain their emergence and submergence, or even to
classify them. The absence of a classification is not very important
in basic psychological research, but it is of practical importance. The
clinical psychologist and the educational psychologist, the psychiatrist
and the neuropsychologist are in need of a reasonable classification
of mental abilities and dysfunctions, if only to direct their patients to
the appropriate wards. A mere list of mental faculties or modules
won't do.
We propose to group the behavioral and mental processes into three
large families: motor, affective, and cognitive. The motor activities are
the specific functions of the neuromuscular system; the affective pro-
98 5. Mentalism

cesses (pleasure, pain, love, hatred, anger, anxiety, etc.) would occur
mainly in the limbic system; and the cognitive processes (perception,
imagery, thought, etc.) would be the specific functions of the cortico-
thalamic and corti co-limbic systems.
Our grouping is not a partition, because there are mixed processes
(e.g., sensorimotor ones). Nor is it a classification. A classification proper
is a conceptual hierarchy in which classes are ordered by the relation of
class inclusion: See Figure 5.3. (See Bunge, I983a for the rules of classifi-
cation.) A possible strategy for building a classification of behavioral and
mental processes is the following, in three steps. First, group abilities and
dysfunctions by their gross manifestations (e.g., aphasia). Second, ferret
out the neural' 'correlates" of such abilities or dysfunctions and, if neces-
sary, regroup them in the light of the neurophysiological evidence. A
result will be a species such as semantic aphasia. The gross manifestation
would be the genus, and the neural "correlate" the species. Third, try to
find some new principle capable of linking together the various families.
Look into the possibility of finding such principle in evolutionary consid-
erations. For example, in the beginning there may have been a single
sensorimotor faculty that evolved eventually into distinct capabilities,
such as chemical sensing-motility (as in bacteria), light sensitivity-motility
(as in Euglena viridis), and so forth. Much later, some senses became
separable from motility. And much, much later some faculties, such as
visual imagery, evolved from sensory ones.
So much for mental faculties. We shall come back to them, construed
as specific functions of neural systems, in chapters 8 and 9. Let us now
address the question of introspection. As we saw in section 4.1, introspec-
tion is a procedure but not a methodological one. Whatever it is, intro-
spection is not what "the little man in the head" sees or says. (The
homuculus hypothesis leads to an infinite regress: The homunculus needs
another of its kind in order to see what it sees, and so on.) We submit that
what passes for introspection is nothing but consciousness, and we ex-
plain the latter as the monitoring of the activity of one neural system by
another. (For details see chapter 11.)

FAMILY

Z ....Jz
0 «Q
wf= UCIl
~~ GENERA -::J
(!J....J
f=~ Ou
c.. ....Jz
~

SPECIES

FIGURE 5.3. A program for classing behavioral and mental abilities and dysfunc-
tions.
5.2. Classical Psychology 99

There is no denying the importance of introspection, particularly in


the task of "mapping the mind on to the brain," that is, in bridging the
chasm between subjective experience and neurophysiology. Therefore
Watson (1913) and his followers were wrong in rejecting self-obser-
vation. But they were right in claiming that introspection is unreliable
(e.g., because of the tricks memory plays, and of interference with
the mental process that is being described by the subject). Besides, self-
observation is limited to humans above a certain age, and it reports
only on conscious events, which constitute just the tip of the mental
iceberg.
The distortions and limitations of introspection have recently been
highlighted by observations on commisurotomy patients, who are unable
to tell us correctly what happens to them (in some parts of their brains).
For example, they may be unable to name colors, yet be perfectly capable
of sorting objects by color. Due to a disconnection between the speech
area and other brain areas, they have lost their ability to report on some of
their subjective experiences.
The data of introspection being incomplete and dubious, they have to
be checked and supplemented with behavioral and physiological ones.
Besides, self-observation, even when correct, yields only a description.
Only a theory, suitably enriched with data, can explain anything. In sum,
we reject not introspection but introspectivism, that is, the thesis of most
classical psychologists, that self-observation gives us a direct, complete,
and incontrovertible knowledge of mind. (See Bindra, 1976, pp. 177-178.)
We conclude with an evaluation of the legacy of classical psychology.
This legacy is ambivalent. On the one hand, it formed the initial body of
problems of psychology, which it sometimes investigated scientifically.
This investigation yielded an important body of data concerning subjec-
tive experience. These data cannot be ignored; they must be checked and,
when found true, explained.
On the other hand classical psychology was too narrow, in particular
for restricting itself to high-level processes in humans. It was theoretically
indigent: Its only important hypothesis was that of the association of
mental "atoms." Although it would be foolhardy to deny the fact of
association, it must be admitted that associationism explains neither the
emergence of the "atoms" themselves nor originality. Classical psychol-
ogy was also methodologically shabby for its excessive reliance on intro-
spection. Because of its lack of interest in the nervous system, it remained
rather isolated from the bulk of natural science, and it consorted openly
with philosophical psychology-in fact it was sometimes difficult to dis-
tinguish from the latter.
Yet, despite its severe shortcomings, classical psychology was the his-
torical root of subsequent developments, which in some respects-partic-
ularly with regard to the experimental method and animal studies-were a
continuation of the last phase of classical psychology.
100 5. Mentalism

5.3 Gestalt Psychology

The Gestalt school of Wertheimer, Koffka (1935), and Kohler (1929) flour-
ished in Europe in the 1920s while behaviorism was taking over in the
United States. The two schools were at odds in every regard except for
their trust in experiment and their mistrust of reason. Whereas the Gestalt
school focused on subjective experience, made ample use of introspec-
tion, and embraced psychophysical dualism, behaviorism disregarded the
mind, abhorred introspection, and denied the existence of the mind-body
problem. Whereas Gestalt psychology was holistic, behaviorism was ana-
lytic; where the former was inclined to theorize, the latter saw no reason
to do so; and whereas the former was strongly influenced by holistic
metaphysics and intuitionistic epistemology, behaviorism was thoroughly
positivistic.
The password of the Gestalt school was of course Gestalt. Whatever it
found was pronounced to be a Gestalt or a law of Gestalten; and whatever
it attempted to explain, it did so in terms of Gestalten held to be unanalyz-
able. Regrettably, not being given to conceptual analysis, the members of
the school often employed the term Gestalt in three very different senses:
whole or system (Ganzheit) , structure or configuration (Struktur) , and
emergent or systemic property (Gestalt-quaIWit). Not surprisingly, a co-
lossal conceptual muddle resulted. Understandably, many philosophers
hailed Gestalt psychology not so much for its genuine findings as for its
obscurity-whereas others rejected it out of hand for the same reason.
Let us attempt to disentangle the concepts designated by the ambiguous
term 'Gestalt'.
A whole, totality, or system, is a complex object the components of
which are coupled, as a consequence of which the object has a definite
structure. A whole differs then from an amorphous aggregate. We must
distinguish systems of three different types: conceptual, material, and
functional. A conceptual system, such as a hypothetico-deductive one, is
made of constructs (e.g., propositions) held together by logical relations
(e.g., that of deducibility). A material system, such as a brain, is com-
posed of material things (e.g., cells) held together by links or bonds of
some kind (physical, chemical, biological, or social). A functional system
is composed of properties that are mutually related by laws.
At this point we are not interested in conceptual systems, such as
theories, which are studied by logic, mathematics, and epistemology. As
for the concept of a material or concrete system, we have elucidated it in
section 3.2. There we defined the structure of such a system as the collec-
tion of all the relations (in particular the bonds or couplings) among its
components, plus the relations among the latter and the relevant environ-
mental items. (A coupling, link, or bond is a relation that makes some
difference to the relata. Causal relations are bonds, spatiotemporal rela-
5.3. Gestalt Psychology 101

tions are nonbonding: See Bunge, 1977a.) It follows then that there are no
structures in themselves, separate from things; every structure is a prop-
erty of a (complex) thing. In other words, 'structure' is an adjective, not a
noun. Consequently it is not synonymous with 'whole,' 'totality,' or 'sys-
tem.'
Let us now clarify the notion of a functional system, which is central to
Luria's thought (Luria, 1973, 1979), and the one that the Gestalt psycholo-
gists, as well as Piaget and Vygotsky, seem to have had in mind when
writing rather obscurely about 'psychological Gestalten (or wholes, or
totalities).' Both Piaget and Vygotsky held that such Gestalten emerge
abruptly in the course of individual development, and in tum give rise to
further wholes. But whereas Piaget saw such emergence as a necessary
outcome of biological maturation, Vygotsky emphasized the action of the
social environment. With hindsight we may find it obvious that the endog-
enous and exogenous factors intertwine. (More on the nature-nurture
issue in section 6.2.) However, our present concern is not with this thesis
but with the concept of a functional system.
We stipulate that a collection of properties is afunctional system if, and
only if, it is a collection of properties of a concrete (material) system such
that, given any member of the collection, there is at least one other
member of it that depends upon the former. A possible formalization of
this concept with the modest resources of the predicate calculus is this. If
IP is the collection of all possible properties of concrete things, then F is a
functional system if, and only if,
F = {P E IP I (3x)(VP)(3Q)
[x is a concrete system /\ Px /\ Q E IP 1\ (Px ~ Qx)]} [5.1]

There are no stray properties. To put it affirmatively: Every property


belongs to some functional system. This is a consequence of two general
philosophical postulates: (a) that every material thing is either a system or
a component of some system, and (b) that every concrete thing satisfies
some laws, which are relations among properties. If we now add the
assumptions that the brain is a system, and that mental faculties are brain
properties, it follows that the collection of mental faculties forms a func-
tional system. The Gestalt psychologists held the same view, though for
obscure reasons.
So much for wholes and their structures. Let us now tackle the third
signification of 'Gestalt,' which we have identified as the concept of an
emergent property. The Gestalt psychologists launched the popular slo-
gan, The whole is more than the sum of its parts. But they did not bother
to elucidate the tricky words 'more' and 'sum': they did not specify what
the extra was, nor in what way it exceeded the 'sum.' However, we
submit that the slogan makes sense in the following translation: "Every
system has some global or systemic properties that are emergent relative
to those of its components, that is, that the latter lack." This ontological
102 5. Mentalism

thesis, which embraces a definition of "emergence," is not restricted to


biology: Emergence occurs on all levels of reality. (See Bunge, 1979a.)
Thus, an atom has properties, such as a radiation spectrum and chemical
reactivity, that its elementary components do not possess; likewise a
society has properties, such as a level of economic development and a
form of government, that the individuals composing it could not possibly
possess. On this point we agree then with the Gestalt psychologists:
Emergence is for real. Where we must disagree is in the mode of under-
standing emergence.
The Gestalt psychologists held that emergence is a sort of atomic fact
that cannot be analyzed, hence understood-that it had to be admitted as
an ultimate. In particular, they denied most emphatically that the emer-
gent properties of a system could be explained in terms of its composition
and the interactions among its components. This was their negative epis-
temological thesis about emergence. This thesis had been refuted even
before the first Gestalt psychologists were born. In fact, basic and applied
scientists had known for a long time at least two major and conspicuous
emergence mechanisms: assembly (natural or artificial) and evolution (bi-
ological or cultural).
In particular, the chemist can explain the properties of a molecule in
terms of those of its precursors (atoms or molecules) and their mode of
binding. And the biologist can explain the properties of the members of a
species in terms of those of their ancestors, as well as in terms of internal
processes (such as embryological development) and environmental ones
(such as natural selection). In general, a knowledge of parts and their
interactions suffices to understand what may be called developmental
emergence. (See e.g., Piaget, 1965, p. 28.) Likewise, a knowledge of
ancestors and their inner constitution, as well as their environmental
circumstances, suffices to explain what may be called evolutionary emer-
gence. In short, there is emergence and it can be explained, at least in
principle. (For more see section 13.3, and Bunge, 1979a, 1981, 1983b,
1985a.)
How has Gestalt psychology fared? Whereas some of its findings have
been partially confirmed, others have been refuted by ulterior experimen-
tal developments, and hardly any of its explanations have survived. For
example, one of the axioms of the school was that perception is an instan-
taneous event. We now know that every perception is a process that takes
at least 100 msec. Another axiom was that perception is a unitary act-
that we perceive things as wholes and analyze them, if at all, later on. The
evidence for this hypothesis (as for that of instantaneity) was exclusively
introspective or phenomenological. Recent experiments involving the
measurement of reaction times to the presentation of visual stimuli sug-
gest that, whereas some objects are indeed perceived as wholes, others
are synthesized from their features. For example: Gestalt-like effects oc-
cur in the perception of apparent motion (Ramachandran & Anstis, 1983);
5.3. Gestalt Psychology 103

o _ NO
~
FIGURE 5.4. Examples of emergence through assembly.

human subjects detect triangles as units or perceptual primitives, but


arrows as complexes (Treisman & Paterson, 1984). Whereas the former
are spotted automatically and preattentively, the latter require attention
and a serial search-which presumably is identical with the sequential
activation of neural assemblies. These and similar results may be taken as
a partial confirmation (hence a partial refutation) of the Gestalt axiom that
all visual stimuli are perceived as wholes.
However, such wholeness, when it exists, is actually preceded by an
unconscious analysis ofthe stimuli into separate features. See Figure 5.4.
In fact, recent psychological experiments show that features come first in
perception, and become integrated at a later stage provided the subject
focuses his or her attention on the object (Treisman & Gelade, 1980). This
finding has a neurophysiological explanation. In fact, it has been known
since the work of the Nobel laureates Hubel and Wiesel (1962) that the
visual system is an analyzer that processes the incoming signals in a
sequential fashion that involves, successively, cells of different types,
some of which synthesize what the earlier ones in the process had ana-
lyzed. The same holds for the auditory system.
In fact, the two previously mentioned sensory systems contain feature
detectors that react only to particular features of the stimuli, such as
edges and colors, and even lines of particular inclinations. However, in
addition to such sensory analyzers, the mammalian brain contains neu-
ral units that put together the signals put out by the various analyzers;
they synthesize them into the perceptual wholes that we perceive. See

\/1;\
• STIMULUS OBJECT

Horizontal lines V'rt;"I1;,,"


'''"',, J ,,,,ct,,

o
Synthesizer

FIGURE 5.5. Perception of a whole as such is preceded by analysis performed by


feature detectors.
104 5. Mentalism

Figure 5.5. So, even when a figure-such as an X in the midst of an array


of Os- "pops out" at the subject, who grasps it automatically (preatten-
tively) in one go, this wholeness is the outcome of a complex analytic
process. Although the final synthesis would have warmed the hearts of
the Gestalt workers, the prior analysis would have disappointed them as
well as their philosophical partners.
The associationists had attempted to explain complex mental phenom-
ena in terms of "mental atoms" and their associations (Mill's "mental
chemistry"). In a similar vein, the reflexologists had tried to decompose
the higher brain functions into elementary reflexes. Neither succeded in
their attempt. The Gestalt psychologists took these failures to prove the
failure of analysis altogether and, in general, of reason. Consequently
they adopted a holistic metaphysics and an intuitionistic epistemology.
The adoption of these philosophies prevented them from analyzing, hence
understanding, the very Gestalten they had found.
Let us finally take a look at the particular version of psychophysical
dualism espoused by the Gestalt school, namely psychophysical parallel-
ism. This is the hypothesis that every mental state (or event) is correlated
with a synchronic brain state (or event). Or, as the Gestalt psychologists
put it: The set of mental events is isomorphic to that of brain states.
(Actually there is a morphism but not an isomorphism because some brain
states, perhaps most, have no mental "correlates.") Note that it is not
assumed that mental events cause, or are caused by, brain events, let
alone that mental events are identical with brain events of some kind;
instead, the synchrony of heterogeneous events is postulated.
Of all the views concerning the mind-body relation, parallelism is the
safest, because it accounts for all possible data. In fact, it cannot be
empirically refuted because there are no mental states without brain
"counterparts." Because the hypothesis is insensitive to empirical find-
ings, we must turn elsewhere for its evaluation. We must ask whether it
has any heuristic, explanatory, or predictive power, and whether it is
consistent with biology and with the scientific worldview.
Clearly, parallelism has some heuristic and predictive power, for it
directs psychologists to search for the neural "correlates" of mental pro-
cesses. Still, the identity hypothesis does this and more, for it explains the
mental as physiological. On the other hand, parallelism explains nothing.
In particular, it does not even attempt to explain mental development
(ontogeny) or mental evolution (phylogeny), for it takes the mind for
granted-whence it is at odds with biology and it invites postulating a
supernatural agency. Nor does parallelism harmonize with the scientific
worldview. Indeed, the latter does not countenance any events other than
changes in concrete things. To speak of mind-body parallelism is like
speaking of movement-body parellelism. (Recall section 5.1.)
In conclusion, Gestalt psychology must be credited with some impor-
tant experimental findings, such as the apparently instantaneous percep-
5.4. Information-Processing Psychology 105

tion of (some) forms, the existence of "good forms" (such as a C, in


contrast to an 0), the Gestalt switches (such as the one that occurs auto-
matically every half a minute or so when staring at a Necker cube), and
the "aha experience" in problem solving. But Gestalt psychology was
unwilling to explain any of its own findings. In fact, it resorted to the word
Gestalt as a sort of magic device to block any attempt to explain the
phenomena (Petermann, 1932). This anti scientific attitude must be blamed
on the obscurantist philosophy adopted by the Gestalt psychology.

5.4 Information-Processing Psychology


Our third example of a mentalist school is information-processing psy-
chology. The birth of this movement in the 1960s can be explained by the
exhaustion of classical behaviorism, the renewed interest in cognitive
processes and in language, and the burgeoning of information engineer-
ing. The new school tackles the vast but neglected problematics of cogni-
tive psychology, from perception to inference, with the help of only a
handful of concepts, mainly those of information and of automaton. It
draws strong inspiration from computer science and artificial intelligence,
and on the other hand it makes no use of physiological or social psychol-
ogy, and it does not pay much attention to experiment. Anyone conver-
sant with folk psychology and the jargon of information engineering can
hope to enter the new field. Many have entered it attracted not only by its
novelty and basic simplicity, but also by its practical promise. It is very
hard to resist the charm of that peculiar combination of folk psychology
with mathematics and technology, particularly when so skillfully adver-
tised as the final solution to all the problems of cognitive psychology-
and a lucrative one to boot.
Information-processing psychology comes in two strengths. In its weak
version it asserts only that all cognitive processes are processes of recep-
tion, transformation, and transmission of information. The strong version
holds that all cognitive processes are computations, or symbol "manipu-
lations," whereby information is processed in accordance with defi-
nite rules (algorithms or programs). Because strong implies weak, if
we find fault with the latter we shall have damaged the former. (Recall
the modus tollens inference pattern: S => W, l W f- l S.) Let us then
start by examining the idea that cognition of any kind is information
processing.
It is an axiom of the school that the mind is a communication system
basically similar to a radio or television set. This being so, Shannon's
statistical-information theory of 1948 must apply to it, and the whole of
theoretical psychology would be nothing but a collection of exercises in
information theory, while the task of experimental psychology would be
restricted to measuring the quantities of information, in bits, entering and
106 5. Mentalism

leaving an organism. In particular, the psychologist would determine the


channel capacity of people for every kind of sensory stimuli, that is, the
greatest amount of information they can give us about a stimulus imping-
ing on them. (How this measurement could be performed on nonhuman
animals was never explained.) As Miller (1956) shows in his delightful
pioneering paper, that capacity is, for humans, only about seven bits. So
far so good. Trouble starts when one has the impudence to ask what
'information' means, for it is not not always used in the sense of statisti-
cal-information theory.
In fact the word 'information' is being used in the contemporary scien-
tific literature in-you guessed it-seven different ways:
information I = meaning (semantic information)
information2 = structure of genetic material (genetic "information")
information3 = signal
information4 = message carried by a pulse-coded signal
information5 = quantity of information carried by a signal in a system
information6 = knowledge
information7 = communication of information6 (knowledge) by social be-
havior (e.g., speech) involving a signal (information3)
All seven concepts seem to be relevant to cognitive psychology-pro-
vided they are not confused. Unfortunately confusion, unwitting or delib-
erate, is frequent. For example MacKay (1969), an early enthusiast of
information-processing psychology, proposed identifying information I
(meaning) with information3 (signal). And neuroscientists often use 'infor-
mation' as synonymous with 'signal' (information3)' In fact, when they
say, for instance, that 'organ A sends information to organ B through
the intermediary of messenger C,' all they mean is that A acts upon B
through the intermediary of some physical or chemical signal C. Given
the multiplicity of concepts designated by the fashionable word 'infor-
mation,' it is mandatory to distinguish them, to indicate their domains
of valid applicability, and to try and find out how they are mutually
related.
In a sense, the weak thesis that cognition is information processing is
trivial, provided 'information' is taken in the third sense, namely to sig-
nify "signal." In fact, the nervous system does pick up, transform, gener-
ate, and deliver signals of various kinds. However, this statement is just
as true, and just as uninformative, as the assertion that the nervous sys-
tem is an energy transfer system. Neither statement describes or explains
anything in detail: Both are global descriptions. The point is to find out
how information (in the various relevant senses) is gained, kept, transmit-
ted, created, and lost by the nervous system. In other words, the problem
is to transform flowcharts into mechanisms. As soon as this endeavor is
undertaken, the marked difference between nervous systems and artificial
communication systems becomes evident.
5.4. Information-Processing Psychology 107

An important difference between an artificial communication system,


such as a telephone network, and a nervous system, is this. Whereas the
former transmits messages that may be distorted by noise along the chan-
nel, the nervous system transmits signals (action potentials) that may
become messages (information4) when processed in the higher centers. In
fact, the signals that propagate along a nerve do not carry any definite
messages. The reason is that the effect of a nerve signal upon its target
depends critically upon the kind and state of the latter-something that
the "sender" could not possibly anticipate. In fact, the final stretch in any
propagation process in the nervous system is a chemical event involving
two items: a neurotransmitter (atom or molecule) released by a vesicle in
the presynaptic neuron, and a receptor in the postsynaptic membrane.
The neurotransmitter binds chemically with the receptor-unless the lat-
ter has been blocked by another, chemically similar, neurotransmitter.
The effect of the neurotransmitter on its target (i.e., the "message deliv-
ered by the messenger"), depends then critically upon the kind of recep-
tor and the state it is in-which is anything but a stationary one. To put it
anthropomorphically, the messenger delivers the message its addressee
wants to hear. For this reason it is impossible to intercept a nerve "mes-
sage" midway by "wiretapping" a nerve with the help of microelec-
trodes. (After all, the neurotransmitters occurring in our exalted brains
are also found in unicellular organisms, where they perform different
functions by combining with different molecules: a good example of the
opportunistic character of biological evolution.)
This important disanalogy between nervous systems and artificial com-
munication systems should suffice to dash the whole of information-
processing psychology, based as it is on the dogma that there are no
important differences between the two kinds of system. But
information-processing psychologists do not usually listen to neuroscien-
tists, so they go undaunted about their business. More on this will follow.
Despite the obvious flaws of information-processing psychology, some
of its box-and-arrow models may perform useful if limited functions.
First, they exhibit graphically and synoptically the results of conceptual
analyses of cognitive processes, such as the transfer from short-term to
long-term memory, the conversion of graphemes into phonemes during
the reading process, or the action of attention on perception. Second,
such flowcharts may suggest the search for neural systems, pathways and
processes; that is, they set psychobiology the task of fleshing out the
boxes as neuron assemblies, and the arrows (or channels) as nerves,
axons, dendrites, synaptic junctions, or what have you. By the same
token, the models suggest looking for the anatomical or physiological
"correlates" of cognitive deficits. When information-processing models
do perform these heuristic and didactic functions, they are useful for
helping pose research problems. They are disvaluable only if taken as
solutions, that is, as goals of research, for they are merely functional:
108 5. Mentalism

They relate disembodied functions, and consequently they explain noth-


ing. In particular, they pay no attention "to questions of why the learning
occurs or what maintains it" (Estes, 1984, p. 624).
Although a refutation of the weak version of information-processing
psychology automatically refutes the strong or computationalist version,
it will be worth our while to examine the latter in its own right, for being
more interesting, influential, and pernicious. We take it that computa-
tionalism boils down to the following axioms. First axiom: The mind is a
computer; that is, all mental processes are computations, or rule-gov-
erned information processes (e.g., the mind computes or "figures out"
what the eye sees and the hand does). Second axiom: Any true theory of
mind is a program, or algorithm, or effective computation procedure. (See
Anderson, 1983; Cohen & Feigenbaum, 1981-82; Dennett, 1978; Fodor,
1975,1981; Haugeland, 1981; Johnson-Laird, 1983; MacKay, 1978; Marr,
1982; Newell, 1982; Newell & Simon, 1963, 1972, 1981; Putnam, 1960,
1975; Pylyshyn, 1978, 1980, 1984; Simon, 1979, 1980; Sloman, 1978a;
Turing, 1950.)
The first axiom is of a substantive nature. It presupposes that all mental
processes are cognitive, or at least that cognitive processes are indepen-
dent of other (mental and non mental) processes-which of course was an
axiom of faculty psychology. And it implies that "our understanding of
the mind will not be further improved by going beyond the level of mental
processes," in particular by studying the nervous system (Johnson-Laird,
1983, p. 9).
The second axiom is of a methodological nature. It presupposes that
theories can be characterized as (not just implemented by) computer pro-
grams, that is, as sets of instructions, rather than as sets of propositions.
And it implies that the only mathematical functions worth considering in
psychology, or at least in cognitive psychology, are the recursive func-
tions, which map nonnegative integers onto nonnegative integers. It also
implies that the only valid explanations in psychology are those formu-
lated as effective (or mechanically performable) procedures.
The first or substantive axiom is open to the following objections: (a) it
leaves out plenty of mental processes, such as motivations, intentions,
expectations, and feelings, that are not obviously describable as computa-
tions or rule-governed operations; (b) it isolates cognitive psychology
from all the other branches of psychology, in particular developmental
and evolutionary psychology-a ghettoization that no psychologist
should welcome; (c) it has been experimentally confirmed only in a few
cases of deductive inference (e.g., Johnson-Laird, 1983)-which is like
testing medicines on healthy people; (d) it has been refuted by experi-
ments on the perception of apparent motion that involve images presented
to viewers at speeds too fast to allow them to think about what they were
seeing (Ramachandran & Anstis, 1986); (e) it has been refuted by all the
studies showing the importance of motivation, expectation, and intuition
5.4. Information-Processing Psychology 109

(in particular shortcuts) in problem solving; (j) it offers no clue as to how


the brain acquires the alleged rules of computation or algorithms-unless
it suggests that the genes are smart enough to "embody" such rules; and
(g) it overlooks the radical difference between natural laws (e.g., those
that brains are supposed to satisfy) and man-made rules (e.g., those com-
puters abide by). In short, the axiom that the mind is (or is like) a com-
puter holds no water.
As for the second or methodological axiom-the program-as-a-theory
prescription-it faces the following objections: (a) it involves a serious
misunderstanding of the word theory, which in logic, mathematics, phys-
ics, and other disciplines, designates a hypothetico-deductive system, not
a sequence of instructions; (b) it involves a frightful impoverishment of
psychology, by depriving it of nonrecursive functions, particularly the
ones studied in analysis, such as the linear, exponential, and sine func-
tions, not to speak of the more complicated ones occurring in theoretical
physics, chemistry, biology-and psychology; (c) it involves eschewing a
critical examination of presuppositions, which are rarely written in a pro-
gram, as a consequence of which "revisions, when indicated, take the
form of ad hoc program tinkering and tuning rather than clear reformulat-
ion of theoretical assumptions" (Erikson & Riess Jones, 1978, p. 72); (d)
the only empirical evidence in favor of the axiom comes from think-aloud
tasks, which force the subject to think serially in a manner remotely
resembling a computer, whereas natural thinking is usually parallel. In
sum, the identification of psychological theories with computer programs
is just as untenable as Chomsky'S identification of grammars with theo-
ries.
There is more. For one thing, computationalist psychology has no ex-
planatory power. Indeed, an explanation is an inference whereby a set of
premises describing some mechanism, and an input to it, imply the propo-
sition(s) describing the output to be explained. Clearly, not being a theory
but a sequence of instructions, a computer program cannot deliver any
explanations. (For details on scientific explanation see section 13.3.) All
the computationalist can produce is an illusion of explanation, produced
by rephrasing psychological descriptions in terms familiar to the com-
puter crowd.
Thus, if a patient loses permanently his long-term memory (i.e., be-
comes a dense amnesic), the computationalist is likely to say that the
subject's retrieval mechanism is out of kilter. However, the neuropsy-
chologist will point out that there can be retrieval without recollection (or
"memoryless memory"). In fact, the amnesic patient can perform certain
tasks without recalling that he knows how to do them. (See e.g., Schacter,
Harbluk, & McLachlan, 1984.) And if the patient loses his long-term
memory and then recovers it, the computationalist may produce an "ex-
planation" to the effect that there is a momentary unspecified "malfunc-
tion" of the "retrieval mechanism," which did not affect "information
110 5. Mentalism

storage. " The illusion that this translation into informationese constitutes
an explanation could have the effect of blocking research into the neural
mechanisms of formation, loss, and recovery of long-term memory.
(More on explanation in section 13.3.)
Furthermore, computationalism involves a confusion between science
and technology, particularly artificial intelligence (AI). The ultimate goal
of AI is not to explain cognition in terms of natural laws-the job of
cognitive psychology-but to design efficient, fast, reliable, and inexpen-
sive machines and programs likely to successfully mimic or replace cer-
tain cognitive processes. The resulting artifact cannot be a good guide to
psychologists because they study animals, which are products of blind,
opportunistic, and tortuous evolution, not intelligent design-whence
such animals are likely to operate in complex, slow, unreliable, and ex-
pensive ways.
Because psychologists wish to understand animals, not machines, they
have little to learn from AI. (On the other hand, whoever wishes to imitate
a human activity must start by becoming acquainted with the latter.) If
they come under the spell of AI, psychologists risk forgetting much: for
example, that animal memory is not a matter of storage and retrieval, for
it is fading and constructive rather than faithful (Bartlett, 1932); that hu-
man thought, far from being always rule-directed, is often erratic and
occasionally original-to the point that it has come up with computers
and robots, whereas no machine is known to have created or studied
humans; and that original problem solving does not proceed according to
rules, whence it cannot be translated into a computer program: Instead,
when faced with a novel task we search and research, we proceed in zig-
zags with the help of scraps of knowledge, visual models, occasional
flashes of intuition, and impelled or blocked by emotions and expecta-
tions. For these reasons, taking the advice of AI fanatics, that empirical
research in psychology should come only after computer programs imitat-
ing mental processes have been devised (Sloman, 1978b), can only cripple
psychology and discredit AI. In fact, that advice has already done consid-
erable harm.
Last, let us dig up the philosophy of mind underlying information-
processing psychology. In the old days mentalism was the enemy of mate-
rialism and, in particular, of mechanistic materialism. Whereas the former
extolled immaterial, immortal, and omnipotent mind, the latter held that
all animals, including humans, are mindless robots, though some more
skillful than others. Knowledge engineering has made it possible for the
former enemies to marry on the basis of the thesis that, far from being a
collection of special brain functions, mind is a collection of programs
detachable from hardware (machine or body). This is of course old psy-
chophysical dualism in fashionable garb.
Indeed, information-processing psychology conceives of mind as an
entity in itself, distinct and separate from the brain, and capable of work-
5.5. Pop Psychology 111

ing autonomously and even looking into itself. Thus Marr (1982, p. 6):
"Modem representational theories conceive of the mind as having access
to systems of internal representations; mental states are characterized by
asserting what the internal representations currently specify, and mental
processes by how such internal representations are obtained and how
they interact." Because information-processing psychology conceives of
mind as self-existing, it explains the mental by the mental (e.g., the
change in an internal representation as a result of the interaction of two
internal representations). The brain is not even mentioned, and the ex-
pression 'interaction of two internal representations' is not defined.
The idea that any given mental process can have different "embodi-
ments" (McCulloch, 1965) or "instantiations" (Pylyshyn, 1984), now in
the flesh, now in the machine, or even in disembodied spirits (Fodor,
1981), goes back to Plato's idealism. Its methodological consequence,
that psychology does not need any neuroscience for being a very special
autonomous discipline (Fodor, 1975), goes back to philosophical or arm-
chair psychology. For all its revolutionary rhetoric, information-process-
ing psychology is then, on the whole, a scientific counterrevolution. Sci-
entists are supposed to push forward, not backward. In particular, the
genuine advances in psychology have progressively estranged it from
mentalism and integrated it with biology, medicine, and social science.
(Forfurther criticisms see Bindra, 1984; Bunge, 1956, 1980, 1985b, 1985e;
Estes, 1984; Hebb, 1980; Paivio, 1975; Weizenbaum, 1976.)

5.5 Pop Psychology


We shall call pop psychology the family of nonscientific beliefs about
behavior and mind that enjoy wide popularity in our time. The best known
members of this family are folk psychology, psychoanalysis, and parapsy-
chology. All these beliefs exhibit or presuppose some version of psy-
chophysical dualism, are conceptually fuzzy, unsupported by experi-
ment, and alien to the body of science, particularly to biology.
Folk psychology is a mixture of wisdom and superstition. Its compo-
nent of wisdom is obvious to anyone who has benefited from advice given
by a perceptive and experienced counselor. It is so important that science
cannot afford to ignore it, much as the engineer cannot afford to ignore all
of the craftsperson's lore. But the psychologists engaged in clarifying,
working out, and testing the valuable insights offolk psychology are likely
to end up formulating hypotheses and explanations going far beyond com-
mon sense. Never mind, for common sense is shackled to appearances, so
it belongs in the dock not in the jury.
Psychoanalysis and parapsychology are better articulated than is folk
psychology, which is totally amorphous. But all three share the same
methodological flaws: conceptual woolliness, lack of adequate experi-
112 5. Mentalism

mental evidence, and alienation from science. Worse, unlike folk psychol-
ogy, psychoanalysis and parapsychology want to be taken for sciences.
(The recent French variety of psychoanalysis, as represented by Lacan, is
an exception; it wishes to pass for a kind of humanistic clinical psychol-
ogy.) Indeed, psychoanalysis passes for being the science of the uncon-
scious and sexuality, and parapsychologists parade their experiments. In
fact, both are prime examples of pseudoscience. Let us give a few reasons
for this claim, focusing on only a few features of these doctrines. (For
details see Bunge, 1985b, 1985c.)
Psychoanalysis is conceptually fuzzy. In particular, psychoanalysts
have made no effort to elucidate the concepts of ego, id (or rather it), and
superego, of repression and resistance, of instinct and consciousness-
not to speak of those of racial memory and the collective unconscious.
None ofthese concepts is a variable in the sense of experimental psychol-
ogy. Consequently they cannot be functionally related to one another, as
is normally done in science. This is why nobody has succeeded in measur-
ing them or in mathematizing even a small fragment of psychoanalysis. In
these regards-measurement and mathematical modeling-psychoanaly-
sis compares unfavorably with behaviorism, neobehaviorism, psychophy-
sics, and physiological psychology.
Conceptual woolliness makes for poor testability. Psychoanalysis con-
tains untestable as well as testable hypotheses, but many of the latter have
never been put to the test; there are no psychoanalytic laboratories.
Freud (1962, p. 4) declared his theory of repression to be irrefutable, and
regarded this feature as a merit. He was right in one respect, for if a
subject fails to manifest complex X, then X is declared to be repressed
rather than nonexisting. Likewise, if the manifest content of the patient's
dreams is not sexual, then the latent or symbolic content is pronounced to
be. The same with aggression: if not overt, then latent. Here the psycho-
analyst's position is as impregnable to fact as is the theologian's.
However, when taken one at a time, some psychoanalytic hypotheses
tum out to be testable and, moreover, false (Bunge, 1967a). For example,
anthropologists and developmental psychologists have found that aggres-
siveness is more learned than innate, and that the Oedipus complex is
anything but universal. Personality experts have found no correlation
between personality type and early toilet training, hence no basis for
Freud's classing of personality types into anal and oral. Clinical psycholo-
gists have found no evidence that all neuroses are caused by repressed
sexuality; moreover, hypersexuality is regarded as morbid. Social psy-
chologists have found that violence and the watching of violent scenes,
far from having a cathartic effect, stimulate aggression. Sociologists and
political scientists laugh at the claim that the root of all social conflict is
the parent-child relationship. And scientific dream researchers have
learned that dreams have no purpose or "meaning," whence they are not
"the royal road to knowledge of the unconscious" (Freud)-that conven-
5.5. Pop Psychology 113

ient scapegoat. Consequently, the attempt to "interpret" dreams is just as


groundless as that of reading the future off cards or tea leaves.
In sum, it is not true that psychoanalysis is totally untestable. Some of
its testable components have been put to the test-albeit, not by psycho-
analysts-and found false. So much for psychoanalytic theory. (More in
Fisher & Greenberg, 1977; Griinbaum, 1984; Perrez, 1979; Rachman,
1963; Van Rillaer, 1980.) As for psychoanalytic therapy, in any of its
many versions, there is no statistical evidence that it is effective-con-
trary to behavior and drug therapies (Eysenck & Wilson, 1973; Prioleau,
Murdock, & Brody, 1983; Van Rillaer, 1980; Wolpe, 1981.) We shall come
back to this point in section 12.1.
Another feature that exhibits the pseudoscientific nature of psychoanal-
ysis is its alienation from science, in particular from biology and experi-
mental psychology. This is no accident: Freud (1929, p. 16) demanded
that psychoanalysis "must dissociate itself from every foreign preconcep-
tion, whether anatomical, chemical, or physiological, and must work
throughout with conceptions of a purely psychological order." In other
words, psychoanalysis is strictly mentalistic despite Freud's occasional
polite nods to neuroscience. Moreover, he encouraged laypersons, totally
lacking in scientific or medical background, to practice analytic therapy.
He was quite right that no such background is necessary, for psychoanal-
ysis lacks a specific background; it is literally groundless. (See section 3.5
for the concept of background.) In particular, psychoanalysis makes no
use of biology because it has no use for the nervous system, and it has no
need for experimental psychology because it is purely bookish and clini-
cal. Of course, it might use statistics-but it does not.
Why then, despite being neither true nor effective, is psychoanalysis so
popular? There are several causes: (a) It requires no scientific training:
Anyone trained in the tradition of psychophysical dualism can understand
it; (b) It is "relevant": It addresses-alas, wrongly-problems about
human nature that academic anthropology and psychology used to ne-
glect; (c) It reassures us by telling us that everybody is abnormal (a
meaningless proposition), so that there is no reason to be ashamed about
being abnormal; (d) It promises to cure disorders that academic clinical
psychology and psychiatry used to ignore or were impotent to treat; (e) It
is a world view with ready and simple answers to nearly everything; and
(f) It extolls instinct and belittles reason, which flatters those who do not
have much reason to spare.
The young Freud's philosophy of mind was psychophysical parallelism,
which he learned from Jackson, who probably took it from Leibniz. On
the other hand, Freud's mature philosophy of mind, which emphasized
psychosomatic interactions, was Cartesian interactionist dualism-the
same advocated nowadays by Popper and Eccles (1977). Parapsychology
too is dualistic, but it is not always clear which version of dualism it
espouses. Whereas telepathy (mind reading) seems to call for epiphenom-
114 5. Mentalism

enalism, telekinesis (movement at a distance) involves interactionist dual-


ism. In any case parapsychology is just as pseudoscientific as psychoanal-
ysis, if only because of their alienation from science.
However, there are some interesting differences between parapsychol-
ogy and psychoanalysis. For one thing parapsychology, unlike psychoa-
nalysis, has no theory to speak of, and consequently it is extremely bor-
ing, whereas the redeeming feature of psychoanalysis is that it is amusing
fiction. In fact, the only hypothesis of parapsychology is that there are
paranormal phenomena, such as precognition and reincarnation, which
defy scientific explanation and moreover contradict certain basic princi-
ples of scientific research (section 1.4). It proposes no mechanisms to
account for its alleged data. For example, in a well-known attempt to
account for telekinesis, Schmidt (1975) proposes that the probability per
unit time of events of a certain kind, occurring in a stochastic system such
as a radioactive source or a random-number generator, alters if the sys-
tem is under the action of a psychic subject. But this is just a redescription
of the alleged phenomenon; it does not offer any possible mechanism for
the alleged action of the psychic.
A conspicuous feature of parapsychological trials from the very begin-
ning is that their outcomes have rarely if ever been replicated by indepen-
dent researchers. (Alcock, 1981; Hansel, 1980; Marks, 1986.) Recurring
problems have been cheating and unconscious sensory cues. Another
problem is that the controls are usually so loose that no valid statistical
analyses of the data are possible (Diaconis, 1978). Whenever proper con-
trols have been introduced, no trace of psi has been found. For example,
Johnson and Jones (1984) refuted the sheep-goat hypothesis, that experi-
mental subjects who believe in ESP (sheep) score consistently at above-
chance level in telepathy and clairvoyance tests, whereas skeptics (goats)
score at below-chance levels. What is true is that parapsychologists con-
tinue to claim that their subjects are inhibited in the presence of skep-
tics-whence objective controls would be impossible.
The most radical of all the criticisms to which bona fide parapsychologi-
cal experiments have been subjected is this (Alcock, 1984). They are not
experiments proper because parapsychologists have no way of turning
their putative psi on or off like an independent variable, or even of know-
ing whether it is in operation at any specified instant during the study.
Besides, in order to be able to conclude that the observed frequency of a
given phenomenon (e.g., guessing cards) is at above-chance level, one
must have a definite model of the situation. Without such models there is
no way of refuting a null hypothesis. But, typically, parapsychologists
produce no such models; they only claim that, sometimes, some subjects
obtain above-chance scores. But this, when it does happen, is normal not
paranormal. In fact, such occasional departures from averages are char-
acteristic of chance events.
In sum, parapsychology has nothing to offer, except by way of method-
ologicallesson: Beware of cheating, self-deception, loose controls, and
5.6. Summing Up 115

sloppy statistics. On the other hand, folk psychology and psychoanalysis


have offered a handful of insights worth pursuing-besides being amus-
ing. Nevertheless, on balance psychoanalysis has been extremely perni-
cious for having reinforced the habits of wild speculation and of recourse
to authority (Freud or Lacan dixit). It has diverted many people from the
scientific path, it has wasted the health, time, and money of uncounted
mental patients. And last, but not least, it has reinforced irrationalism and
even reactionary ideologies (e.g., for holding that war is inherent in hu-
man nature, and that social conflict is inevitable for being a consequence
of the child-parent relationship).

5.6 Summing Up
Mentalist psychology conceives of the mind as a self-contained entity
divided into a number of compartments. It describes these compartments
with hardly any reference to the properties of the subject as an animal; it
attempts to account for the mental by the mental. As a consequence
mentalism places itself deliberately outside natural science and it refuses
to learn from it, particularly from neuroscience.
Moreover, mentalism deals primarily with humans, neglecting other
animals, as well as the processes of development and evolution. And,
except for psychoanalysis, which focuses on affect, mentalism deals al-
most exclusively with cognition, neglecting affect and behavior. In addi-
tion, except for the Gestalt school, mentalism divides cognition into a
number of mutually unrelated faculties.
Because of its detachment from biology, mentalist psychology is at best
semiscientific, at worst pseudoscientific. The two most pernicious ver-
sions of mentalism nowadays are psychoanalysis and computational
psychology, often wrongly equated with cognitive psychology. Unlike
classical psychology and the Gestalt school, psychoanalysis and
computationalism are speculative and dogmatic, and do not care for ex-
periment. Both abuse metaphors (e.g., the censor and the computer),
neither offers genuine explanations, and neither is effective in treating
mental disorders. The explanation as well as the repair of a system call for
some true knowledge of its mechanisms.
Psychoanalysis was counterrevolutionary at the time when classical
psychology was becoming increasingly experimental and getting closer to
physiology. Computationalism came as a counterrevolution six decades
later, when behaviorists were beginning to take an interest in the internal
states of the animal. Yet, for some reason, both psychoanalysis and com-
putationalism were hailed as scientific revolutions. Actually their main
accomplishment has been to propose simplistic solutions to some tough
problems, and to shift attention away from the nervous system as well as
from behavior.
CHAPTER 6

Behaviorism

Behaviorism is the once powerful school that employs the scientific


method to study the overt behavior of animals as a function of external
stimulation. It was born shortly before World War I in the United States,
where it quickly became the dominant trend in psychology and exerted a
strong and often beneficial influence on the social sciences. The early
doctrine was watered down two decades later with the introduction of the
so-called intervening variables and hypothetical constructs. The behav-
iorist movement started to decline in the mid-1950s with the revival of
mentalism in cognitive psychology and the reinforcement of physiological
psychology. At the time of writing, behaviorism is only a shadow of what
it was in the 1930s and 1940s. However, it is still going strong in clinical
psychology (see Chapter 12), and it may come back in basic psychology
as a reaction against the excesses of information-processing psychology.
(See sections 5.4 and 9.4.)
Behaviorism was invented as a reaction against traditional mentalism,
particularly against the proliferation of mental faculties and the predomi-
nant use of introspection. It had a primarily methodological motivation,
namely the wish to transform psychology into "a purely objective experi-
mental branch of natural science" (Watson, 1913, p. 158). The students
who adopted behaviorism solely for this reason, without buying its philos-
ophy of mind (or rather nonmind), may be called methodological behav-
iorists.
Other behaviorists, particularly Watson and Skinner, went further:
They also embraced the metaphysics and epistemology of positivism. The
positivist metaphysics boils down to the thesis that there are only phe-
nomena or appearances, that is, events appearing to some observer or
other. The epistemological consequence of this phenomenalist metaphy-
sics is that we can get to know only phenomena and their relations. (This
was a recurrent theme in Ptolemy, Cardinal Bellarmino-Galileo's main
foe-, Hume, Kant, Comte, Mach, Duhem, and the Vienna Circle.)
Those who embrace these two philosophical theses may be called philo-
sophical behaviorists-or positivists for short.
6.1. Phenomenalism (Black-Boxism) 117

Although the metaphysical thesis of positivism, namely phenomenal-


ism, entails its epistemological component, it is possible to admit the
latter without accepting the former. For example, information-processing
psychology admits mental entities side by side with Turing's functionalist
test of mindfulness, which is consonant with epistemological positivism.
In fact, this test is confined to observing the behavior of the system that is
attributed a mind. (On the other hand, a biopsychological test of mindful-
ness would involve a neurophysiological analysis.)
Whichever their philosophical motivation, the early behaviorists stipu-
lated that the object of study of psychology is the animal as a whole
regarded as an input-output system. Its innards, whether physiological or
mental, were to be ignored. This limitation to controllable stimuli and
measurable responses guaranteed objectivity and shallowness with one
stroke. It avoided the unbridled speculations of mentalism but, by the
same token, it wrote offthe mind altogether, and it blocked the search for
the neuromuscular and neuroendocrine springs of overt behavior.
Whence the ambiguity of the legacy of operationism: A revolution in
methodological rigor and a reaction with regard to the problematics and
aims of psychology.
We shall examine only the three most salient philosophical features of
behaviorism, namely phenomenalism (black-boxism), environmentalism,
and operationism. We shall also study the methodological novelty intro-
duced by neobehaviorism, namely the addition of variables intervening
between stimulus and response, which placed it midway and ambiguously
between radical behaviorism and either mentalism or biopsychology. The
reader interested in a detailed sympathetic study of the philosophies of
science and mind, and even the ideology, inherent in behaviorism, is
urged to read Zuriffs (1985) monograph.

6.1 Phenomenalism (Black-Boxism)


We call black-boxism the strategy of modeling systems as empty boxes
that respond only to environmental stimuli. Black-boxism can be method-
ological or metaphysical. Methodological black-boxism states only that,
whether or not a thing is complex, it must be conceived of, and treated
experimentally as, an empty box with inputs and outputs. Metaphysical
black-boxism adds that things are what they appear to be, as a conse-
quence of which any attempt to conjecture their innards (composition and
structure) is in vain.
The psychology resulting from adopting the black-boxist strategy is
called stimulus-response (S-R) psychology. Any psychology that takes
inner states of the organism into account is called stimulus-organism-
response (S-O-R). If, in addition, it accounts for the consequence the
response has for the way the animal handles the input (i.e., if it takes
118 6. Behaviorism

(a) (b) (e)

FIGURE 6.1. Three models of behavior. (a) Stimulus-response, or S-R; (b) stimu-
lus-organism-response, or S-O-R; (c) stimulus-organism-response-feedback, or
S-O-R-F. The spring symbolizes the inner mechanisms that mediate b'ttween S
and R.

feedback loops into account) the psychology may be called stimulus-


organism-response-Jeedback (S-O-R-F). See Figure 6.1.
Initially behaviorists focused on externalities, that is, Sand R, in their
attempt to rid psychology from mentalist fantasies. Eventually, as some
psychologists-such as the refiexologists and Lashley-persisted in fill-
ing the black box with more or less hypothetical neural mechanisms, the
orthodox behaviorists made it a rule not to "neurologize." For example,
Skinner (1938) stated bluntly: "Psychologists had better give up the ner-
vous system and confine their attention to the end-terms."
Behaviorism has focused its research on learning. The entire behavior-
ist teaching on learning is an ellipse with two foci: classical (or Pavlovian),
and operant (or Skinnerian) conditioning. In classical conditioning the
animal is held immobile and learns, without trial and error, to associate
stimulus with reward. In operant conditioning the animal is free to roam
and learns by trial and error. In the first case, the ad could read: Enjoy
now, pay later; in the second, Pay now, enjoy later.
In both cases the reward, rather than some internal process, is given
credit for the reinforcement. However, the underlying neural mechanisms
are bound to be quite different, if only because of the occurrence of
chance and active search in the case of operant conditioning, and their
absence in the case of classical conditioning. Regrettably, the behavior-
ist learning theories do not include any variables representing the
inner states, let alone the experience, of the animal. (See Luce, Bush,
& Galanter, 1963-65.) A black box makes no room for such extra
variables.
This deliberate neglect of the inner states of the organism has two
important consequences. First, unless the experimental animals are al-
ways (tacitly) prepared in the same way to face the experiment (e.g.,
starved), no pattern is likely to emerge. Thus, whereas one rat may learn a
trick after 10 trials, another may take as many as 100 trials. It is only after
the animal has learned to do the job (i.e., reached criterion) that it can be
relied upon to respond on cue.
6.1. Phenomenalism (Black-Boxism) 119

This explains the many exceptions to the alleged laws of learning found
by the behaviorist researchers. The only invariable pattern seems to be
the so-called zeroth law of rat psychology: "Any well-trained animal, on
controlled stimulation, will behave as he damn well pleases." In other
words, external stimulation is insufficient to determine behavior, hence to
predict it. The reason is not so much that there are no two identical
animals: All of the factual sciences face a similar diversity. The main
reason is that the output of any system, living or nonliving, is a function of
both the input and the internal state. And, because an animal is not
usually in the same state on different trials, it is unlikely to produce the
same output each time. In short, there are no general S-R laws. Lawful-
ness exists only for S-O-R triples.
A second consequence of the neglect of internal variables is that S-R
psychology has nil explanatory power. In fact, by definition, to explain
event A in system B is to exhibit or conjecture some mechanism C, in
system B, that, when triggered by stimulus D, will produce A (Bunge,
1983b, 1985g). Because black-boxism eschews mechanisms, it is at best
descriptive and predictive (like Ptolemaic astronomy), never explanatory
(like Newtonian astronomy). We shall return to the matter of explanation
in section 13.3.
S-R psychology cannot even account for the simplest motor act. In-
deed, every single motor act is the outcome of a complex process occur-
ring inside the organism, and none is the result of mere external stimula-
tion. For example, the maintenance of upright equilibrium and walking
involve a delicate control mechanism including sensors that detect minute
deviations from the equilibrium, (i.e., "errors"). Thus, far from being an
input-output box, the organism is an error-correcting device, that is, one
endowed with feedback mechanisms. (Recall Figure 6.1, part (c).) It is
then closer to a control mechanism, such as Maxwell's governor, than to
a stone. In fact, it does not act in the direction of the stimulus but in the
direction that decreases the "error" or imbalance. (Animal behavior ex-
emplifies Le Chatelier's principle, not Aristotle's mechanics.)
From a biological point of view, then, overt behavior is a manifestation
or external outcome of processes occurring inside the animal. In the case
of animals endowed with nervous systems, behavior is a manifestation,
hence an indicator, of processes occurring in their neuromuscular system.
Shorter: For these animals the organ of behavior is the neuromuscular
system.
The methodological consequence is obvious: If we wish to explain
behavior, not just to describe it in superficial terms, we must study the
neuromuscular system. This holds both for simple behavior, such as turn-
ing the head on hearing a tone, and for complex behavior, such as writing
a poem. The more complex the behavior the more its understanding calls
for a study of its source in the neuromuscular system. In other words, the
study of behavior cannot be confined to behavior; it must embrace all the
120 6. Behaviorism

determinants of behavior. Because in the higher animals some of the


determinants of behavior are mental processes, such as willing, an ade-
quate understanding of behavior involves an understanding of the mind.
This does not imply giving up behaviorism and giving in to mentalism, but
enlarging the narrow perspective of behaviorism to the point where it
becomes psychobiology-for which see chapters 8 and 9.
What holds for overt behavior holds also, a fortiori, for the higher
functions of the nervous system, (i.e., the cognitive and executive ones).
In particular, perception is anything but the passive and automatic record-
ing of external stimuli; it is strongly influenced by emotion, expectation,
and attention. Thus in humans even a simple perceptual task such as that
of discriminating between horizontal and vertical lines requires the
"searchlight" of attention (Sagi & Julesz, 1986). Without this internal
condition the stimuli may remain imperceptible.
Our criticism of black-boxism does not entail a rejection of particular
black-box models (Bunge, 1963b, 1964). A black-box construct, be it a
diagram or a theory, is better than none, if only because it may motivate
the construction of a mechanismic theory accounting for the same facts in
a deeper manner. For example, one of the goals of physiological psychol-
ogy ought to be to deduce Thurstone's famous learning function, the
logistic curve, from a system of hypotheses concerning the change of
synaptic weights or strengths under the repeated application of a stimu-
lus. In sum, a black box should be regarded as the alpha, not the omega,
of scientific theorizing.

6.2 Environmentalism
Environmentalism is the thesis that animal behavior is exclusively deter-
mined by environmental circumstances. To put it negatively: Environ-
mentalism denies that inheritance and internal processes are relevant to
behavior. It is the opposite of nativism, according to which we are mainly,
if not solely, the actualization of our genetic blueprints.
There are two main environmentalist schools in psychology: classical
behaviorism and ecological psychology-for which see Gibson (1950).
Both hold that behavior is the sole effect of environmental stimuli; both
model the organism as an empty box. The difference between them is
that, whereas ecological psychology focuses on perception, behaviorism
is mainly interested in overt behavior.
The environmentalism inherent in behaviorism explains the stand that
this school has taken with regard to instinct, intelligence, personality,
education, and social programs. According to classical psychology and
ethology, behavior can be either of two kinds: instinctive or learned. The
former would be inborn, automatic, and unmodifiable by experience.
Moreover, it would be the same for all the members of a species, and it
6.2. Environmentalism 121

would be characteristics of the species (i.e., species-specific). This is the


view of the classical ethologists Lorenz and Tinbergen, who postulated
the existence of "innate releasing mechanisms" that are triggered by
certain environmental stimuli but, once triggered, coordinate the instinc-
tive act in total independence of the receptors, and keep going until the
goal has been attained.
Behaviorists have never believed in instinct. But, because of the nar-
row focus of their research, they were unable to refute the instinct doc-
trine. The refutation came in the 1930s and 1940s from embryological and
developmental studies (Lehrman, 1953). For example, it was found that
pecking develops before hatching and improves as the chick gains in
strength. Chicks that are kept in the dark and are force fed for the first
twelve days of life starve to death when placed in a normal feeding situa-
tion, because they have had no occasion to associate pecking behavior
with visual stimuli or with food. Second example: In normal situations
rats build nests in a corner of their cages. But when raised in isolation,
and prevented from manipulating or carrying any objects, they do not
build normal nests or retrieve their young normally; they have had no
chance to learn normal maternal behavior from ordinary manipulation
and collection of food and other things. The results obtained with dogs,
monkeys, and other species are parallel. In short, overall, nativism is
wrong: Instinctive behavior is not rigid but can be altered by learning-
nay, it often develops by early learning, and it aborts without it.
We have just concluded that overall nativism is wrong. This is not to
deny that some behavioral and mental traits are totally inborn. Red-green
"color blindness" is a case in point. This has been known of old, but solid
experimental evidence that red-green "color blindness" is caused by
alterations in the genes "encoding" red and green visual pigments was
produced only recently (Nathans, Piantanida, Eddy, Shows, & Hodges,
1986). Only similar molecular genetic studies could establish beyond rea-
sonable doubt the innate character of any other behavioral or mental
traits.
The methodological moral is obvious: By dogmatically decreeing that a
given type of behavior is instinctive or inborn, one is likely to block
experimental research into it. Indeed, one then evades the problem of the
origin of such behavior, and one makes no effort to alter it by experimen-
tal means, or to identify the corresponding genes. Moreover, one adopts a
teleological rather than a causal mode of thinking. In particular, "Lorenz
sees the developing behavior of the animal as progressing toward the full-
blown instinct rather than as developing out of interactions among pro-
cesses present at that stage" (Lehrman, 1953, p. 352). It is far more
instructive to attempt to explain behavior, even instinctive behavior, in
terms of development under genic guidance and environmental con-
straints and stimuli. After all chicks, rats, and people are systems and,
like all systems, they have a composition (partly inherited), an environ-
122 6. Behaviorism

ment, and a structure (Recall section 3.2.) Neglecting either of these


aspects is bound to result in serious distortion.
What holds for instinct holds, mutatis mutandis, for intelligence-
which behaviorists mistake for performance, an indicator of intelligence.
Nobody can reasonably doubt that runners do best if they have inherited
long legs and good cardiovascular systems, nor can anyone reasonably
doubt that such a favorable genetic endowment would be useless without
proper training. But when it comes to intelligence, people tend to reason
differently or not at all, perhaps because they regard it as something
supramaterial, or because of class or race prejudice. Whatever the cause,
the fact is that, until recently, most experts believed that intelligence is
either fully inherited or fully learned. Once again the truth lies somewhere
in between: Intelligent people, just like athletes, are a product of suitable
organs properly trained and given the chance.
The best known evidence for the inheritability of intelligence comes
from follow-up studies of monozygotic twins reared together or apart. But
this evidence is flawed on several counts. First, the twins reared apart
have not been randomized, nor have their foster parents; there was
choice, hence bias, not chance. Second, most ofthe studies rest on intelli-
gence test scores, which cannot seriously be taken as objective measures
of global intelligence. Only very few of the studies use scholastic perfor-
mance as an intelligence indicator, and these have shown that family
background can be stronger than heredity: Children placed in low-income
and poor-educational-level families perform far worse at school than chil-
dren placed in families with a higher intellectual atmosphere, and this
regardless of their biological parents. So, although there is a strong genic
component, there is also a strong environmental one. (See e.g., Bouchard
& McGue, 1981.) Third, due to the intimate intertwining of hereditary and
environmental factors, it is next to impossible to effect a neat separation
of the total variance with respect to a given characteristic, such as intelli-
gence, into a genetic, an environmental, and an interactive part. In gen-
eral, the statistical analyses of this problem have been for the most part so
incorrect that they are practically useless. (See e.g., Gould, 1981; Kamin,
1974; Layzer, 1974.)
We do not know much about intelligence, but we can surmise this
much. First, because the mental is neurophysiological and the bodily is
partly inheritable, we are bound to inherit predispositions or propensities
to excel in certain behavioral and mental tasks while doing poorly at
others. However, so far there is no direct empirical evidence for this
thesis, and none will be forthcoming as long as the various components of
intelligence are not located in the brain. Still, a beginning has been made;
it has been found that, in mice, the velocity of nerve conduction is herita-
ble (Reed, 1983) and so is synaptic transmission time (Reed, 1984). Sec-
ond, we know that, because our brains are largely organized by external
stimuli, in particular social ones, the environment facilitates the actualiza-
6.2. Environmentalism 123

tion or expression of some potentialities while it inhibits that of others. In


sum, whatever empirical evidence we do possess supports the nature cum
nurture view. Further evidence is more likely to come from more studies
in the genetics of the brain and in the physiology of learning than from
more twin studies.
Personality theories too usually have been either innatist or environ-
mentalists. (See e.g., Cartwright, 1979.) Thus the ancient view that
classes personalities according to the predominance of one of the "hu-
mors" is an early form of nativism. On the other hand, the psychoanalytic
fantasy about oral and anal personalities takes the side of education.
Moreover, it is unicausal, for restricting education to toilet training and
ignoring inborn predispositions as well as cognitive and moral develop-
ment. To be sure, a strict and austere education-which among other
things is likely to involve early strict toilet training-may contribute to
molding a serious, industrious, and worry-prone personality, but the ef-
fects of early training may get lost during a rebellious adolescence. Like-
wise, a sloppy child may develop into a finicky adult, but a uniformly
poor, oppressive, and hopeless social environment may churn out individ-
uals with roughly the same apathetic and pessimistic type of personality.
Here again, behaviorism is likely to correct the excesses of nativism.
So much for the observational evidence relevant to the nature-nurture
issue. What about the experimental evidence for or against the behavior-
ist thesis that behavior is a function of external stimulation only? Let us
recall a few classical experiments refuting that thesis. One of the earliest
was Lashley's finding that the efficiency of a stimulus depends critically
upon the "set" (Einstellung, attitude, disposition, preparation) of the
animal, something the experimenter is not always able to control. In fact,
only stimuli for which the animal has been "set to react" are effective. In
particular, unless the animal is paying attention to the visual experimental
stimulus, it may not react to it at all. In electrophysiological terms: For a
stimulus to trigger an action potential, a readiness potential must have
been built before the stimulus is applied.
Another relevant finding is that of Hebb on the effects of sensory depri-
vation. The human subject who is prevented from seeing, hearing, and
even touching, is not reduced to a vegetative existence, but continues to
experience an intense if abnormal mental activity. True, a strict behavior-
ist may write off this experiment because the subject is not allowed to
move, and because the experimenter has to rely on questioning, which is
indirect introspection. But by doing so the behaviorist would only adver-
tise the narrowness of behaviorism.
A much cheaper experiment is to present the subject with an ambiguous
figure such as a Necker cube or Boring's old/young woman. After a few
seconds a gestalt switch occurs and we see the previously concealed
object, even though the stimulus has not varied. Neither the Gestalt psy-
chologists, who discovered this phenomenon, nor the behaviorists, who
124 6. Behaviorism

have kept an embarrassed silence about it, have an explanation for it.
Physiological psychology can explain the phenomenon in terms of neuro-
nal habituation or fatigue. It is well known that, if a neuron is subjected to
a constant stimulus, it becomes "used" to it; its rate of firing dies off. The
gestalt switch can be explained by assuming that there are at least
two neuronal systems involved, which are activated alternatively by
the ambiguous figure. As one of the systems has become habituated,
the other one is ready to take over. An even simpler experiment
indicates the importance of context (frame of reference), experience,
and expectation. Subjects are shown the symbol '0.' Whereas some
of them read the letter 0, others read the numeral zero. All of them
sense or detect the same stimulus, but each subject perceives it dif-
ferently, according to the immediately preceding circumstances and
the current context. (More on the sensation-perception difference in
section 9.2.)
There is one case, though, where even higher animals do seem to be-
have in a behavioristic manner. This is the case of animals that have had
their hippocampus excised. Their behavior is completely controlled by
external stimuli, and they learn only gradually by reinforcement. In fact,
the behavior of "hippocampally ablated animals is held to be everything
which the early S-R theorists could have wished" (Hirsh, 1974, p. 439).
Not so the behavior of normal animals; theirs is controlled not only by
current stimuli but also by their memories, drives, and expectations. For
this reason Bandura (1978) rejects environmentalism and favors "triadic
reciprocal determinism," where environment, thought, and behavior in-
teract.
The idea that thought might be involved in learning and problem solving
was of course anathema to behaviorists and reflexologists alike. The idea
was suggested by the physiological psychologist Karl Lashley (1929) and
put to the test in rats by Krechevsky (1932) and Tolman and Krechevsky
(1933). They fouhd that the rat does not behave haphazardly during the
early stages of a problem-solving process, but proceeds quite methodi-
cally, "attempting various solutions and giving them up when they fail,
until he hits finally upon the 'correct' one" (Krechevsky, 1932). This
work was severely attacked by such behaviorists as K. Spence, as a
consequence of which it was forgotten by American psychologists for
nearly a quarter of a century.
The hypothesis that the higher animals learn by making hypotheses and
putting them to the test resurfaced in the United States with the influential
book by Bruner, Goodnow, and Austin (1956), and was subsequently
investigated both experimentally and mathematically (e.g., Levine, 1974;
Mayer, 1977; Restle, 1976). One finding of this research is that human
learners are not passive; their response depends upon their learning his-
tory and their expectations. Another finding vindicates the Gestalt thesis
that in some learning tasks the subject can learn at a single trial. Both
6.3. Operationism 125

results refute the behaviorist learning theory, which has always been the
centerpiece of the behaviorist movement.
In sum, observation and experiment have refuted environmentalism,
particularly with regard to learning and development. This was to be
expected on the mere strength of an elementary knowledge of physics,
which shows that nothing in the world behaves passively and solely in
response to external forces. For example, the curvature of the path of an
electron entering an external magnetic field depends not only upon the
intensity of the latter but also upon the initial momentum of the electron.
What is strange is that anyone in our century could believe, following
Aristotle, that external forces are omnipotent. Moral: Ignore the neigh-
boring sciences, and the ontologies compatible with them, and you are
bound to embrace some archaic myths.
Nevertheless, by emphasizing, nay exaggerating, the weight of the en-
vironment and denying that of heredity, behaviorism has exerted a benefi-
cial influence on educational psychology, educational policies, and social
programs, both in the United States and in the Soviet Union. (See e.g.,
Suppes, 1978.) Moreover, it has acted as the theoretical bulwark of egali-
tarianism and the bugbear of racism and social elitism. On the other hand,
nativism has always justified social and educational inequalities, and con-
tributed to preserving them wherever it has been dominant, for example,
in the United Kingdom. (See Kamin, 1974 and Gould, 1981.) Presumably,
a more balanced view of the nature-nurture issue will eventually result in
less naive (though hopefully no less fair) social policies aiming at match-
ing the genetic endowment with education, social opportunity, and social
responsibility. However, the nature-nurture controversy is likely to re-
main an ideological hot potato even after it is given a scientific solution.
We shall return to this debate in section 7.3.

6.3 Operationism
Operationism is the semantical and methodological doctrine according to
which "the concept is synonymous with the corresponding set of opera-
tions" (Bridgman, 1927, p. 5). For example, intelligence is declared to be
that which is measured by intelligence tests, and competence is conflated
with performance.
Operationism, a chapter of logical empiricism (or positivism), was pop-
ularized in psychology by Stevens (1935), Pratt (1939), Skinner (1945),
Spence (1948), and a few others. They all followed Bridgman (1927), who
in turn had been preceded by Dingler and others. It has been a constant
companion of behaviorism (Zuriff, 1985). In fact it is the behaviorist pre-
scription for scientific concept formation, as well as the behaviorist se-
mantics, which answers every question of the form 'What is X?' with
another question, namely 'How is X observed, measured, or controlled?'
126 6. Behaviorism

Several objections were raised against operationism. One was that,


because any given property can actually or in principle be measured in
different ways (with the help of different techniques), there would be a
multiplicity of concepts (of e.g., length) even when theory assumes a
single concept. Another problem was that the most powerful and interest-
ing concepts of physics, such as that of electromagnetic field intensity, are
rejected for not denoting directly observable properties-which results in
an intolerable impoverishment of science.
In the face of these and other difficulties Bridgman (1959) eventually
relented and admitted the need for "pencil and paper operations "-a
bashful way of naming concepts. However, operationism is still going
strong in the sciences. Thus, many physicists confuse "defining" with
"measuring," and many psychologists doubt the existence of any subjec-
tive experiences that are not yet amenable to measurement. Worse, many
experimentalists waste precious resources measuring features that have
not been adequately conceptualized, thereby producing heaps of useless
data.
However, the main problem with operational definitions is not that
they fail to yield unique concepts, or that they cut out the most power-
ful concepts, or even that they condone blind measurement. The main
problem is that there just ain't any operational definitions (Bunge,
1967a, 1974a, 1974b). Indeed, a definition is a purely conceptual af-
fair, just as much as a deduction is. Every definition proper character-
izes a construct in a precise and exhaustive fashion; it does not describe
a concrete thing or a property thereof. And it consists in identifying
two items: the new one with a previously introduced construct. Let us
explain.
Definitions can be explicit or implicit. An explicit definition stipulates
by convention that one concept is identical to another, or to a combina-
tion of concepts. Examples: "Somnambulism = df sleepwalking," and
"X = df (1 /n)'i.;J(." An implicit definition is an identity where the defined
concept does not occur alone in the left-hand side (LHS). Examples:
"p~q = df not -p or q" , and "X is an intervening variable = df X mediates
between a stimulus and a response." Axiomatic definitions constitute an
important subset ofthe class of implicit definitions. They occur in axioma-
tized theories, such as probability theory, classical mechanics, and some
learning theories. An axiomatic definition is of the form "x is an F= df X
satisfies A," where A is the conjunction of those axioms of a theory that
contain the predicate F.
All definitions are identities, that is, formulas ofthe type "A = B". The
left-hand side is or contains the defined concept or dejiniendum, whereas
the right-hand side is or contains the defining construct(s) or dejiniens.
These identities are stipulated to hold between constructs, not things or
features of things. In other words, definitions define concepts, not what
these may denote or represent. (For example, we may define a concept of
personality, but not the personality of an individual; the latter must be
6.3. Operationism 127

described.) Definitions are set up by convention or stipulation, not on the


strength of measurements. (For example, we may measure a memory
deficit provided we have a reasonably clear definition of the concept of
memory deficit.) This is why mathematical logic, the most general branch
of formal logic, is in charge of the theory of definitions. (For more on
definitions see Bunge, 1974b. For the formal-factual dichotomy see
Bunge, 1985a.)
Let us return to operationism. Because of its fundamental inadequacies
operationism has had a pernicious influence in physics, where it has (a)
cast doubt on the adequacy of some of the deepest concepts, which refer
to things or properties that are not directly accessible to measurement,
and (b) inspired a semi subjectivistic interpretation of the relativistic and
quantum theories, by demanding that everyone of its formulas be inter-
preted in terms of measuring instruments and observers (Bunge, 1973a,
1985a).
On the other hand, operationism had initially a beneficial influence on
the soft sciences by weeding out woolly concepts and untestable conjec-
tures. It has often led to a better understanding of the need for preparing
theories to confront empirical tests by enriching them with indicators or
diagnostic signs. We submit that such indicators, or links between unob-
servable variables and observable ones, are what operationists call 'oper-
ational definitions.' However, because they are hypotheses to be justified
empirically and theoretically, not stipUlations, they are more appropri-
ately called indicator hypotheses. (Recall section 4.3.)
An indicator hypothesis may be precise or imprecise, according to
whether it is formulated mathematically or not. But in any case it must be
fully testable and may therefore be improved on or even replaced with a
better one. Far from replacing or defining theoretical concepts, indicator
hypotheses are to be added to factual theories in order to render them
empirically testable. Such enrichment of a theory in preparation for its
actual test is sometimes called operationalization. A hypothesis or theory
that cannot be operationalized, or cannot be logically linked to any opera-
tionalizable constructs, is sheer speculation and therefore does not qual-
ify as scientific. I submit that this is the important grain of truth, and the
positive legacy of operationism.
The operationalization of a hypothesis or of a theory, that is, its prepa-
ration for empirical tests, proceeds as follows. Call T the theory to be
tested, and S the set of subsidiary assumptions specifying particular fea-
tures of the referent, for example, the more or less conjectured history
and dispositions of the experimental subject(s). From T and S we build a
model M (or special theory) of the referent(s) to be subjected to tests.
Enter now the set I of indicators ("operational definitions" or criteria)
constructed with the help of the body of antecedent knowledge as well as
of T itself. (Very often a theory suggests some of the indicators that might
link its unobservables to observables.) Enter also the set E of available
empirical data concerning the referent(s) of the theory T and relevant to
128 6. Behaviorism

the latter. Such data are often somewhat removed from the theory. For
example, they may consist of behavioral information, whereas the theory
T is about brain processes.
In other words, the indicator hypotheses I bridge the theory-experi-
ence gap, for they allow us to "read" observable (e.g., behavioral) events
in terms of unobservable ones (e.g., brain events). We may call this a
translation of the available raw data E into the language of the theory.
These translated data E* , that follow logically from (are entailed by) E and
I are then fed into the model M of the referent(s), to yield what may be
called the translated model M*. Finally, this result is translated back into
the language of experience by means of the indicator hypotheses I. That
is, M* is joined to I, to entail T*, the operationalization ofthe theory Tor,
rather, of the model M. Hence not T itself but some consequences of T
together with the subsidiary assumptions S, the data E, and the indicator
hypotheses I, face whatever fresh experience may be relevant to T. See
Figure 6.2. (For details and examples see Bunge 1973b.)
A few examples will show how indicator hypotheses can be used and
misused in psychological research. All vertebrates seem to have been
making use of indicators from time immemorial. In particular, they have
been "interpreting" (reacting to) postures and gestures of con specifics as
guides to social behavior. Mating and fighting rituals are only the most
salient and best studied cases. But we use plenty more indicators in daily
life to "read" moods, attitudes, and intentions off gestures and facial
expressions. (Some of our pets do the same with us.) Of course, such

T s E

FIGURE 6.2. Operationalization of a general theory T. T, together with the specific


assumptions S, yields a model M of the referent (the experimental subject). The
raw data E are translated into the language of the theory T by means of the
indicator hypotheses ("operational definitions") I. The result is E*, which, to-
gether with the model M, yield the testable consequences M*. Finally, these are
translated back into the language of experimental science by means of the same
indicator hypotheses I, to yield T*, which is ready to be confronted with fresh
data.
6.3. Operationism 129

indicators are highly ambiguous. For example, blushing may indicate an-
ger, excitement, or embarrassment.
Scientific indicators, unlike those of folk psychology, are supposed to
be unambiguous. Or, if ambiguous, they are supposed to come in batches,
so that ambiguity can be minimized or even totally removed. An early
example of a scientific indicator in physiological psychology was Pavlov's
use of the number of drops of gastric juice secreted by a dog on presenta-
tion of a bit of food, or of a stimulus associated with it, as an objective
indicator of the effect of the stimulus. (Presumably, he was measuring the
strength of the animal's expectation to sate its hunger.)
Psychological indicators are oftwo types: behavioral and physiological.
The former are seldom any good to disclose mental processes for the
simple anatomical reason that most neuronal systems do not innervate
any muscles. This is why we are forced to supplement behavioral indica-
tors with physiological ones, such as the frequency of firing of certain
neurons.
Finally, we must warn against two common methodological mistakes.
One is the belief that one has set up an "operational definition" just
because the terms occurring in the definiens (or defining formula) are
observational or nearly so. For example, some behaviorists have claimed
that the following is an operational definition of the concept of a rein-
forcer:
Stimulus x is a positive reinforcer of behavior y = df the presentation of
x increases the probability of the occurrence of y.
However, this is an ordinary definition, (i.e., an identity). Moreover, it is
not the only possible definition of "reinforcer"; in fact, one could think of
an alternative definition, in terms of neurophysiological processes.
Another common confusion is that between a mental process or ability
and an objective test or criterion for it. For example, Moore and Newell
(1974) ask how the computer program MERLIN understands, and reply
with this alleged definition:
x understands y = df X uses y whenever appropriate.
But according to this "definition" anyone, animal or machine, that per-
forms an adaptive action, uses a rule, or follows an instruction, can be
said to understand it. Thus a frog uttering a mating call would be said to
understand it, and the psychologist using his or her brain to solve a
problem would be said to understand the brain.
In conclusion, there are no operational definitions, whence any reason-
able version of behaviorism must give up operationism. Instead, there are
objective indicators of unobservable (in particular subjective) properties
and events. Such observable-unobservable (e.g., behavioral-mental) re-
lations are testable hypotheses, not conventional identities. The experi-
mental test of any scientific theory involves the use of such indicator
130 6. Behaviorism

hypotheses, to achieve what has been called the "operationalization" of


the theory. But this operation, far from effecting a reduction of the theory
to laboratory operations, enriches the theory with further hypotheses,
namely indicator conjectures. The closer you want to bring a theory to
experience, the more hypotheses you have to adjoin.

6.4 Intervening Variables


Classical behaviorism, from Watson to Skinner, was strictly a stimulus-
response, or S-R, psychology-or, rather, ethology. It pursued the admi-
rable goal of finding the general (even cross-specific) laws of animal be-
havior without resorting to any of the mentalistic fantasies that eluded
observation. But classical behaviorism failed to attain its aim for ignoring
the internal states of the organism, which are in the habit of changing
relentlessly.
In fact, as noted in section 6.2, it is hardly possible to find constant
objective patterns linking responses to stimuli unless the internal states
are taken into account. Sometimes the input has one given effect, and at
other times it has a different one. Something happens inside the organism,
between one moment and the next, that prevents it from reacting in the
same way to a given environmental stimulus. For example, some key
neurons have become inhibited, or else fatigued (habituated). Because the
aim of science is to find laws or utilize them, the aseptic S-R program of
classical behaviorism had to be altered. The least painful way of doing so
was to introduce a third set of variables, called intervening variables, for
being supposed to mediate between stimuli and responses.

This important advancement within the behaviorist movement was led


by Tolman (1932), Vygotsky (1978), Hull (1952), Berlyne (1975), and a few
others. These innovators were collectively called neobehaviorists, and
they formed the very last phase of behaviorism. (We have included Vy-
got sky in this group, although he is usually counted in the dialectical
materialist school, because he espoused the behaviorist methodology and
paid no attention to the nervous system.)
The neobehaviorists seem to have had two motivations for departing
from orthodoxy and adopting a S-O-R type of psychology. (For an expla-
nation of the S-O-R label see section 6.1.) One was metaphysical and
therefore not willingly confessed: Namely the desire to bring the psyche
back into psychology. This was particularly obvious in the cases of
Tolman, who wrote freely about purposiveness; of Hull, who made use of
dispositional concepts and took valuation seriously; and of Berlyne, who
studied thinking and esthetic judgment. Needless to say, we applaud the
open-mindedness and courage of the neobehaviorists, and only deplore
that they did not go further.
6.4. Intervening Variables 131

The methodological motivation was far more clear: It was the wish to
find S-O-R laws (Zuriff, 1985). A paragon and an example will help under-
stand this point. The ohmic resistance in a direct current circuit may be
said to "mediate" or "intervene" between the input e (the electromotive
force) and the output i (the current intensity), which Ohm's law relates in
the form: e = Ri. In the classical theory, R is an uninterpreted parameter
measurable, for example, via e and i. It is not explained in terms of the
particular structure of the material. On the other hand, solid-state phys-
ics, which analyzes the metal as a network of atoms inside which the
charge-carrying electrons move, does explain R as a particular property
of the system. That is, it regards R not as an intervening variable but as a
hypothetical construct-on which more in the next section.
Our example is this. If stimulus SI sometimes elicits response fl and at
other times response f2, whereas S2 is sometimes paired off to f2 and at
other times to fl, there is no apparent pattern. Lawfulness is restored by
introducing an intervening variable i and conjecturing the correspon-
dences
(SI, i l) ~ fl , (SI, i 2) ~ f2

(S2, i l) ~ f2 (S2, i 2) ~ fl

The intervening variable i is not directly accessible by purely behavioral


techniques. It must be either conjectured or inferred from observed (s, f)
pairs. (Such inference works only for a finite and moreover manageable
table; it fails for a continuous variable.)
A more rigorous and general conceptualization is this. S stimulates or
inhibits an internal variable I, which in turn has an output R in accordance
with the commutative diagram:

s ..------=g--_ r=f(i), i=g(s)

:. r = f(g(s)) = h(s)

f with s E S, i E /, and r E R.
h

where h = fog is the composition of the functions f: I~R and g: S~I.


In this construal the "intervening variable" is not a variable but the set I
together with the functionsfand g. In short, j = (I,f, g).
The main problem with this idea of an intervening variable is that there
are infinitely many possible triples (I,f, g) consistent with the observed (s,
f) pairs. Moreover, I could be as precise but noncommittal as a period of
132 6. Behaviorism

time, or as vague but specific as a habit strength or a drive. The only way
to remove such arbitrariness or indefiniteness is to abandon the purely
formal or syntactical construal of an intervening variable and have it
represent some specific state of the organism. But this passage from inter-
vening variable to hypothetical construct deserves a new section.

6.5 Hypothetical Constructs


Once intervening variables were admitted into the fold of psychology,
there seemed no way to control their proliferation. Psychological theories
risked becoming mere devices for cranking predictions out of data, in
obeisance to the conventionalist prescription. It was necessary to draw
the line between admissible and inadmissible variables inaccessible to
direct observation. MacCorquodale and Meehl (1948) call them interven-
ing variables and hypothetical constructs respectively.
The authors of this crucial distinction between two types of psychologi-
cal variable explain that an intervening variable' 'will involve no hypothe-
sis as to the existence of nonobserved entities or the occurrence of unob-
served processes" (MacCorquodale & Meehl, 1948, p. 103). On the other
hand, hypothetical constructs "are not wholly reducible to empirical
terms; they refer to processes or entities that are not directly observed"
(MacCorquodale & Meehl, 1948, p. 104). (For more on the distinction see
Tuomela, 1973.)
An intervening variable is one that relates stimuli to responses without
being interpreted in either psychological or physiological terms; it is an
auxiliary symbol discharging merely a syntactic or mathematical function.
On the other hand, a hypothetical construct is a variable mediating be-
tween stimuli and responses, and which is assigned a specific psychologi-
calor physiological interpretation (e.g., as motivation or memory). One
characteristic of cognitive psychology is that it is full of hypothetical
constructs, such as mental maps and models. A characteristic of physio-
logical psychology is that all of its hypothetical constructs are assigned
physiological interpretations. In particular, a physiological explanation of
a psychological process will involve some mechanism describable in
terms of properties that happen to be hypothetical constructs-not be-
cause they escape observation but because their observation calls for
neurophysiological (e.g., electrophysiologicai) techniques. See Fig-
ure 6.3.
An important concomitant of the introduction of intervening variables
and hypothetical constructs was the beginning, in behaviorist psychology,
of serious theorizing, and in particular of mathematical modeling. The
early behaviorists had been hostile to theory, just as much as their philo-
sophical mentors, the positivists, had been. (Even as late as 1950 Skinner
denied the need for theories of learning.) Their successors deserved the
name neobehaviorists.
6.5. Hypothetical Constructs 133

S'-IIQ06O\t:r, or r,
s,_. . r, or r,
(a) (b) (e)

FIGURE 6.3. Three stages in the account of overt behavior. (a) Direct S-R relation,
e.g., number of hours offood deprivation~rate of bar pressing; (b) Interposing an
intervening variable (e.g., hunger), with no definite neurophysiological interpreta-
tion; (c) Hypothesizing or finding the neural system possessing the property rep-
resented by the intervening variable in (b). The spring in the box suggests that the
latter, far from being black, contains a mechanism that makes the organism be-
have now in one way, now in another (e.g., press the bar when hungry, but not
press it when hungry and curious about a novel stimulus). (From Bunge, 1985b.)

The earliest and most enthusiastic theorists in the neobehaviorist camp


were Tolman and Hull. The latter seems to have been the first psycholo-
gist of his generation to gain a correct grasp of the meaning of "theory"
and to appreciate the advantages of ax iomatics-which did not endear
him to the profession. Hull realized that the axiomatic presentation of a
theory renders explicit most of its assumptions and facilitates the deriva-
tion and consequently the empirical checking of the logical consequences
of the initial assumptions (axioms).
Hull's (1952) particular "behavior system" can easily be criticized for
not supply a list of the basic (defining or primitive) concepts; that he used
only elementary functions; that time does not occur explicitly in his pos-
tulates; that certain exponents, such as 7.6936 and -1.0796, are likely to
contain several nonsignificant figures, and that his notation was unneces-
sarily cumbersome. But this would be carping: The point is that Hull
dared to theorize in an exact fashion at a time when theorizing was
frowned upon by most psychologists.
Hull's "behavior system" was criticized for the wrong reasons. First,
for containing hypothetical constructs, such as "habit strength." Second,
for coming close to the axiomatic format. Psychologists, along with many
other scientists, had an irrational fear of ax iomatics-they believed that it
is a straitjacket, as Hebb once said. Actually axiomatics, because it puts
all the cards on the table and minimizes hand-waving, facilitates the criti-
cal examination of theories and their experimental test, and consequently
their correction or replacement (Bunge, 1973a).
The orthodox behaviorists were unhappy about both intervening vari-
ables and hypothetical constructs because these concepts go beyond what
is immediately observable without the help of the biological parapher-
nalia. However, the neobehaviorists clung to those concepts because they
wished to explain what they observed. For example, thirst seems to ex-
plain drinking behavior, and curiosity would seem to explain exploratory
134 6. Behaviorism

behavior. In all sciences, the explanation of an observable fact calls for


the conjecturing of some unobserved facts.
Still, there was a problem with the distinction drawn by MacCor-
quodale and Meehl, namely that a number of psychological concepts,
such as those of drive, expectation, and plan, can be construed either as
intervening variables or as hypothetical constructs-that is, as long as the
latter are not linked to neural processes. But the neobehaviorists resisted
such links, and insisted that hypothetical constructs be analyzed in purely
psychological terms. By behaving in this conservative manner they came
close to the old enemy, mentalism. Moreover, they gained only an illusion
of explanation. Thus, to say that a rat is drinking because it is thirsty may
well be true, but it does not tell us anything that we did not know before
starting research.
If the referent of a hypothetical construct is deemed to be a determinant
of behavior, then it cannot be anything other than a neural (or neuromus-
cular, or neuroendocrine, or neuroimmune) process, for behavior is an
outward manifestation of processes occurring in the nervous system and
in systems controlled by it. (We are tacitly referring, of course, to animals
that do possess a nervous system.) But this was far from obvious in the
1940s, except to a few individuals like Lashley. It was not until Hebb
(1949) published his influential book that a substantial number of scientists
realized that if the psychologist "is to work with hypothetical constructs
he must define his constructs neurologically-whether neurology is ready
for us or not" (Krech, 1950, p. 289). But this is another story, to be told in
the next chapter.

6.6 Summing Up
In retrospect behaviorism appears as the culmination of the protoscienti-
fic stage of psychology. It was a mixture of revolution and counterrevolu-
tion. Indeed, it was very progressive in methodics for enhancing experi-
mental rigor, and it was also a step forward with regard to the universe of
discourse of psychology for studying all animals, as well as for investigat-
ing the elementary forms of behavior, which had been disdained by men-
talism. But behaviorism was definitely regressive for eliminating the mind
from the purview of psychology, for discouraging theorizing (hence expla-
nation), and for refusing to look into the sources of behavior, namely the
nervous system. This refusal solidified the wall that mentalism had
erected between psychology and biology. And the rejection of the mental-
ist problematics left a vacuum that was promptly filled by pseudoscience.
(Culture abhors a vacuum: The vacua left by science or technology are
filled with junk.) In short, the behaviorist legacy is ambivalent.
However, since the emergence of cognitivism and generative grammar
in the late 1950s, the great merits of behaviorism have tended to be over-
6.6. Summing Up 135

looked, and behaviorism-bashing has become a fashionable intellectual


sport. Thus Chomsky (1959) and other critics (e.g., Davidson, 1974) have
held that behaviorism is not just narrow and shallow, as we have claimed,
but basically inadequate. Not even classical and operant conditioning
have been spared. Not even the hypothesis that we must learn in order to
know anything has been kept. And disembodied mental entities have
proliferated. We submit that this has been an excessive and obscurantist
reaction against the limitations of orthodox behaviorism; it has thrown the
baby out together with the bath water. We take instead the view that, far
from being wrongheaded, behaviorism was insufficient: that it should
have been expanded and deepened.
As a matter of fact, some of the behaviorists themselves realized some
of the severe limitations of radical behaviorism and attempted to over-
come them. Indeed, behaviorism evolved from stimulus-response ethol-
ogy to stimulus-intervening variable-response, to stimulus-hypothetical
construct-response psychology. This liberalization process culminated in
the 1950s, when it should have been clear that there were only three ways
open to the behaviorist movement.
One possibility was to cling, with Skinner, to the declining orthodoxy,
which had become theoretically empty and experimentally boring. Such a
conservative stand was no mere stubborness: There were still lots of
kinds of behavior to be discovered, and behaviorism was just starting to
harvest its most sensational practical fruit in the field of clinical psychol-
ogy, namely behavior therapy (e.g., Wolpe, 1958), which will occupy us
in section 12.1. The second way open to the movement was that taken by
such neobehaviorists as Hull and Tolman. This was to get closer to men-
talism, sometimes dangerously close to the wild conjecturing that was
becoming fashionable in cognitive psychology and in psycholinguistics.
The third possibility for the behaviorist movement was to remain faithful
to the scientific attitude of orthodox behaviorism, while attempting to
widen its problematics and its methodics, and endowing it with a theoreti-
cal nucleus by forming a close alliance with neuroscience. This was the
way first taken by Lashley: that of biopsychology or psychobiology
(Lashley, 1941). This coalescence will be our next concern.
IV

Biopsychology
CHAPTER 7

Neurobiology

Biopsychology, or psychobiology, is the scientific study of behavioral


and mental processes as biological processes, hence as falling within the
province of biology. More precisely, the driving assumption of biopsy-
chology is that the behavior of animals endowed with a nervous system is
controlled by the latter, and that their mental or subjective life, if any, is a
collection of neural processes. This assumption incudes then the strong or
emergentist identity hypothesis that mental events are brain events (sec-
tion 3.1).
The basic assumption of biopsychology has acted as a powerful heuris-
tic guide and it has never been refuted. However, being so sweeping, it
will never be fully confirmed for all types of behavior and all kinds of
mental process. It is in the same boat with other sweeping scientific
hypotheses, such as those of (local) energy conservation, and of evolution
by genic recombination and mutation and natural selection. The hypothe-
sis is, if you will, an act of faith. But an act offaith that, farfrom chaining
us to archaic and barren dogmas, has sparked one ofthe most exciting and
fruitful research projects of all times: that of explaining behavior and
subjective experience in a scientific fashion.
As we saw in the two preceding chapters, mentalism and behaviorism
are inadequate for refusing to investigate the nervous system in order to
find out how it controls behavior and does the minding. From a biological
viewpoint, any study of vertebrate behavior or mentation that overlooks
the head is wrongheaded. Still, there is no denying that both mentalism
and behaviorism have made valuable contributions to our knowledge of
behavior and mind. Only, their lasting findings have covered a narrow
ground; they have been confined to the molar level and they have been
purely descriptive. Let us explain.
In every science dealing with systems we are forced to account for
certain events in terms of two or more levels: one of wholes or systems,
the other of parts or components. For example, we describe evaporation
as the conversion of a liquid into a gas, and explain it as the molar
outcome of the escape of the more energetic molecules from the liquid
140 7. Neurobiology

body. We describe a given chemical reaction as a transformation of sub-


stances, and explain it as the cumulative effect of molecular rearrange-
ments. Likewise, in psychology we describe vision as an internal repre-
sentation of things in the external world, and attempt to explain it as the
specific activity of the visual system. And we describe speech as a form of
behavior that conveys information, explaining it as an activity of the
speech centers in the brain and the vocal tract. In all such cases we link
levels, and we explain in terms of components and their interactions what
we had first described in terms of wholes. In physics and chemistry,
explanation may take us down to the level of elementary particles or field
quanta, whereas in psychology it may take us down to the level of synap-
tic boutons and neurotransmitters.
Mentalism and behaviorism investigated behavior and the mental with-
out inquiring into their neural "substrates" or "correlates." Biopsychol-
ogy investigates not only behavior and mentation but also their neural
mechanisms, in an attempt to explain the former. (Remember: No expla-
nation without mechanism, and no mechanism without matter.) Whereas
mentalism and behaviorism are one-level disciplines, and more precisely
molar ones, biopsychology is a discipline that studies processes on sev-
erallevels: molecular, subcellular (e.g., synaptic), and cellular, as well as
those of the neural system (e.g., the hippocampus), the brain, the neuro-
endocrine system, and the whole organism. And, whereas mentalism and
behaviorism are at best descriptive, biopsychology is explanatory as well,
at least when it succeeds in uncovering neural mechanisms.
The difference between pre biological and biological psychology is par-
allel to that between classical and quantum mechanics, or between classi-
cal and molecular genetics. The similarity concerns not only the matters
of levels of organization and of explanation versus description. It extends
to the need to keep molar research going, though of course now with the
help of some of the ideas and methods of biology. Just as we have not
exhausted the macrophysical study of extended bodies, such as viscous
fluids and plastics, so we may never exhaust the molar study of behavior
and mentation. This study is interesting in itself as well as being a source
of new problems for biopsychological research. (Think of such problems
as uncovering the neural mechanisms of sadness, doubt, and moral con-
flict.) The path from molar description to neurophysiological explanation
is sometimes called the top-down strategy, whereas the converse path,
from neural systems to behavior or mentation, is called the bottom-up
strategy. These two strategies are not mutually exclusive but complemen-
tary.
Most university curricula give the impression that biopsychology is just
one of the many branches of psychology, along with psychophysics, cog-
nitive psychology, developmental psychology, and others. This impres-
sion is mistaken, for all psychological problems can profitably be
approached in a biological fashion. This does not entail that
7.1. Brain & Co. 141

biopsychologists are out to dislodge workers who are not particularly


knowledgeable in neuroscience; it only suggests that their work will be
most useful if placed in the wider context of biopsychology. We must
discover and describe psychological phenomena before we endeavor to
explain them in biological terms. The point, then, is to work toward a
merger of the various branches of psychology on the understanding that
ultimately all behavioral and mental processes must be explained in bio-
logical terms because they happen to be biological processes. As a matter
offact, this is how a few highly successful teams, such as those at McGill
University and the University of Oxford, have been working for the past
few decades, namely by combining the ideas and methods of psychology
with those of physiology. (More on integration in section 13.2.)
So much for advertising copy. Let us now peek at the most intricate,
intriguing, and least well known of all biosystems: the central nervous
system, or CNS for short. We shall focus our attention on the CNS ofthe
higher vertebrates (i.e., mammals and birds). And we shall stress some of
its peculiar emergent properties, particularly its plasticity. We shall also
make some observations about its ontogeny (development) and phylog-
eny (evolution). Finally we shall describe in outline the job of identifying
the neural systems that perform behavioral or mental functions.

7.1 Brain & Co.


The brain is the most complex and interesting component of the body. It
controls or influences many visceral functions and, via the peripheral
nervous system, most of behavior as well; and it does all the feeling,
knowing, willing, and much more besides. Therefore the brain deserves to
be studied by anyone seriously interested in behavior or in mind. (See
e.g., Kandel & Schwartz, 1981.) However, a description of the brain
would be beyond the scope ofthis book as well as way above this writer's
head. We shall therefore focus our attention on a handful offacts relevant
to certain philosophico-scientific issues.
Although comparatively small, the brain is one of the biggest energy
consumers in the entire body. The human cortex alone, a sheet 2.5 mm
thick and the size of a scarf, weighs less than 1.5% of the body total
weight, consumes more than 15% of the body's total energy. This datum
should suffice to cast doubts on the tenet that mental work, unlike muscle
labor, must be immaterial and describable in terms of information rather
than energy.
Brain tissue is composed of neurons and glial cells, the latter probably
working essentially in the "service" ofthe former. Microscopic examina-
tion fails to exhibit two exactly identical neurons. But, of course, neurons
can be grouped by similarity. Grouping by shape results in classes such as
pyramidal, double bouquet, basket, and chandelier neurons. A physiolog-
ical classification results in excitatory and inhibitory neurons. (The mere
142 7. Neurobiology

existence of inhibitory neurons disqualifies the input-output models, typi-


cal of behaviorism and information-processing psychology, which pivot
around excitation while ignoring inhibition.) Finally, a biopsychological
(or functional) classification yields classes of neurons with very special
abilities, such as detecting different colors or lines of different inclinations
(i.e., feature detectors).
Far from being a homogenous body honoring the holistic metaphysics,
the brain is composed of a number of anatomically distinct molar subsys-
tems, such as the brain stem, the hypothalamus, the thalamus, and the
cortex. The latter in turn is divided horizontally into different "areas"
and vertically into well-differentiated layers-an average of six in the
case of humans.
Each of the large brain subsystems is composed of smaller multi-
neuronal components. In particular, the primate cortex is composed of
columns, which are in turn composed of minicolumns. There are an esti-
mated 600 million minicolumns in the cortex, everyone of which is com-
posed of about 110 neurons (except in the striate cortex, where the num-
ber is 260). A mini column works as a whole and it is related to neighboring
cortical columns as well as to subcortical systems such as the thalamus
and the limbic system. Each column may discharge functions different
from those of the adjacent columns, and receives stimulation indepen-
dently from its neighbors. This allows it "to do its own thing" while at the
same time participating in a collective or mass activity. (See e.g., Schmitt,
Worden, Adelman, & Dennis, 1981).
The old view was that the brain could be activated only by external
stimulation. The recording of single cell activity showed long ago that this
static picture of the brain is false: that the normal state of every neuron,
be it excitatory or inhibitory, is one of activity. Moreover, neurons can
fire spontaneously; that is, they can produce an output without an exter-
nal cause. The effect of external stimulation is not to initiate brain activity
but to change its pattern, for instance, by altering the connections among
neurons. Such neuronal spontaneity suffices to refute the dogma that the
brain is a Turing machine, for the latter can change states only under the
action of an input. (In fact, one of the axioms of the theory of Turing
machines is "T(s ,0) = s," where T is the next state function, s an arbitrary
machine state, and 0 the null input.) Neuronal spontaneity also refutes the
philosophical dogma that all changes require external causes-one of the
tenets inherent in behaviorism.
Brain activity is of many kinds, from metabolism-involving the syn-
thesis and breakdown of biomolecules of several kinds-to intra- and
intercellular communication, to the sprouting and pruning of dendrites
and synaptic boutons. Let us recall a few facts involving such activities.
Blood flow, being the conveyor belt of the inputs and outputs of meta-
bolic processes, is a faithful indicator of metabolism. In turn, cerebral
7.1. Brain & Co. 143

blood flow can be imaged, e.g., by having the subject inhale radioactive
xenon. An increase (decrease) in the level of radioactivity indicates an
increase (decrease) of overall activity in the brain region that is being
studied. In this way it has been found, among other things, that the
forebrain (or executive center) of a subject sitting with eyes closed and
doing nothing, is between 20% and 30% more active than the average of
the entire brain. Using the same technique it has also been shown that,
when faced with tough problems, nearly all of the regions of the brain
become equally active; the cognitive centers muster all the support they
can get. (The matter of localization of functions will be taken up in sec-
tion 7.5.)
There are dozens of interneuronal "messengers," from light atoms
such as sodium and calcium to heavy molecules such as serotonin and
dopamine, and even heavier ones such as the brain peptides. Calcium is
the Hermes of them all; it is light and swift, it activates nearly everything
it touches, and-unlike the heavier neurotransmitters-it can penetrate
cellular membranes. (Actually Ca++ can do all this.) In particular, calcium
activates enzymes and, because of its capacity to cross neuron mem-
branes, it acts as a mediator between neurotransmitter molecules and
target cells. If calcium is added, a number of processes occur; if put out
of action (by adding substances that combine with it), neural activity
decreases. Moral: If you wish to keep a sound mind, watch your Ca
diet.
As for anatomical changes in the brain, the most spectacular recent
findings are those concerning dendritic branching and axon regeneration.
Modern methods of imaging in vivo have made it possible to observe
directly, and even to film, some of the changes in neuronal connectivity
that had been hypothesized to explain learning and other facets of mind.
For example, Purves and Hadley (1985) succeeded in visualizing the same
neurons in young adult mice at intervals of 3 to 33 days. They found that
some dendritic branches formed de novo, others elongated, and still
others retracted over intervals of two weeks or more beyond the develop-
ment period.
As for neuron repair, it has been known for a long time that peripheral
nerves can regenerate, but it had been assumed that injured cells in the
mammalian CNS could not regenerate. Ramon y Cajal doubted this belief
and put it to test. He inserted segments of peripheral nerves into the brain
and observed that they become temporarily innervated. He even ex-
pressed the hope that experimental neurology might one day remedy
artificially the deficiencies caused in the eNS by injury or disease. Recent
experimental work has strongly confirmed Cajal's conjecture and given a
basis to his hope. In fact, Aguayo (1985) and others have shown that
injured neurons of certain kinds in the mammalian CNS can elongate their
axons when peripheral nerve segments are grafted; those neurons grow
144 7. Neurobiology

through the graft toward the target. The in vitro manufacture of living
neural prostheses is getting close. (However, see Rakic, 1985.)
The permanent activity of the brain on various levels is relevant to the
question of whether humans are naturally active or passive. This question
can only be decided scientifically, but it is also philosophically and ideo-
logically interesting. In fact, if humans are naturally passive, then envi-
ronmentalism-one of the assumptions of behaviorism (section 6.2)-
cannot account for the phenomena of boredom, impatience, and
rebellion. And if the majority are really passive, then racism and classism
become vindicated in their claim that progress calls for the domination of
the many by the few.
There is plenty of behavioral evidence against the natural passivity
thesis. Healthy animals are naturally active. This explains our dislike for
monotonous tasks, eagerness for novelty (in moderation), and the various
ways in which humans and other animals manage to keep busy even when
they are not ordered to work (e.g., by playing, strolling, tinkering, and
daydreaming). Such behavioral evidence is explained by the permanent
activity of the brain. Only the sick or undernourished brain may never get
bored or impatient; the healthy brain seeks stimulation and action.
Brain dynamism is due not only to the spontaneous activity of the
neurons but also to the interactions between the brain and other body
systems, such as the endocrine one. The latter synthesizes some of the
molecules used by the brain, which in tum stimulates or inhibits the
endocrine glands. Something similar holds for other body systems; the
brain is distinguishable but inseparable from them. It acts upon, and is
influenced by, the rest of the body. In particular, the homeostatic regula-
tion of the milieu interieur-without which, as Claude Bernard noted, we
would be totally at the mercy of the environment-takes a lot of brain
work. In sum, brain activity can only be understood by regarding the
brain as a subsystem of an animal immersed in its environment. (By the
way, this refutes the strong modularity thesis discussed in section 5.2.)
All this now sounds trite, but every time an interaction between the
brain and some other body subsystem is discovered, it comes as a great
surprise. After all, sectorial or nonsystemic thinking, fostered by hyper-
specialization, usually prevails. Two recent episodes will drive home this
point. The first concerns what used to be called 'visceral learning. ' The
so-called autonomic nervous system is in charge of such functions as the
regulation of breathing and the heartbeat. That system was assumed to
function independently of the brain, so that an animal could not possibly
learn to slow down breathing or metabolizing. It has now become routine
laboratory practice to condition a number of animals to do precisely this,
though not exactly in the conditions described in the early publications.
(See Dworkin & Miller, 1986.) Yogis, who had previously been regarded
with suspicion, were tested and found to be able to teach themselves to
7.1. Brain & Co. 145

lower their own metabolic and heartbeat rates. The popular explanation
was of course that this was evidence for the power of mind over matter.
The fact that similar results were attained with rats and pigeons, without
any obvious intervention of consciousness and willpower, refuted the
popular explanation. In any event, the moral is clear: No subsystem of the
body is fully autonomous.
The second finding is of similar philosophical import. It concerns the
action of higher brain functions on the immune system. It is an old tenet of
folk medicine that changes in mood affect sensitivity to pathogens and, in
general, resistance to illness. For example, recently bereaved people are
more likely to get sick than others, cancer patients deteriorate more rap-
idly if they are pessimistic, and generally a decreased "will to live"
lowers the natural (immune) defenses. By the same token, "positive"
feeling and thinking help healing. When Harvard paleontologist Stephen
Jay Gould learned that he was suffering from a rare incurable cancer, he
asked Nobel laureate Sir Peter Medawar what the best prescription was.
His answer: "A sanguine personality. "
Such cases were hardly investigated because they seemed to be nothing
but old wive's tales in support of the myth that mind can move matter.
Most scientists tend to believe that the mental is an epiphenomenon that
cannot react back on the somatic. Only those who adopt explicitly the
psychophysical identity hypothesis in either of its forms understand that
mind can move matter because it is material or, rather, a process in a
material thing. In any event, recent work has altered the opinion accord-
ing to which illness will follow its course regardless of the way one feels or
thinks about it. (See Locke & Hornig-Rohan, 1983 for a bibliography of
neuroimmunological research.)
Indeed, it has been found that there is an intimate coupling between the
nervous system and the immune system. This coupling is effected not
only by products of the immune system, such as interferon and inter-
leukin, that affect neurons. It is also effected by nerves that control the
thymus gland (where lymphocytes mature), the lymph nodes, the spleen,
and even the bone marrow. (Apparently these innervations had escaped
earlier anatomists. Perhaps they were not looking for them.)
One of the most spectacular experimental findings on the action of the
nervous system upon the immune one is this: Immune responses in rats
can be altered by conditioning to the point of extinction. Thus, rats given
cyclophosphamide, which causes nausea and also suppresses immune
responses, were fed a saccharin solution. Predictably, they quickly
learned to avoid the latter. After a while they resumed drinking saccha-
rine solution-an indicator that they were forgetting their taste aversion
experience. But as they did so, unexpectedly they started to die at an
unusually high rate. The "conclusion" (hypothesis) is that they had be-
come conditioned to suppressing their own immune responses (Ader &
146 7. Neurobiology

Cohen, 1985). Such experiments as this help explain clinical findings such
as the one that stress produces immune suppression, and that the recently
widowed exhibit lower resistance to sickness. In the cases of stress the
mechanism seems to be this: Hypothalamus ~ pituitary ~ adrenal
glands ~ immune system. In the case of bereavement the pathway is
bound to start higher up, namely in the cortico-limbic system. In any
event the so-called psychosomatic disorders involve the brain, not the
mythical immaterial psyche. (See e.g., Fox & Newberry, 1984.)
In emphasizing the interactions between the brain and other bodily
organs we should avoid the holistic fallacy that, because the brain cannot
work in isolation, it cannot be said to be the organ of mind. The fallacy
inheres in the neuromuscular model of mental activities, which involves
Watson's thesis that thinking is silent speech (McGuigan, 1978). The fact
that some mental processes have muscular inputs or outputs does not
make them into neuromuscular processes. Likewise, the fact that all men-
tal processes depend on support systems, such as the cardiovascular and
the gastrointestinal ones, does not entail that mentating is a neurocar-
diovascular or a neurogastrointestinal process. There is such thing as the
specific function or activity of an organ (i.e., that which the given organ,
and none other, can perform-though of course not in isolation from all
other bodily organs). The body is a system composed of numerous sub-
systems-among them the brain-and everyone of them has its specific
function(s).

7.2 Plasticity
The boundless variety of human experience and creativity seems to con-
tradict the psychoneural identity hypothesis. After all, the number of
neurons in the human brain, though huge (roughly 10 12), is finite; it would
seem that there are not enough of them "to do all that." Although this
argument has been repeated ad nauseam, it is invalid because it rests on a
false presupposition. This tacit assumption is that neurons, like the ele-
mentary components of a switching circuit or a computer, can be either
on (firing) or off (not firing), and consequently can combine with one
another in a finite number of ways. True, a system composed of two
switches connected in parallel can be in any of only four different states-
if the infinitely many transient states are neglected. But a system of two
neurons connected in parallel or in series can be in any of a nondenumera-
hie infinity of states, because the synaptic junctions can change in a con-
tinuous manner, either spontaneously or under the action of stimuli. Let
us have a quick look at this matter.
Any neuron in a neuronal system can make contact with its neighbors
through a thousand synaptic junctions. The human brain contains at least
10 14 synapses. The collection of connections of a system of neurons is
called its connectivity, or also its wiring, by analogy with the manner in
7.2. Plasticity 147

which the components of an electrical system are connected. We shall call


Cmn (I) the strength of the connection between neurons m and n at time t.
'*
(In general Cmn (t) Cnm (t).) That strength varies in the course of time,
and it is 0 only if the two neurons become disconnected from one another.
The connectivity of a system composed of N neurons can be represented
by the N x N matrix II Cmn (t) II, every element of which can in principle
vary in the course of time. A rough global measure ofthe connectivity of a
neural system at time t is the fraction of non-zero off-diagonal elements of
its connectivity matrix. This number may be called the connectance (a
term used in mathematical ecology). The stronger the connectance of a
neural net, the more the latter acts as a single unit.
Two kinds of neuronal connectivity or wiring are distinguished with
regard to modifiability: hard and soft. Hard-wired neuronal systems are
assumed to be genetically controlled and impregnable to experience.
They can be momentarily excited, but this excitation falls off quickly and
leaves no trace. On the other hand soft-wired systems are supposed to be
sensitive to experience and, in particular, capable of learning from it.
Actually, the hard-soft dichotomy is as simplistic as the on-off one, for
all neuronal systems are soft-wired to some extent. In other words, the
strength of all synaptic junctions changes in the course of time, some
more and others less. Even the invertebrate neuronal systems, the hard-
est of all, are not totally rigid. But of course the higher vertebrate brain is
the system with the greatest capacity to modify and even reorganize itself.
So, the hard-soft dichotomy is a (very) rough approximation.
Depending on the kind of synaptic junction, a suitable stimulus, such as
a burst of electromagnetic impulses, will cause either a short-term or a
long-term change in synaptic effectiveness. In the first case we shall speak
of elastic, in the second of plastic synaptic junctions. In turn, a plastic
synaptic junction may be excitatory or inhibitory according to whether the
long-term change consists in a strengthening or a weakening of the con-
nection. In short, we propose the following partition:

Elastic (short-term changes) or E

Synaptic iUDCtions( /Excitator y or L

Plastic (long-term changeS)\

Inhibitory or H

A burst of impulses acting on a synaptic junction of type E (elastic)


causes a brief excitation; the latter decays quickly leaving no trace. If the
same stimulus acts on a synaptic junction of type H, it causes long-term
inhibition or weakening of the original synaptic strength. Finally, a stimu-
lus acting on a synaptic junction of type L causes a long-term excitation or
148 7. Neurobiology

potentiation; there is a long-term strengthening of the initial synaptic


effectiveness. This potentiation is particularly strong in response to
successive high-frequency bursts applied at intervals of 200 millisec-
onds: it seems that the first burst "primes" the cell (Larson & Lynch,
1986). Such long-term changes occur both in vivo and in vitro. See
Figure 7.1.
We conjecture that the synaptic junctions of type H constitute the
neural mechanisms of habituation or adaptation. If a sea slug is poked
with a stick, it withdraws at first; but, as the stimulation is repeated, the
response weakens until it becomes extinct. We may think of this behavior
as a manifestation of the blocking or inhibition of certain inherited neural
pathways. This kind of "learning" might be described by a nativist as the
forgetting of some items of innate knowledge. We prefer not to include
habituation in learning, for it is far too basic, pervasive, and negative a
type of behavior to qualify as learning. We prefer to restrict the word
'learning' to denote the capacity to perform new behavioral or mental
tasks, not just to suppress inherited reflexes. In other words, we adopt
Hebb's hypothesis, that learning is a lasting strengthening of synaptic

TIME t
H

FIGURE 7.1. Three types of synaptic junction: elastic E (retains no trace of stimu-
lus); excitatory plastic L (or learning); and inhibitory plastic H (or habituation).
After stimulation, E goes back to its initial state, L is strengthened, and H is
weakened. The three cases are covered by the following equation for the rate of
change S of the synaptic strength S under the action of a stimulus of intensity e:
T S+ S - ae = 0,

where T is a time constant, and a a real number. For a type E synapse, a = 0; for
L, a > 0; and for H, a < O. If a constant stimulus is applied suddenly at t = 0 (i.e.,
if e(t) = M(t), where b > 0), the synaptic strength decays exponentially:
Set) = [S(O) + ab le- tlr + ab [V(t) - 1],
where U is the unitary step function. If the inertia lIT of the synapse is small, the
asymptotic value S(O) + ab is reached promptly. In the figure we have assumed
S(O) = o.
7.2. Plasticity 149

junctions. Shorter: Only cell assemblies held together by type L synaptic


junctions can learn.
It is possible that the nervous system of the newborn animal, regardless
of its level on the phyletic scale, has synaptic junctions of all three kinds.
In particular, the visual system of the newborn fly is so plastic that, if kept
in continual darkness, or in a lighted but unpatterned environment, the
animal is never able to discriminate visual patterns (Mimura, 1986). Thus
learning may occur in the early stages of development of all animals, even
invertebrates. But after such a critical period is completed, plasticity is
lost in all but the higher vertebrates; the neural network becomes "hard-
wired," which is the neurophysiological way of saying that an old dog
cannot learn new tricks. The animal has become set in its ways and views,
if any.
Apparently adult invertebrates and lower vertebrates have only synap-
tic junctions of types E and H. Consequently they can habituate, or get
used to not doing certain things, but they may not learn anything new.
(See Hoyle, 1976; Kandel, 1976; Young, 1973.) Thus, when a honey bee
identifies a food source and subsequently revisits it by flying in a straight
line to it, it has not learned its way but has just blocked the neural circuits
controlling flight in alternative directions. Not having learned to perform a
task, the bee cannot be said to possess knowledge of it.
But of course the bee has memory. Experience has left a lasting trace in
its brain and, more precisely, in its tiny mushroom bodies. It is even
possible that such memory is a sort of "cognitive map" of the animal's
territory (Gould, 1986). Moreover, such memory can be transplanted to
other animals of the same species by substituting the mushroom bodies
(Martin, Martin, & Lindauer, 1979). In short, the bee can habituate and
remember, but not learn; it remains forever ignorant. The same can be
said of all the other invertebrates, even of the smart octopus. (Caveat:
Most ethologists and psychologists are likely to disagree.)
The higher vertebrate nervous system seems to be the only one contain-
ing synaptic junctions of all three kinds, E, H, and L, during the entire life
span of the animal. In particular, our own peripheral nervous system
contains H synaptic junctions (i.e., where use decreases the effectiveness
of stimulation). This explains our quickly getting use to the commonplace.
For example, we notice when we grasp a pen or let go of it, but hardly
notice the pressure we apply while holding it. So, we do habituate all the
time. But in addition to possessing junctions of types E and H, the higher
vertebrate CNS contains L (learning) synaptic junctions. Hence only the
higher vertebrates can be said to have the ability to learn and therefore to
gain and lose knowledge (in the sense of the collection of learned items).
We have taken it for granted that the key to habituation and learning is
plasticity. This hypothesis will be explored in the next two chapters. For
the time being let us note the ambiguity of the word 'plasticity' and the
various ways in which something can be said to be plastic. In fact, we
150 7. Neurobiology

must distinguish at least three different though related kinds, hence con-
cepts, of plasticity in our field of inquiry: neural, functional, and behav-
ioral.
Neural plasticity is the ability to modify the strength of synaptic junc-
tions, be it by changing the rate at which neurotransmitters are released
into the synaptic clefts, or by budding (or destroying) synaptic boutons, or
by sprouting (or pruning) dendrites. Functional plasticity is the ability to
recover certain functions after stroke, injury, surgery, or infection. And
behavioral plasticity is the ability to either habituate or learn.
In our view neural plasticity, or the ability to change the interneuronal
connectivity, explains (is the mechanism of) both functional and behav-
ioral plasticity; the last two would occur only as a result of massive
neuronal changes. However, the kind of evidence (relevant data) favor-
able to one kind of plasticity may not be the same as for the other, whence
the fundamental oneness of plasticity may not always be recognized. For
example, observing synaptic or dendritic growth, particularly in vitro,
tells us nothing about learning unless the latter occurs in an experimental
group but fails to occur in the corresponding control group. And func-
tional recovery can occur either through morphological changes in large
numbers of neurons, or through' 'rewiring" (reorganization of the neural
"circuitry"), or through reallocation of functions of certain neuronal sys-
tems. Let us say a few words about these three kinds of process.
Synaptogenesis, or the budding of new synaptic boutons, can be spon-
taneous or induced (i.e., endogenous or in response to external stimuli),
and has been observed in vivo as well as in vitro (Baranyi & Feher, 1981;
Bliss, 1979; Bliss & L~mo, 1973; McNaughton, Douglas, & Goddard,
1978). Such observations have refuted once more the obsolete idea that
neurons are just as passive as the elementary components of a computer.
As for induced synaptogenesis, it has been observed to occur as a result
of electrical or chemical stimulation, as well as in response to experimen-
tal lesions (Flohr & Precht, 1981). The nervous system repairs itself not so
much by whole-cell regeneration (which seems to occur only in the pe-
ripheral nerves) as by synaptogenesis and "rewiring" (i.e., qualitative
changes in the neural connectivity).
It may be speculated that synaptogenesis and "rewiring" are some-
times adaptive responses to an adverse alteration of the way in which the
stimuli impinge upon the animal. For example, normally if a person ro-
tates his head to the right, his eyeballs move automatically to the left with
the same velocity, so that his outside world remains stable. This inborn
reflex is called the vestibulo-ocular reflex. (There is no auditory analog
because we do not have earballs. Hence, when rotating the head, the
sound source appears to rotate around us. Beware of appearances.) If
now a subject is made to wear horizontally reversing prism goggles, in the
beginning he continues to rotate his eyeballs in the customary reflexive
way. Consequently he now sees the outer world moving in the direction of
7.3. Development 151

his head-an experience which even subjectivists are unlikely to enjoy.


However, after some time, usually a couple of weeks, the vestibulo-
ocular reflex is considerably reduced, and consequently the apparent
movement of the external world is reduced as well. Sometimes the vesti-
bulo-ocular mechanism is reversed altogether, so that the subject recuper-
ates normal vision; presumably he has suffered a rewiring of his vesti-
bulo-ocular system (Berthoz & Melvill Jones, 1985; Melvill Jones, 1977).
This finding refutes once more the classical hypothesis that all reflexes are
processes in hard-wired neural systems. Hardness is a matter of degree.
As for functional reallocation, it has been defined as the use of nervous
tissue for non normal functions. A remarkable finding is that unilateral eye
removal in neonatal hamsters causes an increase in the number of somato-
sensory units in the contralateral superior colliculus. Such functional real-
location, though far more frequent in young animals than in adults, occurs
in the latter as well. (For a review see Burnstine, Greenough, & Tees,
1984.) It might well be the mechanism of some of the remarkable recover-
ies seen daily in the neurological clinic.
To conclude, the higher vertebrate nervous system has certain proper-
ties that distinguish it from all the other organs. One of them is the wide-
spread occurrence of inhibitory components: short-axon neurons, neuro-
transmitters such as GABA, and inhibitory synaptic junctions. As a
result, the effects of stimulation are greatly diminished or contained in-
stead of propagating throughout the body like the ripples generated by a
stone thrown into a pond. This is the phenomenon of lateral inhibition,
discovered by Mach and studied intensively by von Bekesy, but totally
ignored by S-R psychology. Another peculiarity of the higher vertebrate
brain is its self-organizing ability, unthinkable in any other organs. Neu-
rons can bud or nip synaptic boutons, sprout or prune dendrites, and
increase or decrease the release of neurotransmitters. All such elemen-
tary processes, some spontaneous and others induced by external stimuli,
keep changing the connectivity of the brain. Consequently any given
stimulus rarely elicits the same response, and neural systems acquire and
lose global (systemic) properties. The occurrence of such events of emer-
gence and submergence is particularly notable in the course of the individ-
ual's development. But this subject deserves a new section.

7.3 Development
Development is the process that starts at conception and goes through
stages until adulthood and beyond. From a neurobiological viewpoint,
development is a process of maturation and reorganization of the nervous
system, particularly the brain. In the early stages neurons grow in size
and number, and at times millions of them die and are replaced by others.
In all stages they change shape through variations in the numbers of
152 7. Neurobiology

synaptic boutons and dendrites, and they make new connections and cut
off old ones. These processes are manifested in the successive acquisition
and loss of abilities and skills.
U ntiI recently, development was defined as the process that starts at
birth and ends on reaching adulthood. This convention presupposed that
the embryological process is automatic (i.e., regulated exclusively by
genes). In recent years we have learned that the chemical environment
exerts a powerful influence on the fetus: Witness the deficits exhibited by
children of drug-using mothers. The intrauterine environment, particu-
larly the sex hormones and toxic substances in it, is so important that it
may account for some cerebrallateralization patterns (Geschwind & Ga-
laburda, 1985).
As for the length of the development process, it has recently been
shown that the human brain continues to be plastic throughout life. There
are parallel processes of growth and involution at all ages-barring senil-
ity. In particular, whereas the number of dendrites of some neurons in-
creases, that of others decreases. And animal experiments confirmed
what could have been suspected from human learning, namely that an
enriched environment favors dendritic growth. In short, development
spans the entire life of the higher vertebrate.
One tends to say that the neuronal systems "subserving" reflex or
automatic activities, such as the visceral functions, are prewired, or es-
tablished once and for all from birth. This is certainly the case with the
neuronal systems of simple invertebrates, but it does not apply to verte-
brates. In the latter the development of the nervous system does not
follow a genetically predetermined path; instead, it is a process of rear-
rangement of synaptic connections determined by internal and external
factors. The decisive internal factor seems to be interneuronal competi-
tion; neurons compete for synaptic contacts, and only those that establish
a minimum number of contacts survive (Rager, 1981). The other factor is
external; the rearrangement of contacts is influenced at first by the intrau-
terine chemical milieu, and after birth by environmental inputs. In short,
it is not true that each axon "knows" beforehand in which direction it has
to grow, and that each synapse "knows" where it has to form (Easter,
Purves, Rakic, & Spitzer, 1985).
As the brain develops, a number of new faculties or skills emerge while
others submerge. For example, as infants grow their motor coordination
improves, whereas their prehensile reflex becomes all but extinct. But of
course on the whole the period between birth and the end of adolescence
is the one of fastest behavioral and cognitive growth-environmental
circumstances permitting. The human neonate is a vegetative, motor,
affective, sensory, and precognitive animal. Neonates detect a number of
stimuli but they may not be able to perceive ("interpret") them. They are
aware of a variety of external and internal stimuli but not conscious of
anything. And, pace Freud and lung, they are not born with images,
7.3. Development 153

concepts, or symbols. But by the time they are four months old human
infants have learned to perceive, locate, and discriminate objects of vari-
ous kinds, from toy and food to human faces. Cognitive development, an
aspect of neural development, proceeds from lower to higher and, in
particular, from iconic to abstract and from specific to general ideas.
However, if the nervous system and its maturation are lost sight of, it may
be possible to interpret the results of certain experiments on infants as
confirming the view that development goes from higher to lower-as
suggested by Bower (1974).
The developing brain is remarkably plastic, as suggested by the pace of
learning and as shown by some experiments. For example, if frontal
lesions are performed on adult monkeys, they exhibit a permanent and
serious deficit in performing delayed response tasks. But if the same
lesions are performed on young monkeys, no impairments result (Gold-
man 1971). This finding has been confirmed by operating on monkey
fetuses. Postmortem anatomical examination shows that unilateral abla-
tions in utero are followed by radical reorganizations of the nerve fibers.
In particular, the callosal fibers deprived of their normal targets in the
contralateral hemisphere become redistributed, so that a shift of function
occurs (Goldman-Rakic, 1982). A last example of plasticity is this. It
appears that we learn to speak with the two hemispheres, lateralization
(usually in the left hemisphere) being a rather slow process that may not
be completed until the age of about to. In fact, hemispherectomy per-
formed on children under 10 years old does not seem to impair severely
their speech; and, when impaired, the right hemisphere can gradually take
over the language functions. But beyond that age the right hemisphere
seems to be mute in most cases.
However, there are limits to plasticity. In fact, many psychologists
have suggested that there are critical periods (i.e., periods comprised
between definite ages for each animal species), during which learning of
certain kinds is possible (or at least fastest), and before or after which it is
impossible (or at least slowest). Language learning is a case in point.
Foreign languages can hardly be mastered if studied beyond adolescence.
However, behavioral observation or even experiment is inconclusive, for
it may be interpreted in alternative ways.
Only a biopsychological study involving an examination of nervous
tissue may be able to decide whether or not certain qualitative changes do
occur at certain ages, that make it possible (or impossible) for an animal to
learn something. In this respect the classical experiments of Hubel and
Wiesel (1962) on visual deprivation in kittens are decisive. They reveal
that there is a critical period for learning to see (roughly the first three
months), past which the animal cannot perceive certain objects. There is
an olfactory analog of that experiment (Van der Loos & Woolsey, 1973).
The existence of critical periods supports Piaget's well-known hypothe-
sis that cognitive development is not continuous but proceeds through
154 7. Neurobiology

qualitatively different stages (sensorimotor, preoperational representa-


tion, concrete operations, and formal operations). These stages can be
shortened or lengthened by environmental factors but they cannot be
skipped nor their order inverted. So far so good. Dissent starts when
assessing the impact of the social milieu on cognitive development.
Whereas Piaget held his stages to be biological, hence unavoidable and
cross-cultural, Vygotsky (1978) claims that they are culturally condi-
tioned, whence the higher stages might not emerge at all in children raised
in culturally deprived environments. In fact, his own field research, as
well as that of his disciple Luria (1976), confirmed this view. (See also
Scribner & Cole, 1981.)
This controversy belongs to the old nature-nurture issue discussed in
section 6.2. There we saw that it is impossible to take sides for either the
nativist or the environmentalist, because each holds a part of the whole
truth. The behavioral and biopsychological evidence supports the hypoth-
eses that there are critical periods, and consequently qualitatively differ-
ent stages as well. And the evidence coming from social psychology and
personality studies shows that the social environment is so strong that it
can thwart development altogether or, on the contrary, get good results
from poor genetic material. In short, we must strike a balance between
nativism and environmentalism. The reason is simple: Although the brain
has a dynamics of its own, it does not exist in a vacuum, but in a natural
and social environment that stimulates its development in some regards
while inhibiting it in others.
One of the most spirited recent controversies on the nature-nurture
issue concerns language learning. Chomsky and his followers embraced
nativism merely because the behaviorist psychology of the 1930s was
incapable of accounting for language acquisition. But the failure of a
particular hypothesis about learning is not a sufficient reason to assume
innateness-as Lehrman (1953, p. 343) notes in criticizing the nativism of
the classical ethologists. In general, arguments from ignorance are in-
valid. On the other hand, the genetics of the brain has much to say about
this issue. In fact, iflanguage, or at least the mythical universal grammar,
were inheritable, then all the systems involved in the formation, utter-
ance, and understanding of speech-or at least the neural ones-should
be inherited in a single block; that is, there should be a single gene, or a
block of interlinked genes, regulating the development of all the many
components of the speech system. But it is a well-known axiom of genet-
ics that the genetic transmission of anatomical components is independent
even if they are intimately connected. Thus, it is possible to inherit
Wernicke's "area," or part of it, from one parent, and Broca's, or part of
it, from another. Such independent inheritability results sometimes in
mismatches that become manifest in speech impairments or idiosyn-
crasies (e.g., good speech comprehension but poor delivery, or con-
versely).
7.4. Evolution 155

But the most sensitive area where the nature-nurture controversy has
been raging from time immemorial is of course that of intellectual ability.
The nativists (e.g., Eysenck, 1971) hold that intelligence is wholly inher-
ited, whereas the environmentalists (e.g., Kamin, 1974) claim that it is
entirely a matter of training. The genetic evidence for nativism is nonex-
istent: (a) no genes controlling intelligence have been discovered, (b) the
gross features of the brain, such as size and the shape of the convolutions,
do not correlate with intelligence, and (c) Burt's celebrated results con-
cerning identical twins reared apart were shown to be fraudulent. How-
ever, this does not entail that intelligence has nothing to do with the
genes-as it should if mind were immaterial. No doubt, genes regulate the
development of the brain, as of every other part of the body, but they do
so in cooperation with the environment. We are not born intelligent, any
more than we are born tall, nimble, or talkative. But we are born with the
genetic potential to become intelligent or stupid, tall or short, nimble or
clumsy, talkative or taciturn, and so forth.
The genetic makeup of an individual fully determines his or her life
history only in the case of serious genetic disorders affecting the organiza-
tion of the brain, such as the Down syndrome. (There are about 3,000
known genetic diseases.) Except in such cases, which are comparatively
infrequent, what will become of a newborn depends as much on his or her
environment as on his or her genes. In sum, we subscribe the report ofthe
ad hoc committee of the Genetics Society of America on genetics, race,
and intelligence, which rejected "doctrinaire environmentalism" and
warned strongly against "the pitfalls of naive hereditarian assumptions"
(Russell, 1976).

7.4 Evolution
The neuroscientist and the psychologist who experiment on nonhuman
primates usually wish to learn from them something about humans. They
assume that apes and even monkeys are adequate "models" or analogs of
man in some respects (e.g., in motor behavior and vision). This assump-
tion is well-founded. It rests on a huge body of observations and experi-
ments showing that, in fact, all primates are kindred. In other words,
experimental work in neuroscience and biopsychology presupposes evo-
lutionary biology and in turn contributes to it even when it does not
engage in tracing family trees.
At first sight evolutionary neurobiology and psychology are impossible
for lack of fossil records of nervous tissue and behavior. Although this
lack is a serious obstacle, it is not an insurmountable one, for we can
"see" evolution by studying homologous systems, and their functions, in
phylogenetically related species with living representatives. One result of
such comparative study is the explanation of "fossil" neuronal systems,
that is, systems left over from earlier stages in evolution and which are
156 7. Neurobiology

now devoid of adaptive value. Another finding is that certain neuronal


systems control nowadays behaviors that differ considerably from the
original ones. (See e.g., Dumont and Robertson, 1986.)
The mere acknowledgement of biological evolution reoriented neuros-
cience and psychology by stimulating animal studies. It opened the door
to comparative or evolutionary ethology and psychology-an early off-
shoot of Darwinism founded by Darwin himself. It disqualified psycho-
neural dualism because, if mind is anything other than a bodily function,
then either it does not evolve or, if it does, its evolutionary mechanisms
are unknown and possibly unknowable. (However, they would have to
match those of biological evolution.) Finally, the adoption of an evolu-
tionary perspective in psychology writes off all anthropomorphic views of
mind, such as panpsychism (e.g., Teilhard de Chardin's), as well as Wat-
son's opinion that thinking is just silent speech-a thesis that makes it
impossible for animals to think. In short, the acknowledgement of evolu-
tion is a watershed in the history of neuroscience and psychology. (See
Jerison, 1973; Masterton, Campbell, Bitterman, & Hotton, 1976a; Mas-
terton, Hodos, & Jerison, 1976b; Roe & Simpson, 1958.)
Evolutionary neurobiology is being studied on all levels: molecule, cell,
organ, and whole organism. Let us give a few examples of each. To begin
with, until recently it was believed that hormones and neurotransmitters
are produced only by highly specialized cells in vertebrates. It is now
known that unicellular organisms, even bacteria, contain both, and even
opioid peptides-as well as ion channels and pumps. (See Schwartz,
1984.) With evolution, cells, in particular neurons, have become more
complex and have specialized, but the biochemical units have been con-
served although they have often changed functions (Roth, LeRoith, Shi-
loach, Rosenzweig, Lesniak and Havrankova, 1982, p. 524).
The nervous and the endocrine systems are the two coordination and
integration systems in vertebrates. (The immune system too is regulative
but not integrative.) These two systems are so strongly coupled together
that they constitute the components of a supersystem that in many re-
gards functions as a whole, namely the neuroendocrine system. (More-
over, the diffuse neuroendocrine system contains cells with properties
common to both neurons and endocrine cells.) This coupling is not acci-
dental but an outcome of evolution. It has been hypothesized that either
of the systems evolved from the other, or that both evolved from a com-
mon ancestral system serving as a single coordination system in more
primitive organisms. The data concerning modern organisms seem to sup-
port equally well the two hypotheses. The psychologist need not take
sides in this controversy, but should keep in mind that an adequate expla-
nation of a number of behavioral and mental phenomena, such as fright
and stress, call for a consideration of the interactions between the ner-
vous and the endocrine systems. And it will help the psychologist to know
something about the evolution of these two.
7.4. Evolution 157

From a biopsychological viewpoint, the main landmarks in evolution


have presumably been the emergence of the first neurons, of the first
neuronal systems, of warm-blooded animals, and of primates. We know
next to nothing about the first two events. However, if modern inverte-
brates are any indication, it is plausible to suppose that the first neurons
and neuronal systems were highly specialized. (The "multipurpose" neu-
rons characteristic of the uncommitted human cerebral cortex are a very
recent development.)
We take it that the emergence of warm-blooded animals was revolution-
ary because, as we know from experiment, the chemical reactions occur-
ring in the synaptic cleft, which are essential for the discharge of mental
functions, slow down and ultimately stop altogether as the body tempera-
ture drops to about 20°e. (In fact, local cooling to this temperature is a
standard technique for stopping mental processes.) A fairly constant tem-
perature of between about 30°C and 40°C may therefore have been crucial
for the emergence of mind. This suggests that the first organisms with
mental functions were mammals and birds, or their immediate ancestors,
the mammal-like and bird-like reptiles respectively. This is why we attrib-
ute mentality exclusively to mammals and birds. However, this hypothe-
sis may still prove to be false.
As we ascend from cold-blooded or lower to warm-blooded or higher
vertebrates, a few general trends appear. One of them is encephalization
(i.e., the increasing importance of the brain as we climb up the phyloge-
netic tree). For example, whereas the frog retina has ",bug detectors," the
higher vertebrate needs a cortical system to discriminate bugs from other
visual stimuli. And, because our brain is extremely plastic, we can learn
to see better, and even relearn to see properly when wearing image-
inverting goggles. (Recall section 7.2.) The same holds for other sensory
modalities and for motor control; the encephalization process is also one
of increasing plasticity, that is, softening of much neural wiring.
Regrettably little is know about the details of the encephalization pro-
cess. The only general and precise hypothesis about the evolution of the
brain is the popular stratification hypothesis. According to it the brain has
evolved by the successive addition of new layers, the newest of which is
the neocortex. Moreover, the older layers would retain essentially their
original functions, and the new arrivals would serve basically the function
of fine tuning the older functions, for instance, to perfect motor control or
vision. They would interact with the older parts but on the whole the
latter would hold sway over the newer ones. This hypothesis is quite
popular; we all have heard about the reptilian mind lurking in our modern
mind and ready to snap at the latest arrival, namely reason.
The stratification hypothesis does not fit in with psychology. It is sim-
ply not true that deep down we feel, perceive, or behave like alligators or
like trout. Furthermore, the geological metaphor ignores the finding that
the organization of the brain, in particular the cortex, is not only by layers
158 7. Neurobiology

but also by vertical columns. It is more likely that, with evolution, the old
systems got reorganized, hence changed their functions in some respects,
and that the newer systems sometimes got the upper hand, as shown by
the fact that we are able to control emotions and correct perceptions.
The difficulties besetting the reconstruction of our family tree become
staggering when it comes to cognitive, linguistic, and moral abilities.
From a biological viewpoint there can be no doubt that these abilities have
evolved from more primitive forms of behavior and mentation, and that
their evolution is only one aspect of the general process of evolution.
Moreover, in some cases we can put forth plausible evolutionary hypoth-
eses. For example, it is likely that all the specialized sensory systems
except for the sense of balance have evolved from primitive tactile sys-
tems, that is, that they are specializations of the skin. And one may
conjecture that human language evolved on the basis of a few rather
general (i.e., cross-specific) neural mechanisms plus a limited set of spe-
cific mechanisms and constraints that differentiate the manner in which
we get to communicate among ourselves (Lieberman, 1984, 1985). How-
ever, we must always keep in mind the speculative nature of such ac-
counts, and guard against two pitfalls.
One pitfall is the rather widespread belief that cognitive development
recapitulates cognitive evolution, so that both our remote ancestors and
the modern primitives feel, think, and act like children. This view is an
application of Haeckel's "law" that ontogeny recapitulates phylogeny. If
this "law" were true we would be able to infer the way our remote
ancestors behaved from watching the behavior of our own children. For
better or worse, the "law," though still popular, has been refuted long
ago. Hence its application to cognitive evolution must be written off, the
more so because we must suppose that the hominids were capable of
doing many things that no young child can do, such as procreating, fash-
ioning tools, and defending themselves.
Another possible mistake is forgetting the social component of human
evolution. Unlike ferns and snails, hominids were gregarious. Conse-
quently, in order to reconstruct our remote past we must use not only the
categories and principles of the synthetic theory of evolution, such as
those of genetic mutation and recombination, natural selection, and sex-
ual competition, but also those that account for social evolution. These
include psychological categories, such as those of empathy, thinking, and
planning, as well as sociological ones, such as work, social organization,
and defense.
We are ignorant of almost everything about the evolution of the brain
and its functions. However, we can no longer ignore that it did occur. The
mere existence of evolutionary biology has changed the way many psy-
chologists view behavior and mind. It has transformed psychology into a
natural science-or nearly so. On the other hand, the disregard for the
evolutionary perspective has a number of negative consequences on the
7.4. Evolution 159

psychological way of thought. A psychologist without an evolutionary


frame of mind is likely to indulge in vitalism, dualism, and misplaced
teleology, whereas a psychologist mindful of evolution will avoid such
mistakes.
For example, psychologists trained to think in evolutionary terms will
not say that we have memory in order to better face the future. Instead,
they will surmise that the animals deficient in memory had poor chances
of surviving, hence of leaving offspring. And they will not say that we
possess consciousness in order to better control our feelings, mentation,
and behavior. Instead, they will suggest that the higher the degree of
consciousness, the better the chances of keeping emotion, thinking, and
acting under control during emergencies, whence the better the chances
of survival.
One last example. Suppose we were to adopt Helmholtz's view that all
perceptions involve unconscious hypotheses and even inferences-a con-
jecture obviously at variance with the evolutionary ranking of mental
abilities. In this case one would be likely to assume, with Gregory (1973),
that there are two radically different kinds of perceptual illusion: physio-
logical and cognitive. The former would consist in neural malfunctioning
(an obviously material condition) whereas the illusions ofthe second kind
would derive from a wrong cognitive strategy-presumably an immaterial
item. Thus disregard for evolution, hence for comparative psychology,
may easily lead to psychoneural dualism-and conversely.
The explicit adoption of an evolutionary perspective is bound to shed
light even on psychological corners that, at first blush, have nothing to do
with evolution. A case in point is personality theory. The differences in
the nervous systems of people are partly inherited and partly acquired.
Hence, if personality is characterized by the animal's behavioral and
mental repertoire, it follows that it too is partly inherited and partly ac-
quired. In an evolutionary perspective the situation is this. Some person-
alities are better adapted than others to cope with environmental de-
mands. Thus some personality traits have, in a given milieu, a better
chance of survival (i.e., of transmission to the progeny) than others. Now,
conceivably, the bands of hominids and primitive men had far fewer
people than the Neolithic villages, and the challenges posed by nature to
the former were more uniform, whereas the social challenges were far less
numerous and exacting. This must have favored the propagation of gener-
alists and conformists, while hampering that of specialists and deviants,
among hominids and primitive men. (This would explain the slowness of
their evolution.) As human societies grew in complexity, they made room
for more differences in personality. Such "variability" has peaked in
modern society (despite our constant complaints about uniformity). And
it contributes to explaining the quick pace of change in modern times.
The adoption of an evolutionary perspective is bound to have an impact
not only on the way behavioral and mental traits are explained, but also
160 7. Neurobiology

on experimental design. For example, mating behavior is usually studied


in either of two ways: behaviorally (in the wild or in the laboratory) or
physiologically. The behavioral study yields data that cry for explana-
tions. The latter can be found by investigating the hormonal and neural
mechanisms, as well as the environmental triggers, of mating behavior.
But the mere existence of such mechanisms poses the problem of their
evolution. Speculation on the latter may in turn suggest hypotheses con-
cerning the ways internal factors combine with physical and social cues to
cause or alter sex behavior (Crews & Moore, 1986). The same evolution-
ary viewpoint suggests making experiments to find out, for example,
whether female mating preference in a given species is random, genetic,
or learned. Indeed, if this preference has an important genetic compo-
nent, then it must be possible to alter it in successive generations by
artificial selection. This is apparently the case with ladybirds (Majerus,
O'Donald, & Weir, 1982). Some researchers claim that in this respect
humans are no better than ladybirds.
Finally, research strategies are bound to be strongly influenced by the
adoption of an evolutionary perspective. Thus, because humans and rats
have common ancestors, rat psychology has much to say about human
nature. On the other hand, because humans and computers have no com-
mon ancestors, information engineering has nothing to say about human
nature. Only an utter disregard for evolutionary biology could make it
possible for anyone to believe that the study of machines could do more
for human psychology than the study of people and their evolutionary
relatives. And only a firm conviction that learning is free from "biological
constraints," and therefore must have been the same from the moment
the first organisms capable of learning emerged, could have led the behav-
iorists to look exclusively for commonalities in the learning patterns
of the various species, and to write off the warning of the ethologists
that different species might learn in different ways (Bitterman, 1975,
1984).

7.5 Functional Localization


The traditional view of the brain is holistic. It holds that the brain is a
homogeneous mash working as a whole. It was only natural to hold such a
view before neuroanatomists showed the brain to be composed of many
anatomically distinct subsystems. It was also natural to distrust localiza-
tionism for having first been proposed as a wild fantasy by F. J. Gall,
who claimed to be able to make personality diagnostics by examining
the bumps on the skull. Besides, holism went well with psychophysi-
cal dualism for, if the various mental faculties are not localized in the
brain, then the mind too had to be conceived of as a unitary entity, and
it seemed vain to try to find the neural "correlates" of the various men-
7.5. Functional Localization 161

tal functions. This explains why, whereas holists have usually been
dualists, most localizationists have been monists of the materialist
variety.
The nineteenth-century holistic tradition is still going strong in neuro-
psychology. There are two arguments in its favor. One is that the normal
brain works synergically when tackling complex tasks-at least when it
does so successfully. A second forte of holism is the remarkable func-
tional recovery some patients make after having sustained severe lesions,
particularly if they are young. However, all this proves is that the brain is
a tightly knit system, some components of which are plastic. It does not
prove the "equipotentiality" of all the regions of the brain, or even of the
cortex, any more than the remarkable integration of the various compo-
nents of a car disproves the principle that everyone of them performs a
specific function. We feel, think, and move as units, just as a car moves as
a whole. But this only shows that, although brains and cars have many
components, these are well coordinated; they are systems, not aggregates
or structureless wholes.
There is plenty of experimental and clinical evidence for the localiza-
tionist hypothesis, and it is mounting steadily. One of the earliest was P.
Broca's discovery in 1861 that strokes or lesions in what is now known as
Broca's area (i.e., the foot of the third frontal convolution of the left
hemisphere), can cause severe impairments of speech delivery. Some
years later C. Wernicke (1874) discovered that speech formation and
understanding is in the charge of what became known as Wernicke's area
(i.e., the posterior part of the first temporal convolution of the left hemi-
sphere). A stroke or injury in this area may cause the subject to process
auditory information 10 times slower than normal, and may even render
the subject incapable of understanding what he or she is being told, or of
uttering meaningful sentences. Lesions in different parts of the speech
areas cause aphasias of different kinds. For example, cerebral atrophy in
a certain area can cause a patient to have serious difficulties with verbs,
although she may retain her ability to handle nouns. There are even
different brain areas for short "connecting" words (articles, pronouns,
and prepositions) and for "content" words (nouns, adjectives, and
verbs). In fact, patients who have sustained lesions in certain parts of the
motor cortex find it difficult to handle words of the former kind but not of
the latter. Almost every issue of the journal Neuropsycho[ogia contains
reports of this kind. In short, neurolinguistics has amply corroborated the
localizationist view.
What holds for speech holds, mutatis mutandis, for one of the most
basic and ancient of all mental functions, namely emotion. In 1927 W. R.
Hess found that he caused rage and attack by electrically stimulating the
hypothalamus of a cat. Later on, J. Flynn found that, by stimulating a
nearby region (the central grey), rage without attack ("sham" rage) was
caused. And in 1954 J. Olds and P. M. Milner discovered that pleasure too
162 7. Neurobiology

is deep in the brain and, more precisely, in the anterior medial hypothala-
mus. Pain is also in the brain even when felt elsewhere.
Another example: Face recognition is the task of a distinct population
of neurons, which in the macaque is contained in the inferior temporal
cortex. The response of those neurons is roughly independent of the
position and size of the stimulus, and not greatly reduced if some of the
components of the latter, such as the eyes, are removed from the image.
On the other hand, the scrambling of the internal features of the face, in
the cubist style, greatly reduces the response. (See e.g., Desimone, Al-
bright, Gross, & Bruce, 1984.)
Recent laboratory and field studies have shown that skill learning and
conscious recollection of the learning process are the jobs of different
memory systems located in different regions of the brain. For example,
amnesics who cannot recall recent events can use some of their skills and
even acquire new ones while being unable to remember the circumstances
of the learning process. (See e.g., Schacter, 1983; Squire, Cohen, & Zou-
zounis, 1984; Squire & Cohen, 1985.) Thus an amnesic golf player in the
early stage of Alzheimer's disease can playa good game but cannot re-
member, after half a minute's delay, where her ball landed. Others can
learn to read mirror-inverted words without recollecting how they learned
this skill; the learning "center" of that skill is different from that of
"episodic memory." Neurological patients of a different type can draw
the things they see but cannot name them. This suggests that their visual
system has become disconnected from their speech areas. (See Ges-
chwind, 1974.)
In short, localizationism has been amply vindicated. However, it will
continue to be challenged, if only because the task of locating mental
functions involves tricky logic. The basic assumption of localizationism is
that every mental process is the specific activity of some subsystem of the
brain. It follows that, ifthe neural system malfunctions or is destroyed, or
is absent from birth, the corresponding function is abnormal or even
nonexistent. (In symbols: If F, then S. Now, not-So Hence, not-F.) How-
ever, if the normal function is absent, it does not follow that we can blame
the corresponding brain system. It may well be the fault of some other
system connected with it, even of a support system such as the cardiovas-
cular one, which may not be supplying the required amount of blood to
the brain. (In symbols: If F, then S. Now, not-F. Nothing follows.) The
absence of normal function-or, in neurological terms, the appearance of
a symptom-is merely an ambiguous indicator of the possibility of a
lesion in the corresponding subsystem of the brain. (For more on this see
section 4.5.)
The experimental and clinical evidence favors localizationism, but the
latter comes in two strengths. The strong, topographical or mosaic locali-
zationist hypothesis is that every behavioral or mental function is dis-
charged by a distinct anatomically concentrated neural system with well-
7.5. Functional Localization 163

defined borders (i.e., a center, nucleus, or "area"). The weak


localizationist hypothesis is that every behavioral or mental function is
performed by some neural system that may be concentrated or distrib-
uted. Note the shift of emphasis from place to system. The first hypothe-
sis implies the second, as A implies A or B.
According to weak localizationism, some of the neural systems dis-
charging behavioral or mental functions could be composed of neurons,
or neuron assemblies, located in different places-provided of course that
such components be held together, if only by thin filaments of nervous
tissue. (See e.g., Squire, 1986.) Moreover, as Hebb (1949) and Bindra
(1976) speculated, it might well be that the membership of some such
systems varies in the course of time; one and the same neuron might
belong now to a given system, later to another. After all, such a possibility
is strongly suggested by the experimental evidence for neural plasticity
(section 7.2). See Figure 7.2. Furthermore, the specific function of some
such systems might consist in the interaction among their components.
For example, it has been suggested that the sleep cycle control system is
such an interacting neuronal population, i.e. that sleep is its specific activ-
ity or function (Hobson, Lydic, & Baghdoyan, 1986).
The experimental evidence available at the time of writing seems insuffi-
cient to choose between the strong and the weak localizationist hypothe-
ses. However, given the dynamic nature of interneuronal connection,
weak localizationism looks more plausible than the strong version. We
may then adopt it for the time being, if only because it is the more cau-
tious of the two. In any event, localizationism, in some form or other, is
in, whereas holism is out. And, whether in its weak or in its strong
version, the triumph of localizationism is also that of systemism.
A priori there is an alternative to both holism and systemism, namely
atomism (section 3.2). Neurobiological and biopsychological atomism is
neuronism, or the view that a single neuron can discharge a behavioral
or mental function. No doubt, some kind of neuronism does work for
simple vertebrates, every neuron or neuronal ganglion of which inner-
vates a given muscle or muscular system. It may also work for some

o
(a) (b)
t,

(c)
t2

FIGURE 7.2. Three possibilities for the localization of a mental function. (a) Con-
centrated system: nucleus, center, or area. (b) Distributed system with constant
composition (i.e., same neurons). (c) Distributed system with varying composi-
tion (i.e., different neurons at different times).
164 7. Neurobiology

functions in the lower vertebrates. A classical example is that of the


neurons in the frog retina that respond singly to the presentation of
insects.
But neuronists have gone further, speculating that at least the neurons
of certain kinds could perform singly certain complex mental functions.
Thus Blakemore (apud Dimond, 1980, p. 26) has claimed that feature
detectors, such as those discovered by Hubel and Wiesel in the visual
cortex, "have knowledge. They have intelligence, for they are able to
estimate the probability of outside events . . . . Neurons present argu-
ments to the brain based on the specific features they detect, arguments
on which the brain constructs its hypotheses of perception." (See also
Barlow, 1972.) However, there is not a shred of evidence for this specula-
tion. On the other hand, there is plenty of evidence for the hypotheses
that most whole human brains are incapable of estimating the probability
of events or of constructing valid arguments. (See e.g., Johnson-Laird &
Wason, 1977.)
One reason for the appeal of neuronism is quite sound, namely reduc-
tionist philosophy, which can actually work a long way-until it hits
emergence. Another reason is less respectable, namely attachment
to a well-established laboratory technique: It is far easier to record the
activity of single neurons than the synergic activity of a group of neu-
rons. In fact, the latter calls for using bundles of electrodes, one for
each cell, as well as the investigation of correlations among their
recordings. However, preliminary results of multineuron experiments
are encouraging. For example, they have shown that certain neuron
assemblies in auditory cortex can discriminate "melodies," e.g., the
tone sequences ABC and BAC, whereas single neurons in the same
region cannot (Gerstein, Aertsen, Bloom, Espinosa, Evanczuk, &
Turner, 1985).
To conclude, the brain is neither like a bowl of broth nor like a gaseous
body. The brain is a supersystem composed of numerous subsystems,
everyone of which is coupled to some other subsystems, but none of
which is directly coupled with all the other brain subsystems. Every brain
subsystem has its specific function, or peculiar activity, in addition to
performing general "household" functions. But it cannot discharge its
specific function in a normal way without the support of a number of other
systems, some nervous (like the brain stem) and others non-nervous (like
the heart, and the endocrine or immune glands). Ultimately, it is the
whole animal that moves and perceives, breathes and digests.
Given the postulated identity of mental function and specific brain func-
tion, the consequence of all that for psychology is obvious. The mind is
neither a single homogeneous block nor a collection of independent mod-
ules a la Gall or Fodor (1983). Instead, the mind is a functional system
(section 5.3), that is, a collection of distinct but interlocked brain pro-
cesses. However, we are trespassing on the next chapter.
7.6. Summing Up 165

7.6 Summing Up
Adoption of the psychoneural identity hypothesis encourages the study of
the nervous system and, in particular, of the brain. More specifically, it
encourages carrying on with the job of "mapping the mind on to the
brain," or identifying the neural systems that perform illental functions
(Olds, 1975). This vast research project calls for the cooperation of
neuroscientists and psychologists, both pure and applied. It involves
studying not only whole organisms but also some of their components,
particularly the neural ones, as well as analyzing the latter on several
levels, even down to that of the molecule. This is so vast and difficult a
project that it is likely to occupy all the generations to come. Working on
it is a much more rewarding occupation than playing the old obscurantist
game of writing about the mysteries of mind.
CHAPTER 8

Basic Functions

The attempt to "correlate" behavior and mind to brain functions is not


new. Hippocrates rejected spiritualism and adopted Alcmaeon's material-
ist hypothesis that the brain is the organ of mind, from which it follows
that mental disorders are brain maladies. The Hippocratic views on this
and other matters were perfected by Galen and remained in force until the
seventeenth century. Thus, in a book published in 1575, that was trans-
lated into several languages and remained in print for more than one
century, the Spanish doctor Juan Huarte conceived of the various mental
faculties as so many functions of different brain subsystems: He was a
localizationist. And when Cardinal Richelieu died in 1642, the team that
performed his autopsy reported that the Cardinal's brain had twice the
normal number of ventricles, which proved that "il y faisait double quan-
tite d'esprit en general." (See Beaulieu, 1983.)
But outside the medical profession dualism was dominant. Most of the
philosophers and psychologists who wrote about behavior or mind had no
use for the brain. At most they admitted that it is the "material basis" of
mind, but they did not bother to elucidate this obscure metaphor. Even
nowadays the majority view in the cognitivist (or information-processing)
school is that the brain is confined to processing sensorimotor informa-
tion, cognition being ajob of the mind, conceived of as a set of immaterial
programs. (Recall section 5.4.)
The first scientific endeavors to map behavior and mind on to the brain
were those of Broca and Wernicke in the past century. However, the first
sustained efforts by entire teams of researchers were not made until
Pavlov initiated his experimental investigations into the physiology of
behavior, and Hebb started to speculate on the physiology of cognition.
The basic hypothesis underlying the entire project is of course the psy-
choneural identity hypothesis (section 1.3).
In this chapter we shall examine some of the basic behavioral and
mental functions, leaving the higher ones for the next. We count move-
ment, feeling, sensation, attention, and memory among the basic func-
tions because all the animals endowed with a nervous system, even inver-
8.1. Movement 167

tebrates, seem to perform them. On the other hand, learning (as distinct
from habituation), perception, and concept formation would seem to be
the privilege of the higher vertebrates. The basic-higher distinction corre-
sponds roughly to the hard-wired-soft-wired one-only roughly because,
as we saw in section 7.2, hardness is a matter of degree. It also corre-
sponds roughly to the inborn-learned, or reflexive-cognitive one. Again,
this distinction is not a dichotomy, for all the higher verebrates can learn
to alter some of their reflexes.
The higher-lower distinction has not been easy to come by, because
there is a natural tendency to explain animal behavior in terms of human-
like purpose and cognition-or else to deny animals all cognitive abilities,
usually on theological grounds. The need to draw the higher-lower dis-
tinction, and to avoid anthropomorphism, was emphasized by the founder
of comparative psychology, one of the disciplines spawned by the Dar-
winian revolution: "In no case may we interpret an action as the outcome
of the exercise of a higher psychical faculty, if it can be interpreted as the
outcome of the exercise of one which stands lower in the psychological
scale" (Lloyd Morgan, 1894, p. 53).

8.1 Movement
Movement is the most basic mode of change and the easiest to observe,
though not always the easiest to explain or predict. Therefore the descrip-
tion of movement, in psychology as well as in physics, is prior to every-
thing else. But in both cases movement is an effect, output, or manifesta-
tion, that can only be explained in terms of mechanisms, be they physical,
chemical, or biological. In other words, although the study of kine-
matics-the description of motion-precedes historically that of dy-
namics, only the latter can explain the former.
In fact, dynamics entails kinematics, so that the latter has no indepen-
dent existence except in the early stage of the history of a discipline. Just
as Newton's dynamics explains the overt behavior of falling bodies, pen-
duli, planets, and other things, so only neuroscience can explain overt
animal or human behavior, for the latter is a manifestation of complex
internal-particularly neuromuscular-processes. For example, the
movement of a cat that lands on its feet after having been dropped upside
down can be filmed, and its main features can be described both verbally
and mathematically. But the explanation of that remarkable feat calls for
an investigation of the cat's neuromuscular system as well as of the dy-
namics of its whole body.
Motor or overt behavior is defined as the observable activity of muscles
and organs (in particular glands) of external secretion. Walking, turning
the head, reaching for a thing, blinking, and weeping are examples of
overt behavior and, as such, processes in need of explanation. It is not
enough to describe an animal's walk; we want to know how he does it and
168 8. Basic Functions

whether he does it automatically or voluntarily. A satisfactory explana-


tion of the neuromuscular mechanism of overt behavior has been found
only for some simple invertebrates. For example, the motoneurons of the
sea slug Aplysia have been identified, and its basic reflexes, including the
avoidance reaction and its extinction (i.e., habituation), have been un-
veiled. (See Kandel, 1976.)
As we climb up the evolutionary tree the task of accounting for behav-
ior becomes ever more complex, hence more difficult. For example, hu-
man walking involves all the reflex mechanisms found in some lower
animals plus affective, cognitive, and volitive processes. You have de-
cided to attend a certain lecture because the subject or the lecturer inter-
ests you for some intellectual or sentimental motive. In order to reach the
lecture room you gather some bits of information that you plug into your
spatiotemporal map. And you use this map, as well as some clues that you
find and recognize along the way, to guide your walking. Thus even a
banal activity such as walking enlists in humans not only the motor sys-
tem but also the limbic system, some or even all of the sensory systems
and their secondary areas, the cognitive system, and the so-called execu-
tive centers located in the frontal lobes.
A physiological account of motor behavior involves, among many other
items, a redefinition of the key concepts of stimulus and response. A
possible redefinition, for animals endowed with a nervous system, is this.
An external stimulus s acting on an animal b is an event in the environ-
ment of b, that activates some neuromuscular system in b which is di-
rectly or indirectly sensitive to s. And an overt response of an animal b to
an external stimulus s acting upon b is a change in the ongoing overt
behavior of b brought about or influenced by the neural system in b
activated by s. Finally, biopsychology modifies the S-R hypothesis by
interpolating the nervous system as the supreme intervening "variable."
The new version reads: Given a stimulus s acting on an animal b, b
contains a neural system n controlling the performance of an overt re-
sponse, in such a way that the latter is determined jointly by s and by the
state of n at the time the stimulus (or its transduction) reaches n.
Much of behavior is automatic and stereotyped, and proceeds to com-
pletion even when it is not biologically rewarding. For example, a wood-
pecker may peck repeatedly at a metal plate that it has mistaken for the
hideout of a worm; a goose may continue to perform the movement of
rolling an egg with its beak even after the egg has been removed; and a
bureaucrat may keep going through a totally useless routine. Such auto-
matic behavior is explained in terms of prewired units called motor pro-
grams (or tapes), the nature of which is far from having been disclosed.
Being prewired, every single motor program is inborn. However, expe-
rience may elicit novel combinations of such units: It may put several
motor programs together in the proper order; and lack of experience may
lead to the animal assembling the units in an inappropriate way (Gould &
8.2. Affect 169

Marler, 1984). Thus learning to swim, fly, or walk, as well as learning to


hunt, avoid predators, recognize kin, or sing, may consist in combining
innate elements in adaptive ways.
In addition to innate reflexes and their combinations there are, of
course, learned automatisms. It has been conjectured that the automatiza-
tion of a motor activity in the higher vertebrate consists in (is identical
with) the formation of a reflex-like control loop in the motor cortex
(Evarts, 1973). The higher vertebrate nervous system is so plastic that it
can learn to modify some of the inborn reflexes. In fact, animals of many
species can learn to regulate their own visceral rythms (e.g., their heart-
beats). The neural plasticity of the higher vertebrate is so pronounced,
that it has been postulated, "When a reflex competes with a learned
response, the learned response will take precedence over the reflex"
(Engel & Joseph, 1981). For example, sphincter contraction is a learned
response that in adult humans and many pets takes precedence over
natural incontinence of stool.
The behavior of any animal with a nervous system can only be under-
stood by investigating its neuromuscular system. For example, the study
of voluntary movement involves not only that of the muscles that execute
it, but also the study of the centers in the frontal lobes that do the willing,
planning, and attention giving involved in "instructing" the muscles to
contract in an effective way. To separate muscular activity from its neural
sources is as artificial and superficial as to dissociate the study of radio
waves from that of broadcasting antennas.
A favorite ploy of dualistic philosophers of mind is to claim that, al-
though physiology may give a causal explanation of behavior, it does not
yield an understanding of the latter because it remains silent concerning
the subject's intention-which is presumably an immaterial entity and
thus beyond the reach of physiology. Shorter: The claim is that science
does not fully understand behavior, which involves final causes, because
it acknowledges only efficient causes. This argument is fallacious for two
reasons. First, much of human behavior is automatic or semiautomatic
(i.e., unintentional). Second, as intimated earlier, and as will be seen in
some detail in section 9.5, biopsychologists regard intentions as brain
processes and investigate them accordingly with the help of microelec-
trodes and other tools that would certainly not catch any immaterial entit-
ies. What is true is that, as long as psychologists confined their attention
to the description of behavior, they gave philosophers an excuse for deal-
ing with intention in a pre scientific manner.

8.2 Affect
The term affect is used to denote a large variety of kinds of experience:
drives such as hunger and sex, emotions such as pleasure and anxiety,
and feelings such as empathy and love, as well as moral feelings such as
170 8. Basic Functions

shame and compassion. The early behaviorists ignored affect, but the
neobehaviorists, particularly Hull and Tolman, realized that behavior
cannot be understood without a consideration of drives. At the time of
writing the study of affect is being neglected once more by psycholo-
gists-though not by physiologists-this time as a consequence of the
cognitivist upheaval. If brains are computers, and the latter do not experi-
ence drives, emotions, or feelings, there is no point in bothering over
affect-except perhaps as a cognitive disorder best left to clinical psy-
chologists.
The scientific study of affect is important for the following reasons.
First, affect is a source of behavior, often a more important one than
environmental stimulation. Second, all cognitive processes, such as lis-
tening and problem solving, are motivated, whence no study of cognition
can be complete unless it includes an account of affect. Third, a remark-
ably large portion of the higher vertebrate brain is "implicated" in affect.
Fourth, because of its survival value, affect is likely to be phylogeneti-
cally very ancient, hence widespread across species, so that its anatomi-
cal "seat" is likely to be very ancient.
The physiology of affect was inaugurated in 1927 by W. R. Hess, who
caused rage and attack in a cat by electrical stimulation of a site in its
hypothalamus. Ten years later J. W. Papez conjectured that all of the
affective functions are processes in the limbic system. (This compara-
tively large and rather complex subcortical system is composed of the
amygdala, the hippocampus, the fornix, the cingulate cortex, and the
septal area.) Since then we have learned that the hypothalamus too is part
of the affect system, particularly with regard to hunger, thirst, and sex.
And, because the limbic system is coupled to the rest of the brain, in
particular to the cortex, it should not come as a surprise if it were found
that electrical stimulation of the cortex causes (indirectly) emotional pro-
cesses. In fact, in 1963 Penfield and Perot found this to be the case. These
and many other experimental findings obtained by physiological psychol-
ogists and neuropsychologists have both suggested and confirmed the
hypothesis that affect is a specific function of the supersystem composed
of the limbic system and the hypothalamus.
In addition to electrical stimulation, chemical stimulation and surgical
lesion have been used to locate affect "centers." For example, the injec-
tion of tiny amounts of acetylcholine in the septal region or in a few other
subcortical organs has been found to cause multiple orgasms. (The electri-
cal stimulation ofthe same organs has the same effect.) And the implanta-
tion of suitable hormones in the medial preoptic region of a castrated cat
restores copulatory behavior. On the other hand, surgical damage of the
amygdala causes hypersexuality-whence it has been conjectured that
the amygdala is a sex inhibitor.
Sex has long been recognized as a primary drive, but it is only in recent
times that physiologists and psychologists have undertaken to study it
8.2. Affect 171

scientifically. In the lower animals sexual behavior is elicited by physical,


chemical, or social signals acting on the endocrine system, which in turn
triggers the nervous system. These mechanisms can be unveiled by tam-
pering with the animal's neuroendocrine system. Such experimental inter-
ference with the normal course of events can produce hypersexuality,
deviance, or impotence. An indirect manner of altering sexual behavior is
to alter the stress level of the subject; in fact, it is well-known that stress is
a sexual inhibitor. Another is to have other drives, such as hunger or fear,
compete with the sex drive.
Human sexuality is particularly complex, therefore interesting. In fact,
in addition to being a neuroendocrine affair it is socially conditioned and it
involves the active and more or less conscious search for partner and
opportunity. Because human sexual activities are complex affective-cog-
nitive-volitional-muscular processes, the mere observation of overt sex-
ual behavior sheds no light on it. An understanding of sex calls for a shift
of emphasis from the genito-pelvic tissues to the nervous system, in par-
ticular the limbic system (Davidson, 1980). See Figure 8.1.
Because the limbic system is connected to the rest of the brain, it must
be expected to take part in learning. In fact, it has been known for a long
time that the stronger the motivation, the faster the learning, and that the
more closely associated with emotional episodes, the more vivid the

Ol Ol
c c
ii "E
ou o
u
Q)
Q)
a: a:

FIGURE 8.1. Hypothesized connections among the components activated during


sexual intercourse in male mammals. C = cortex, HL = hypothalamic-limbic
supersystem, S = spinal cord, G = genital system. (a) Normal orgasmic experi-
ence. (b) Pleasure without ejaculation. (c) Ejaculation without pleasure (occurring
in subjects who have suffered section of spinal cord). G-S-HL arrow: genital
sensation. HL-S-G arrow: affect modulates sexual reflexes. HL-C arrow: altera-
tion of consciousness (loss of contact with environment and loss of control over
selO. C-HL arrow: cognitive input. Inspired by Davidson (1980).
172 8. Basic Functions

memory. (Casanova, the famous 18th century womanizer, tells us that his
father spanked him while pointing to a salamander in the hearth, so the
boy would never forget the alleged fact.) Just as affect drives or inhibits
learning, so knowledge may in turn modify some of our motivations.
Hence there is a bridge, not a chasm, between affect and cognition.
(Moreover, it could be that in mammals the bridge is the hippocampus.)
The mere existence of the affect-cognition connection refutes faculty
psychology and invalidates much of cognitivism.
Valuation may be regarded, at least among the higher vertebrates and
when spontaneous, as an affective process. Actually all organisms, even
bacteria, show preferences (e.g., for a neutral over an acidic medium, for
light over darkness, and so on). We may speculate that every biospecies,
or at least every animal species, is characterized, inter alia, by a definite
value system manifested as a hierarchy of behavior types. Thus, normally
an animal presented simultaneously with a noxious (or merely fearsome)
stimulus and food, will withdraw; withdrawal (or even fleeing) behavior
dominates feeding behavior. However, if the animal is given no choice, it
may eventually learn to reverse its priorities (i.e., it will eat, paying the
price of some discomfort or even pain). In short, value systems are plas-
tic: just as plastic as the nervous systems in which they are implanted.
The concept of a value system has remained rather vague in the anthro-
pological, psychological, and philosophical literature. It can be elucidated
in a clear and simple manner by assuming that what organisms value is
being in certain stales. Consequently they only value external items
(things or events) insofar as they are instrumental in attaining those inter-
nal states. Accordingly we define the value system "If K of an arbitrary
biospecies K as the totality SK of internal states in which members of K
can be, together with the relation ~ of preference that characterizes them.
es ~ t' is read: 'Organisms of the given kind prefer state s to state t'.) In
short, "IfK = (SK, ~). In addition to the value system shared by all the
members of K, we may have to consider one idiosyncratic value system
"Ifb for every member b of K. Presumably, whereas "If K is inborn, "Ifb is
learned.
The concept of reward, which is unintelligible in strictly behavioristic
terms (except with reference to the experimenter), can be understood in
terms of the concept of value system. Indeed, the comparative concept
"more rewarding than" can be defined as follows. Let x and y be external
stimuli acting on an organism of kind K in state s, such that, whereas x
causes the organism to jump from state s to state I, Y causes it to go from
state s to state u. (In short, (x,s) ~ I, and (y,s) ~ u.) Then stimulus x is
more rewarding than stimulus y to organisms of kind K when in state s
= df organisms of kind K, when in state s, prefer t to u. (In short, x 2: sY
= df I ~ u.)
A simple physiological explanation of the fact that some states are
preferred to others, and that behaviors of certain types dominate over
8.3. Sensation 173

behaviors of others, is this. State or behavior of type A dominates state or


behavior of type B if, and only if, the neural system that controls A
inhibits (blocks) the neural system that controls B. In turn the occurrence
of such inhibitions, when innate, can be explained in evolutionary terms.
On the other hand, the reversal of innate dominance relations, as well as
the emergence of new ones, is to be explained in terms of learning, which
in turn may be explained as a lasting alteration of certain neural connec-
tions. We shall defer the study of learning to section 9. t.

8.3 Sensation
All organisms have sensors (detectors) of various kinds: of heat, pressure,
gravity, acidity, and so forth. Whereas some sensors are subcellular,
others are single cells and still others-particularly in our own sensory
systems-are complex neuronal systems. Sensors enable animals to mon-
itor environmental changes, or some of their own internal processes, and
to behave accordingly. Usually behavior is adaptive; for example, bacte-
ria tend to move toward the regions of greater food concentration or lower
acidity. At other times behavior is maladaptive. For example, many in-
sects cannot help circling around lights until they burn, and some people
prefer to die fighting than live working. Such cases of maladaptive behav-
ior refute the contention that the sensors' 'have been designed in order to
ensure survival." The evolutionary explanation of the apparent perfec-
tion of some sensory systems, such as the vertebrate eye, is that those
organisms that failed to have adequate sensors were not prepared (pre-
adapted) for sudden changes and perished before being able to reproduce.
With animals capable of habituating, and particularly of learning, the
story is somewhat different; they could adapt quickly to some of the new
environmental circumstances. Their behavior was thus not only a result
of evolution but also a motor of it (Piaget, 1976).
It takes a nervous system, though not necessarily a very complex one,
to form sensory maps capable of guiding adaptive behavior. For example,
certain invertebrates can be trained to take the right turn in a T maze, by
electrifying the left arm of the T. This is a typical avoidance reaction, and
it can be explained as a case of habituation, hence of inhibition or block-
ing, that does not involve learning proper. (Recall section 7.2.) In any
event, an animal that has been trained to always take the right turn in a T
maze has formed a sensory map of this part of its environment-though
not necessarily a mental model of it.
Sensory maps are of two basic kinds: those representing events in the
external world, and those representing happenings in other parts of the
body. The visual and auditory maps are of the first kind, whereas the
proprioceptive maps are of the second. See Figure 8.2. Note the key
words event and map. What the sensors detect is neither properties nor
174 8. Basic Functions

External
events

Internal
events

FIGURE 8.2. Neuronal system (a) records some external events, whereas system
(b)monitors some internal events. The lines joining the two represent axons or
nerve fibers.

states of things, not even things. They detect only changes in the state of
external or internal things: i.e., they detect events, such as the bouncing
of photons or the impinging of sound waves. Moreover the sensors map
events as further events, namely events of the neural kind. And the term
'map' is apposite, for the correspondence between the represented and
the representing events is precisely a map or function. We may call an
'atlas' the collection of all the sensory maps of an animal. (See Bunge,
1980, chap. 4, sect. 4 for details.)
The construction of somatotopic maps is a typical task of biopsychol-
ogy. We are all familiar with the "homunculus" spread over the human
cortex. There are similar maps for other animals. The technique consists
in stimulating the body surface and recording the resulting electrical activ-
ity in individual cortical neurons by means of microelectrodes. A recent
arrival is the discovery of the somatosensory map in a species of bat
(Calford, Graydon, Huerta, Kaas, & Pettigrew, 1985). This is of particular
interest because it exhibits traces of the evolutionary process leading
from walking to flying mammals. The differences found in bats relative to
other mammals reflect the lifestyle of the former and have resulted from a
major reorganization of certain nerve fibers that conduct signals to the
cortex.
Body maps or schemas are partly learned and they can be altered
experimentally. For example, the somatotopic map of a monkey changes
rather quickly following the amputation of a finger (Merzenich, Nelson,
Stryker, Cynader, Schoppman, & Zook, 1984). It is even possible to
cause intermodal compensations. (See Burnstine et aI., 1984.) For exam-
ple, rats reared in the dark develop a more complex auditory cortex than
light-reared rats; blinding reduces the weight of the visual cortex but
increases that of the somesthetic cortex, and it also increases the soma
size and activity of the neurons in the motor cortex. And blind human
subjects trained to read Braille process tactual signals at a much faster
rate than do normal subjects. All this confirms the remarkable plasticity of
the mammalian nervous system, and the corresponding modifiability of
the sensory and behavioral abilities.
8.3. Sensation 175

"
I I

distance

(a) (b)

FIGURE 8.3. Lateral inhibition. (a) A single prick on the palm of a human hand:
The excitation fades quickly with distance, and even turns into inhibition (numb-
ness ring). (b) Two pricks felt as one: The two curves add up to a single excita-
tion-inhibition curve (dotted line).

Discrimination is a more complex process than is detection. For exam-


ple, it takes human subjects between 2 and 5 ms to distinguish successive
tones; below this threshold one hears a single tone. Another example: If
we prick the palm of a human hand at two different points less than 1 cm
apart, the subject feels a single prick. The traditional explanation was that
the pressure sensors are not densely distributed. The correct explanation
is that this is a case of lateral inhibition, a typical property of the nervous
system (von Bekesy, 1967). See Figure 8.3. The neural mechanism of
lateral inhibition is this. If an array of neurons is stimulated, the central
ones inhibit those around them, so that sensation is confined. See Figure
8.4. Lateral inhibition is likely to be involved in higher processes too
(e.g., in mental concentration).
Scientific psychology started with the study of sensation. Its first
achievement was Fechner's famous law relating the felt intensity IjJ of a
physical stimulus to the strength S of the latter: IjJ = k log (S/So), where k

RESPONSE

11%STIMULI

FIGURE 8.4. Neuronal mechanism of lateral inhibition. The lower central neuron
inhibits its lateral neighbors, as a consequence of which the excitation propagates
along the central line only.
176 8. Basic Functions

is a constant and So the value of the sensory threshold. One century later
S. S. Stevens claimed that the correct formula is 1jJ = k (S-So)P, where pis
a parameter characteristic of the sensory modality. It is now believed that
the two formulas correspond to different tasks, and that both overlook the
important effect of context. (Human and animal subjects respond differ-
ently to one and the same stimulus object when placed in different con-
texts: McKenna, 1985; Zoeke & Sarris, 1983.) This has suggested to some
that what is actually being measured is not raw sensation but perception.
(See however, Laming, 1985.) In any event, psychophysics, which was
thought to have been exhausted, is now making a vigorous comeback, it
has been extended to animals, and it is questioning much of received
knowledge. This renaissance is partly due to having started work with
animals, which has forced investigators to invent objective substitutes for
introspective reports.
Color vision is still the most obscure of all sensory processes. Accord-
ing to the classical (or physical) hypothesis, what determines the color we
sense is the wavelength of the light that hits the retina. Thus, a thing will
be seen as red if it absorbs all the wavelengths except those comprised
within the red band, and it will appear multicolored if it reflects equally all
the wavelengths. Land's sensational experiments in 1959 refuted this hy-
pothesis. Take two black-and-white photographs of the same scene, one
through a red filter and the other through a green filter; then project these
two black-and-white pictures in superposition on a screen, interposing a
red filter in the path of the light from the projector. What you see is a
colored scene similar to a standard color photograph. This experiment
suggests that color is in the brain, not the retina. So much so, that split-
brain patients do not always see the same colors as normal people (Land,
1983). These findings are shifting the focus of research on color vision
from the retina to the visual area. (See e.g., Zeki, 1980.) Thus, psy-
chophysics is becoming rooted in neuroscience.
Philosophers have always been interested in color vision. There are
essentially three philosophical views on color and, in general, on qualia or
sensible (or phenomenal) qualities. They are naive realism, phenomenal-
ism, and scientific realism. Naive realists hold that blood is red, and they
take it for granted that this property can be explained in physical terms.
On the other hand, phenomenalists hold that blood only looks (appears to
be) red: that this property, like every other property of material objects, is
in the mind. Moreover they deny the independent existence of material
objects, which they take to be, in Mill's famous phrase, nothing but "per-
manent possibilities of sensation." Clearly, physics, chemistry, biology,
and other sciences reject subjectivism with regard to material things; they
are realists. (See Bunge, 1983b, 1985a.) But psychology confirms the
phenomenalist thesis that qualia have no independent existence.
Scientific realists, from Galileo on, have drawn a distinction between
primary and secondary qualities. A primary quality is a property that a
8.4. Attention 177

thing possesses whether or not it is being observed by a sentient being.


The number of components and the electric charge of a system are pri-
mary and intrinsic qualities of the system; its mass and temperature,
though relative or frame-dependent, are equally primary or subject-inde-
pendent qualities. A secondary quality is a joint property of a material
thing and a sentient being. Color, smell, taste, and smoothness are sec-
ondary; they appear differently to different observers or to the same
observer in different contexts.
Scientific realists solve the conflict between naive realists and phenom-
enalists by keeping the autonomous reality of material objects while ad-
mitting that some of the properties we assign them are joint or relational
properties of things and their observers in a given context. As well, scien-
tific realism indicates the way we are to understand the well-known claim
that each animal "constructs its own Umwelt (environment)" (von Uex-
kiill, 1921). Actually there is no such multiplicity of worlds: the octopus's,
the owl's, the human being's, and so on. There is a single world, com-
posed of things possessing primary properties, which is sensed and
mapped in many species-specific ways-in fact, as many as animal spe-
cies. The world exists by itself, whereas the maps of the world are pro-
cesses in brains. Whoever denies this realist thesis has no use for the
experimental checking of our conceptual models of things, and cannot
explain the history of science. Worse: He or she risks being referred to a
psychiatrist.

8.4 Attention
Looking and listening are more complex than are seeing and hearing
respectively. The former involve attention. In primates, gazing involves
not only the visual system but also the so-called eye fields, located in the
frontal lobes. In all the higher vertebrates attention may be accompanied
not only by motor processes but also by a number of higher functions,
such as memory, expectation, imagery, and volition. However, attention
can be drawn automatically by sudden environmental or bodily changes in
any animal endowed with a nervous system. Hence it is likely to be a very
basic and ancient faculty.
Two of the neural systems involved in attention in primates and other
higher vertebrates are the thalamus and the anterior part of the frontal
lobes. The thalamus contains novelty detectors, that is, neurons that
respond only to novel stimuli and habituate rapidly (i.e., cease to respond
on repetition of the same stimuli) (Jasper & Bertrand, 1966). Not surpris-
ingly, lesions in the thalamus impair an animal's alertness. When an ani-
mal is presented a cue that leads it to expect an event, particularly one
that will require it to perform a movement such as reaching for food, a
slow wave of activity appears in its anterior frontal lobe. This wave,
178 8. Basic Functions

technically known as CNV, has been called the 'expectancy wave' and it
is used as an objective indicator of a state of expectation. (See Evarts,
Shinoda, & Wise, 1984.) By the way, we have become accustomed to
thinking of attention as a state of readiness. Actually it is a process; it is
the ongoing activity of a group of neurons.
Attention may be general or specific; in the first case it is also called
'readiness,' in the second 'set.' General attention or readiness may be
explained as the simultaneous activity of all the attention systems. Spe-
cific attention or set results when all the attention systems except the one
sensitive to stimuli of a certain kind are inhibited. According to this,
"attending to stimuli of kind S" is identical to "inhibiting all the sensory
channels except that on which items of kind S impinge." In turn, whether
a receptor of a certain kind is ready to receive stimuli seems to depend
upon the state of a further neural system, a selector, located higher up in
the central nervous system-perhaps in the thalamus (Crick, 1984a;
Hebb, 1972; Milner, 1957). By the way, it has been known for quite some
time that, besides having a limited attention span, we cannot pay attention
to more than about half a dozen objects at anyone time (Mandler, 1984;
Miller, 1956).
One kind of "set" that has been explored from a biopsychological
viewpoint is the motor set, or preparation for motor behavior. Whereas
the dualist (e.g., Libet, 1985) regards this preparation as a state of the
immaterial mind, the biopsychologist takes it to be a brain process. More-
over, the latter is likely to conjecture that preparation for motor behavior
is a particular collection of processes occurring in a highly specialized cell
assembly. This conjecture leads the biopsychologist to search for such a
neural system, and this search may eventually lead to finding it.
As a matter offact at \east one such system has already been identified.
Indeed, Wise and Mauritz (1985) have localized 70 neurons in the premo-
tor cortex of the rhesus monkey, the activity of which seems to be identi-
cal with the preparation for movement. (See also Evarts et aI., 1984.) This
finding and others related to it show that preparation for movement (a)
though a mental process, is a brain activity, and (b) is anatomically and
physiologically different from actual motor behavior. These results are
damaging for behaviorism but confirm the psychoneural identity hy-
pothesis.
It is difficult to dissociate attention from curiosity, and the latter from
exploratory behavior-at least until we disclose their neural mechanisms.
We know very little about these matters. But we do seem to know this
much. First, attention is necessary but not sufficient for curiosity; an
animal may be paying attention to stimuli of a certain kind, and yet habitu-
ate quickly to them. (Even pleasurable stimuli may bore us after a while.)
Second, very old or severely sick animals lose curiosity; they prefer to
stay in familiar surroundings and avoid surprises. Thus one and the same
stimulus may elicit different behaviors, one in a healthy animal, a differ-
8.5. Memory 179

ent one in a sick organism; it elicits exploratory behavior in the former


while inhibiting it in the latter.
We should avoid confusing attention or alertness with consciousness.
An attentive or alert animal is aware of its surroundings and of itself, but it
may not be conscious of either for the simple reason that consciousness is
not the same as awareness (or the noticing of external or internal stimuli).
Consciousness is the monitoring of one's own perceptions and thoughts.
Many invertebrates are capable of attention (e.g., when stalking a prey or
searching for a mate), but they can hardly be attributed consciousness.
Yet some psychologists conflate consciousness with attention; that is,
they write about experiments on "conscious" (meaning "alert") rats. We
shall return to this point in chapter 11.
Finally, an item bordering on the next section. Memory and learning
depend not only on external stimulation but on the animal's state of
readiness: "Signals leave traces [engrams] only when the animal pays
attention to them and uses them for the control of behavior" (Singer,
1982, p. 221). In other words, the environment can induce no lasting
changes in the nervous system of an animal-or at least it cannot do so
easily-unless the latter is attentive. Moral for the animal experimenter:
Make sure the stimulus catches your subject's attention. (See also Daw-
son & Furedy, 1976; Dickinson & Mackintosh, 1978.)

8.5 Memory
Few subjects in psychology have attracted so many workers as memory.
However, the results of so much effort have been unimpressive until
recently. One reason for this scarcity of results is that' 'If X is an interest-
ing or socially significant aspect of memory, then psychologists have
hardly ever studied X" (Neisser, 1982, p. 4). Bartlett (1932) was of course
the outstanding exception to the tradition of failing to ask the right ques-
tions and to advance interesting hypotheses about memory. Fortunately,
physiological psychologists, starting with Hebb (1949), and neuropsychol-
ogists, in particular Luria (1973), saved the day. The study of memory and
its pathologies has become an important and rapidly moving chapter of
biopsychology.
In reflecting upon memory we should start by recalling that the nervous
system is not the only one capable of keeping a record of events. Fold this
page and you will produce an engram of sorts in it. Rocks are testimonials
of geological processes, and DNA molecules are archives of biological
evolution. The immune system records some of the onslaughts the envi-
ronment has inflicted upon the organism, and even the muscular system is
a lifestyle indicator. However, all such records are coded; in order to
"read" them out we need theories as well as observations. On the other
hand, the nervous system files experiences in a direct manner; it needs
neither theories nor new observations to remember the face of a friend or
180 8. Basic Functions

to phone her on her birthday. Paradoxically enough, it is precisely this


immediacy that makes animal and human memory so intriguing and so
different from, say, the hysteresis or the fatigue of a chunk of steel.
A dualist would claim that there are two radically different kinds of
memory: somatic and mental, the first being coded in the body, the sec-
ond being kept in the mind. Biopsychologists have no use for this dichot-
omy: They will argue that all memory is a lasting change in some material
system-in the CNS in the case of mental memory. Of course they will
grant that the change is not the same in both cases; for example, magnetic
tapes do not store information the same way brains do. However, in both
cases the "storing" is a material process and, in the case of the nervous
system, memory is not a state but a process, one likely to fade out or even
to suffer qualitative changes (e.g., embellishments). But if the dualist
were to ask what exactly the "storing" mechanism in the CNS is, biopsy-
chologists would have to admit that they do not have a solid answer; that
all they have is a collection of data and hypotheses. Still, these are not
wild speculations but are based on our (very imperfect) knowledge of
interneuronal contacts, and they are being investigated experimentally on
various levels: those of molecules, synapses, dendrites, whole cells, cell
assemblies, whole organs, and even organ systems.
It is common to assume that memory is the disposition (probability)
that a cell assembly will be reactivated when the stimulus, which was
present when the item was learned in the first place, reappears. This
assumption presupposes that we can remember only that which we have
learned. But this tacit assumption does not fit in with what we know about
systems which, like rocks, genomes, and magnetic tapes, have memories
although they cannot learn. It is also incompatible with our knowledge of
inborn reflexes. Thus, the pupillary reflex system does not need to learn in
order to function properly. On the other hand, it would be impossible to
learn without memory of some kind. We must therefore put the cart
(learning) in the correct place, namely after the horse (memory), not
before it.
Most biopsychologists agree that memory is a systemic or cooperative
affair rather than a property of individual neurons: "Memory resides less
in neurons than in relationships between neurons. A given memory re-
sides in a contingent of neurons only as much and as firmly as they are
interconnected" (Fuster, 1984b, p. 285). We shall assume then that mem-
ory is a lasting (plastic) modification of the strength of the synaptic junc-
tions in a multineuronal system. (Shorter: Memory is a plastic change in
the connectivity of a neuronal system.) Recent experiments on cats sug-
gest that every memory is identical with a change in the connectivity of a
system containing between 5 million and 100 million neurons located in
different parts of the brain (John, Tang, Brill, Young & Ono, 1986). There
is also experimental evidence for the hypothesis that such lasting modifi-
cation occurs only if the afferents to the system cooperate (Goddard,
8.5. Memory 181

1980). This is a version of Hebb's principle that neurons that fire together
stay together.
The biopsychological view of memory is at variance with the storage
metaphor made popular by information-processing psychology. Accord-
ing to the latter, to memorize an item is to file it away, be it in a temporary
file (short-term memory) or in a permanent one (long-term memory). Ac-
cordingly, to recall the item is to retrieve it from its storage. This view is
inadequate for at least two reasons. First, this account is just a metaphor
that tells us nothing about the "filing" or the "retrieving" processes. The
second reason that this account is inadequate is that, as Bartlett (1932)
notes, human memory is constructive rather than passive. This is due to
the fact that engrams are not isolated from one another but interact with
one another. In the process some memories become impoverished while
others are enriched. For example, the mistakes we sometimes make in-
volving synonyms and homonyms may be explained as follows. The cor-
responding traces being similar, as they fade out they may become either
more similar (in which case we confuse them) or more different (in which
case we forget that they are synonymous).
No matter how specialized and spatially localized the neural systems
may be, there can be no specialized and localized organ of general mem-
ory, precisely because the other functions are localized. (The same holds
for learning.) In fact, if we experience and learn now this, now that, using
a different part of the brain every time, memory must be a property of all
the neural systems that do the experiencing or the learning, as well as of
the hard-wired or committed ones. In other words, we may be able to
localize memory for this or that, but not memory in general. Shorter:
There is no general memory system but there might be several specific
memory systems.
As a matter of fact, it has been discovered in recent years that the
primate brain contains neural systems specialized in engramming skills,
and others in episodes. Thus, whereas motor experiences are recorded in
the motor strip, the medial-temporal region, and the cerebellum, mental
experiences are recorded in the cortico-limbic system. Such retention
systems have been identified by various experimental means, from mild
electrical stimulation to the injection of drugs and surgical lesions. For
example, electrical stimulation of the left pulvinar (a component of the
thalamus) impairs short-term verbal memory; on the other hand, the cool-
ing (to about 20 e) of the frontal cortex causes a deficit in short-term
0

visual memory. Another method is postmortem neuroanatomical exami-


nation. Thus the role of the mammillary bodies in memory was discovered
by examining those organs in individuals who had suffered from the Kor-
sakoff syndrome, characterized by amnesia, disorientation, and fabula-
tion. The existence of two retention systems, one for skills and the other
for memories proper, explains why damage to the latter need not impair
the former. This is why some amnesics can make use of skills without
182 8. Basic Functions

remembering the circumstances of their learning; theirs are cases of


"source amnesia" or "retrieval without recollection" (Schacter, 1983).
Certain striking experiments on monkeys, confirmed by neurological
observations on humans, have shown that there are two distinct neural
systems for retention, one of which involves the limbic system and an-
other that does not (Bachevalier & Mishkin, 1984; Hirsh, 1974; Malamut,
Saunders, & Mishkin, 1984; Mishkin, Malamut, & Bachevalier, 1984;
Mishkin & Petri, 1984; Weiskrantz, 1982). The first, or cortico-limbic
system, "stores" memories proper. It is the system in charge of what has
been called 'episodic memory' (Tulving, 1983). Its destruction causes
amnesia; the subject cannot recognize items that had been presented to
him only a short while ago, but he can still discriminate and even learn
new discrimination tasks. In particular, as first illustrated by the famous
case of Henry M., bilateral removal of the hippocampus causes loss of
recent memory but not of habits (Scoville & Milner, 1957).
Removal of the amygdala has similar consequences for recognition
memory (Saunders, Murray, & Mishkin, 1984). Thus, a human subject
who has sustained bilateral ablation of the amygdala is able to distinguish
two formerly familiar faces but cannot tell their names. Further investiga-
tion has revealed that, as far as visual recognition is concerned, there are
at least two parallel and nearly equivalent retention systems: one involv-
ing the hippocampus, the other the amygdala (Bachevalier, Parkinson, &
Mishkin, 1985). In fact, destruction of one of them has nearly the same
effect as destruction of the other.
The second, or cortico-striatal system (or rather supersystem), is in
charge of habits and know-hows, such as walking and driving. Destruc-
tion of this system causes the subject to lose certain habits and know-
hows-whence the name 'habit system' proposed for this second reten-
tion system. It has been conjectured that this system is phylogenetically
the older of the two, and it has been found that it develops earlier than the
episodic memory system. This is why, "whereas infants can readily ac-
quire habits, they are seriously deficient in forming memories" (Mishkin
et aI., 1984, p. 74).
How do we forget? Freud held that all forgetting is an effect of repres-
sion: The super-ego would suppress disagreeable or shameful memories.
This flight of fancy is far of the mark, because (a) forgetting occurs in all
animals, not only in those few capable of feeling shame, and (b) shameful
episodes are among those we usually remember most vividly. The most
popular answer to our question is that forgetting is the natural decay of
memories. However, this is a definition, not an explanation. Nor do the
behaviorists explain anything when they note (as it happens, correctly)
that forgetting follows insufficient reinforcement. This is only a descrip-
tion of an association. On the other hand, the biopsychologist does have
an explanation proper, namely this. Forgetting is identical with the fading
out of engrams (or neural traces), which in turn can be caused by such
8.5. Memory 183

events as spontaneous firing of nearby neurons, lateral inhibition pro-


duced by newly-formed engrams, and interruption of certain neural path-
ways (i.e., disconnection).
To conclude a story that has hardly begun, using the classical methods
of psychology, students of memory have made some interesting discov-
eries: For example, that there is short-term memory and long-term mem-
ory; that the former is limited and the latter unreliable; that the more
memorable events are the more emotional ones; that items are best re-
called when understood, and that the learning of new items may interfere
with the recall of previous experiences. All this descriptive and molar
material calls for explanations.
The best explanation, albeit still a sketchy one, is that memories are
engrams or lasting changes in the activities of certain systems of neurons
held together by plastic synaptic junctions, as described in section 7.2.
Now, the molecular turnover is quite rapid; for example, the average
lifespan of brain proteins is about one hour. Hence the engrams cannot
consist in molecules that do not change place or function. Rather, they
must consist in constant patterns of activity. It is likely that the molecules
in the synaptic cleft "interact in such a way that they can be replaced with
new material, one at a time, without altering the overall state of the
structure" (Crick, 1984b).
Until recently the engram hypothesis did not sound convincing because
it had been presupposed that the engrams were located in the cortex, and
the microscope had failed to locate them there. In the meantime the
clinical and pathological study of amnesics was showing that the memory
pathologies are located in subcortical organs, particularly the mammillary
bodies, the hippocampus, and the amygdala. Intensive experimental work
on monkeys confirmed these findings and revealed a number of distinct

FIGURE 8.5. Mishkin's model of visual memory in monkey. Flow of visual "infor-
mation" from primary cortical area (OC) through secondary areas (OB, OA, and
TEO) to the highest order visual area (TE), and from there to the medially located
amygdaloid complex and the hippocampal formation. From Mishkin (1982).
184 8. Basic Functions

memory systems. It has also suggested at least one biopsychological


model of visual memory. See Figure 8.5. As well, it has suggested that
deep amnesia, characterized by total loss of episodic memory, might be a
disconnection syndrome brought about by the interruption of the path-
ways from the temporal cortex-where habits seem to dwell-to the
frontal lobes, which lodge some of the centers of conscious cognition
(Warrington & Weiskrantz, 1982).
The just-mentioned findings and hypotheses are at variance with the
dualistic view according to which memory is a faculty of the immaterial
mind. They also refute the information-processing view, that memorizing
is storing, and recalling is retrieving. On the other hand, they confirm the
psychoneural identity hypothesis, as well as localizationism.

8.6 Summing Up
Many of the processes studied by psychologists are automatic, either
from birth or once they have been learned. There is nothing conscious and
nothing purposeful about such processes as turning the head when hear-
ing a tone, withdrawing the hand from a source of intense heat, getting
angry, detecting a pebble under the foot, remembering how to swim,
recalling a painful experience, or even falling in love. All these are basic
functions of the nervous system, even though everyone of them can be
influenced by prior or concomitant cognitive processes. A basic function
is one carried out by a neuronal system that either is prewired or has
formed at some early stage in development and has kept its integrity
through repetition.
The very way such basic functions are objectified in the laboratory
presupposes as well as confirms the psychoneural identity hypothesis.
Thus one assumes and confirms that a monkey is seeing a given stimulus if
certain neurons in its visual cortex are firing vigorously. (Moreover, one
may assume that the sensed intensity of the stimulus is equal to the firing
frequency of the corresponding neural system.) And one assumes as well
as confirms that the animal is "set" to detect stimuli of a certain kind, or
to perform a motor act of some sort, if a certain potential is recorded.
The fruitfulness of the neurophysiological approach is perhaps most
apparent in the case of memory. It has done more than suggest engram-
ming mechanisms that explain some of the facts described by prebiologi-
cal psychology. It has also come up with some startling new findings, such
as the existence of several memory mechanisms, located in different
though large parts of the brain. In particular, it has discovered that there
is memory for habits and know-hows, and memory for episodes and
know-thats. Thus the biological approach to mind is not only yielding
explanations of well-known molar psychological phenomena, but is also
increasing our knowledge of the latter.
CHAPTER 9

Higher Functions

We count learning, perception, conception, cogmtIon, and intention


among the higher functions of the higher vertebrate brain. Those mental
functions are called 'higher' because they are the most complex of all, and
for being peculiar to the most advanced animals in the phylogenetic tree.
And they are called 'functions,' and more precisely 'specific functions,'
because they are processes in special subsystems of the brain, much as
breathing is the specific function of the lungs.
There is another sense of 'higher function' that will not be used here,
namely as denoting an ability allegedly peculiar to human beings. To be
sure, we can do lots of things that no other animals on our planet can do,
such as building houses and theories, designing machines and institutions,
composing poems and paintings, and so on. However, these activities are
rather young, perhaps not more than about 10,000 years old. And they are
likely to have been invented by brains that were not much different from
our own. We must then assume that all of the highest mental functions
were learned by combining and refining simpler abilities that we share
with primitive human beings. In other words, it seems that in the case of
humankind the pace of cultural evolution has been far swifter than that of
biological evolution over the past few millenia.
From an evolutionary viewpoint it is likely that primitive mentality was
a refinement of hominid mentality, which in turn was a refinement of the
mentality of the ancestors common to both hominids and the modern
apes. Recent improvements in mental testing have shown that mammals
and birds have amazing perceptual and problem-solving abilities, not to
speak of memory and locale maps. There is even evidence that some
mammals and birds can form concepts.
We assume that, unlike habituation, which occurs even in inverte-
brates, learning is a higher function of the nervous system. We also as-
sume that learning is the most basic of all the higher functions: that even
perception presupposes learning for, unlike sensation-which can occur
automatically-perception can be perfected by experience. And, in line
with Tanzi and Lugaro, as well as Hebb and Bindra, we assume that
186 9. Higher Functions

learning an item is identical with the formation of a specialized system of


neurons held together by excitatory plastic synaptic junctions. (Recall
section 7.2.)
We conjecture likewise that the remaining higher functions are activi-
ties in neural systems. To be sure, so far we have only rather hazy ideas
about how the brain performs these activities. But we do know that most
of them are fairly well localized; that they are learned; that they can be
drastically altered by various means (electrical stimulation, drugs, cool-
ing, etc.); that they interact with one another as well as with the lower
functions; and that they decay with sickness and age. All this and more
has been learned by combining ideas and techniques of classical psychol-
ogy with those of neuroscience. Thus neuroscientists and psychologists
are finally implementing the research project laid down by Maudsley in
The Physiology of Mind (1876).

9.1 Learning
According to behaviorism, learning is a lasting change in behavior brought
about by practice. We cannot use this definition because it is too restric-
tive in one sense and too broad in another, and because it ignores the
nervous system. The definition is too restrictive because it does not in-
clude conceptual learning, which is certainly not identical with behavior
change even though it may occasionally be manifested as such. For exam-
ple, my learning a learning theory mayor may not change my behavior in
some respects, but in any case what it does is to change the furniture of
my mind-that is, to enrich some of my brain processes.
The behaviorist definition of "learning" includes habituation (or adap-
tation), which should not count as a kind of learning. We do not learn
anything when getting used to sitting on a hard bench, or to ignoring street
noises, or to living with the threat of nuclear annihilation. These are cases
of habituation, the simplest of which can be explained as elastic (as op-
posed to plastic) neuronal changes caused by repetitive stimuli. (Recall
section 7.2.)
The definition we are criticizing applies even to certain physical, chemi-
cal, and biological processes that have nothing to do with learning proper.
For instance, with prolonged weathering "practice," a stone becomes
ever smoother or, on the contrary, it acquires an increasingly contorted
shape. Is this learning? As a product-inhibition chemical reaction pro-
ceeds, its velocity decreases until it stops. Is this learning? And as a
consequence of repeated exposure to germs of certain kinds, a hospital
nurse becomes immunized to them. Has the nurse learned to cope with
them?
We conceive of learning as a lasting modification in a neural system,
other than either habituation or memory, that enables its owner to have
experiences it could not have had before learning. More precisely, let S be
9.1. Learning 187

a kind of (external or internal) stimuli that an animal b can sense or detect,


and call E a kind of event or process in a neural system of the same animal
b. Then b has learned e in E in the presence of s in S during the time
interval [I" IzJ if, and only if,
(a) e did not occur in b in the presence of s before II; and
(b) after Iz, e occurs in b whenever b detects s (i.e., b has memorized s or
some change caused by or associated with s).
This definition embraces all kinds of learning: motor, affective, and
cognitive. It includes learning about oneself, not only about one's envi-
ronment. (On the other hand, it excludes societal learning; all learning is
individual, even though in the case of gregarious animals it occurs in
society.) Because the definition makes room for internal stimuli, it does
not demand that all learning be an effect of external stimuli. Hence, unlike
the behaviorist definition, it applies to taste-aversion learning, which is
widely recognized as being inconsistent with behaviorist learning theory.
As is well-known, Garcia and Koelling (1966) paired taste and sound
with nausea caused by radiation, and found that the animal "overlooked"
sound but associated taste with radiation-induced nausea. This associa-
tion occurred even though the nausea was experienced several hours after
the stimuli were applied. This showed that animals are not passive trans-
ducers of environmental stimuli, but can ignore some (actually most) of
them. And it suggested looking for the neural mechanism "embodying"
the association. We now know that the taste-sickness association occurs
because "gustatory and visceral afferents converge directly to the brain
stem, indicating an intimate relationship among tastes, ingestion, and
emesis" (Rusiniak, Palmerino, Rice, Forthman, & Garcia, 1982). The
exact mechanism remains yet to be unveiled, but at least we know where
it resides.
Our definition covers both single-trial and multiple-trial learning. Each
of these two modes of learning has been studied by a separate school.
Single-trial learning has been studied by rationalists and intuitionists (in
particular the Gestalt workers), whereas multiple-trial learning is the spe-
cialty of behaviorists. Psychologists have known for some time that hu-
man beings and other higher vertebrates can learn in both ways: suddenly
and gradually. Moreover, we have learned recently that each mode of
learning is the specific function of a distinct neural supersystem. The
cortico-striatal system, which specializes in habits and know-hows, does
the step-by-step learning, whereas the cortico-limbic system, which spe-
cializes in know-thats, is capable of learning on a single trial (the "aha
experience' ') and retains not only what it has learned but also a memory
that it has learned the item concerned. (It is likely that both systems are
active at the same time, though probably with different intensities.) This
is the neurophysiological explanation for the fact that both behaviorists
and cognitivists have been able to make contributions to learning theory:
188 9. Higher Functions

The former to the more primitive forms of learning, the latter to the most
advanced ones. (See Hirsh, 1974; Mishkin & Petri, 1984.)
Note that we have distinguished learning from memory, and recall that
we can acquire habits without recalling the circumstances of learning
them (i.e., having no episodic memory of such events). Memory is a
component of learning: If we have forgotten to do (or feel or perceive or
think of) X, we must try and relearn X. Memory is thus more basic than
learning. An organism without memory of any kind would be incapable of
learning anything. But many organisms endowed with memory are inca-
pable of learning anything; all they can do is habituate or adapt. (Recall
section 7.2.)
At the neural level the difference between learning and memory would
seem to be this. Whereas learning is the process of formation of a new
neural system (cell assembly), or the emergence of a new activity pattern
in an existing one, memory is the preservation of either for some time, or
the ability to reactivate the activity in question. See Fig. 9.1. At the
cellular level the change involved in either the emergence of a new cell
assembly, or the appearance of a new pattern of activity in it, is likely to
be identical to a variation in the number, size, or relative position of the
synapses and dendrites. (See Greenough, 1984; Chang & Greenough,
1984.)
The following experimental findings corroborate this hypothesis. First,
rats reared in complex environments, where they have learning opportu-
nities, have about 20% more synapses per neuron in certain brain re-
gions-those doing the learning. Second, rats trained to run mazes exhibit
dendritic changes in the visual cortex. Third, in his now classic paper "A
Brain for All Seasons," Nottebohm (1981) reports that male canaries
change their repertoires from one spring to the next. So far, the behav-
ioral description. Now comes his explanation based on a neuroanatomical
investigation: When the canaries change their repertoires, their neural
nuclei for song control nearly double in volume relative to the fall, when
they stop singing.

Rd
I
+I
I ..
R1 R1
(al (bl (el

FIGURE 9.1. Three types of emergence of new behavioral or mental activity. (a)
Habituation: One of two neural systems is inhibited, whence the corresponding
response (R 2) does not occur. (b) Combination: Two neural systems combine to
produce a resultant response R that neither of them could have produced by itself.
(c) Creation: A new neural system is formed, with a radically new activity pattern.
9.1. Learning 189

Fourth, Morris, Anderson, Lynch, & Baudry (1986) used aminophos-


phonovaleric acid (APS) to block the N-methyl-D-aspartate receptors in
the neuron membrane, which respond to the neurotransmitter glutamate.
The net effect on rats is that the animals became unable to learn certain
tasks that the controls could learn rather easily. The experiment involved
swimming rats placed in a large pool and trained to escape by swimming
up to a platform. After they had learned this task the position of the
platform was changed. The animals in the experimental group took much
longer to find escape routes than did those in the control group. The
learning was selective; the rats lost their spatial learning ability, although
they kept the ability to learn visual cues. The learning impairment was
similar to that caused by hippocampal lesions. (See O'Keefe & Nadel,
1978, for the hypothesis that the hippocampus is the seat of locale or
"navigational" maps.) All four findings confirm the hypothesis that learn-
ing is the same as the reorganization, through chemical processes and
anatomical changes, of plastic neuronal systems.
The empiricist account of learning is that it consists in the association of
sensory impressions, which were regarded as the knowledge atoms. Re-
flexology and behaviorism refined associationism, by stating that learning
consists in the pairing of stimuli to responses-physiological in the case
of reflexology, and behavioral in the case of behaviorism. But neither
classical associationism nor its modern versions account for the learning,
let alone the creation, of abstract ideas, such as the concept of psychology
and the proposition "Psychology is becoming scientific." Moreover, nei-
ther of them propose a learning mechanism.
Eventually a neural mechanism was proposed. It was assumed that the
regions of the cortex lying between the primary sensory areas do the
association-hence the name 'associative cortex.' However, things
turned out to be more complicated. For one thing, subcortical organs,
in particular the limbic system, were shown to play an important
role in memory and learning. For another, the associative cortex,
which had been assumed to be the most recent because of its psy-
chic function, turned out to be more ancient than other regions of the
cortex.
Still, there is no doubt that we do learn some items by association. In
particular, we learn to associate sensory impressions with one another as
well as with internal (in particular visceral) processes, and either with the
outcomes of behavioral responses. (Strictly speaking we do not learn to
associate stimuli with responses, but the effects of stimuli with those of
responses.) Therefore, it behooves the biopsychologist to literally flesh
out such associations by conceiving of them as associations between two
or more neural systems, or as the formation or activation of new neural
systems.
The very first step in uncovering the neural "correlates" of associa-
tions is to look for gross anatomical components. Thus we can explain
190 9. Higher Functions

that the sight of a beloved object arouses emotions because it is known


that the visual system projects on to the limbic one. And we can under-
stand that the sound of cannon makes us reflexively take cover because
there are nerves connecting the auditory system with the motor one. (In
both cases cognitive centers. are involved as well.) In general, we may
postulate that two kinds of physiological events, at least one of which is of
the mental type, are or become associated if, and only if, there is a neural
connection between their corresponding loci. It follows that, when such a
connection weakens or is damaged, the corresponding association weak-
ens or disappears altogether. But a detailed explanation of associative
learning calls for more than molar anatomy: It requires going down to the
cellular level and even below it.
A general and comparatively simple theory of associative learning is
this (Anderson et aI., 1977). Consider two neural systems, a and b, such
that each is composed of N neurons, and that every neuron in a is poten-
tially connected with every neuron in b. (Such symmetry is assumed only
for the sake of mathematical simplicity.) The strengths of these N x N
connections can be displayed by an N x N connectivity matrix with ele-
ments Cmm where Cmn is the strength or weight of the connection of
neuron m with neuron n. The Tanzi-Hebb hypothesis, that the strength of
such connections increases when the two neuron assemblies are simulta-
neously active, can be exactified as follows.
The strength Cmn(t) of the m-n connection at time t will be assumed to
equal that at time 0 plus a term proportional to the current activities of m
and n. Calling Am and Bn respectively such activities, we have then
Cmn (t) = Cmn (0) + c Am (I) Bn (t), [9.1]
where c is a positive constant. (Remember that we are dealing with learn-
ing, not memory, hence with strengthening, not fading.) Assuming that a
and b were initially disconnected (i.e., C mn (0) = 0 for all m and n), and
setting c = 1, the learning equation simplifies to
Cmn (t) = Am (t) Bn (I). [9.2]
Actually this is only an arbitrary member of a system of N2 simultaneous
equations, which can be written in matrix notation:
C= BA, [9.3]
where A and B are the state functions of the neural systems a and b
respectively, and A is the transpose of A. (A state function for a system is
the list of functions representing its relevant properties. It can be written
as a column matrix. The outer product B A occurring in [9.3] is a matrix,
unlike the inner product A B, which is a number.)
According to the preceding, learning is identical with the strengthening
of interneuronal connections in accordance with [9.3]. The induced activ-
ity ofthe second or output neural system b is B = C A = BAA. If, for the
9.1. Learning 191

sake of simplicity, we assume A to be normalized to unity (i.e., if we set


A A = 1), then the pattern in the output neural system reduces to
B = CA. [9.4]
In words: After the synapses have been modified in a lasting manner,
every time activity A occurs in system a, activity B = C A occurs in
system b. This is how Anderson's theory captures the Tanzi-Hebb hy-
pothesis. In other words, the latter has become embedded in a hy-
pothetico-deductive system. (For an alternative formalization using graph
theory, see Palm, 1981.)
However, one should try to construct more sophisticated mathematical
theories of associative learning, including the following features: (a) the
spontaneous activity of the target-which inclusion can be effected by
adding an A-independent term to the RHS of [9.1], and (b) the fact that
every neuron in a is actually connectible to only some neurons in b-but
then by about a thousand synaptic junctions. Moreover, a more refined
theory should be probabilistic rather than deterministic, because of the
strong random component of both the spontaneous and the induced neu-
ral activity. Finally, the model should explain some of the well-known
regularities found by psychologists; perhaps it should entail Thurstone's
logistic function. In this way the molar regularities of learning would be
shown to emerge from neurophysiological microprocesses.
Presumably, the Tanzi-Hebb learning mechanism operates in every
plastic region of the brain-by the definition of "plastic." In other words,
there is no single learning center, just as there is no single memory center.
Habituation, memory, and learning can occur anywhere in the entire
cortico-limbic and the cortico-striate systems. However, as we noted a
while ago, there are different types of learning (e.g., of habits and of
know-thats, of sensorimotor and of cognitive tasks, and so on). Given the
overwhelming evidence for localizationism, we may safely assume that
the various types of learning are so many specific functions of special
neuronal systems.
Because there are different kinds of learning in any given higher verte-
brate species, and a fortiori in the whole animal kingdom, there must be a
large family of specific learning laws, each describable by one specific
theory (model). This multiplicity of special learning laws is consistent
with the existence of a single basic universal (cross-specific) learning
mechanism. In fact, the Tanzi-Hebb hypothesis can be supplemented with
specific hypotheses concerning the anatomical peculiarities of the various
plastic neuronal systems, the kinds of stimuli (when existing), and even
the specificities of the neurotransmitter-receptor pairs.
(We find a parallel in modern continuum mechanics and materials sci-
ence, which contain a family of models sharing the same basic equations
of motion. Every model is singled out by certain "constitutive relations"
describing the specific properties of the type of material in question. For
192 9. Higher Functions

example, there is one such "relation" for water, another for crude oil,
and so on. The logical aspect of such multiplicity of specific theories or
models Mi bound to a common general theory G is this. Every model Mi,
for i = 1, 2, . . . ,n, equals the set of consequences of G conjoined with
Si, where Si is the set of specific hypotheses that individuate the referents
of Mi. For details see Bunge [1983b].)
This is nothing but a research program, but one that should satisfy the
behaviorist yearning for a universal (cross-specific) learning law, and at
the same time answer the objection of the ethologist that animals of differ-
ent species are bound to learn different "things" in different ways. In
fact, our proposal combines generality with specificity. In particular, it
makes room for the facts that learning depends on the internal state of the
animal as well as on the "significance" the stimulus has for it. We have
repeatedly referred to the former factor; let us now deal quickly with the
second.
A tacit assumption of classical (behavioristic) learning theory is that
any two stimuli can become associated: that the choice of conditioned
stimuli is arbitrary. Experiment has repeatedly refuted this hypothesis,
showing that the individuals of every species respond only to stimuli of
certain kinds, whence they can learn only tasks of certain kinds. For
example, a pigeon that learns readily to peck a key for food may be unable
to peck a key to turn off an electric shock or a burst of rock music.
Another tenet of classical learning theory is that reinforcement of the
response is both necessary and sufficient to learn to associate it with a
conditioned stimulus. The principle looks self-evident for being a reword-
ing of the alleged axiom that all animals are hedonistic, that is, pleasure-
seeking (or maximizers of utility). Observation has repeatedly refuted that
tenet: Reinforcement seems to be only sufficient. Thus, a pigeon will
continue to peck at a key even after the latter has ceased delivering food
(i.e., in the absence of reinforcement). Presumably it cannot help pecking
even if it gets nothing out of it (Williams & Williams, 1969). The pursuit of
happiness requires more plasticity than that of a pigeon's brain.
Single-trial learning, a common experience in many higher vertebrates,
is another counterexample to classical learning theory. Furthermore, it
poses the interesting methodological problem of identifying instinctive
behavior. The usual criteria for deciding that a type of behavior is instinc-
tive are as follows: (a) the item appears without previous training, in
particular without repeated trials followed by reward, and (b) the item
appears at a very young age, though not necessarily at birth. However, it
would seem that both criteria are satisfied by single-trial learning as well.
If so, they do not suffice to discriminate instinct from learning. And if so
we need an improved concept of instinct, perhaps one making room for a
learning component at least in the case of the higher vertebrates.
Conventional wisdom has it that if behavior pattern X is innate, it is also
fully unlearned and it will be exhibited regardless of circumstances. In
9. 1. Learning 193

view of a mountain of adverse empirical evidence we should say instead


that, if X is innate, it will occur provided the animal finds itself in the
environment to which all the members of its species are normally ex-
posed. Should this environment alter in some radical manner, the animal
is likely to exhibit a behavior pattern other than X. For example, a gosling
deprived from its mother will attach itself to some other animal (e.g., to
Konrad Lorenz). This suggests that what is instinctive is not attachment
to mother but attachment to some animal. The environment will specify
the object of attachment.
In other words, instinct and learning do not exclude one another. If an
animal is born with the ability to do X, it will actualize this potential
provided it is reared in a suitable environment-otherwise not. (For ex-
ample, dogs and monkeys reared in isolation may attempt to copulate but
do not succeed.) Abilities other than primary reflexes neither mature
automatically nor are learned regardless of circumstances.
For example, in the rat nest-building probably does not mature autonomously-
and it is not learned. It is not "nest-building" that is learned. Nest-building
develops in certain circumstances through a developmental process in which at
each stage there is an identifiable interaction between the environment and or-
ganic processes, and within the organism; this interaction is based on the preced-
ing stage of development and gives rise to the succeeding stage. (Lehrman, 1953,
p. 344)
Does learning affect evolution, and if so how? We know one thing for
certain on this matter: that learned behavior is not inheritable. It is not,
because what we learn does not get coded into our genes. And it does not
because learning is a biological change on a supramolecular level: that of
dendrites and synaptic boutons. Such change involves, among other
things, the synthesis, breakdown, and transport of proteins, but not any
genic mutations or recombinations. On the contrary, for the supply of
proteins to be adequate, the overwhelming majority of genes must remain
relatively unaltered. Moreover, most mutations are maladaptive or even
deleterious, and some of them cause serious learning disabilities.
Hence when primitive human beings first learned to make axes, or to
keep fire, or to speak, their genes did not alter as a consequence. There-
fore their offspring did not inherit anything beyond the ability to learn to
make axes, to keep fire, or to speak; the newcomers had to do all the
learning from scratch. Whence Chomsky's hypothesis, that human beings
are born with a knowledge of universal grammar, is at variance with both
genetics and evolutionary biology. What holds for language holds for
every other cognitive faculty or social ability as well. Every one of us
starts as a blank slate with regard to both cognition and sociality. What-
ever gets inscribed on the slate after birth is to be credited to ourselves
and our teachers, not to our genes.
However, learning does affect the gene pool, hence evolution, in an
indirect manner. Indeed, by enhancing or impairing the chances of repro-
194 9. Higher Functions

duction, learned behavior can affect the distribution of genes in a popula-


tion. For example, both extreme meekness and extreme aggressiveness
will in general lower the chances of reproduction, possibly to the point of
extinction. So will overspecialization and a hypercritical attitude. In gen-
eral, acquired traits, if markedly idiosyncratic and maladaptive, are likely
to distribute at random and thus be unlikely to offset the genetic balance
of the population. But when new acquired traits are adaptive, they tend to
spread until they are shared by a significant fraction of the population; a
definite trend emerges in the gene pool, and evolution can take a new
turn. In short, learning can affect the genes but only indirectly, through its
social impact.
Two final questions: What do we know and what can we know? Our
answer to the first question is this: We know whatever we have learned.
This includes some skills, such as walking and eating, that have instinc-
tive roots but must be practiced and controlled in order to be mastered.
On the other hand, knowledge does not include inborn reflexes, such as
the knee jerk, which we can at most learn to control. In short, the knowl-
edge of an animal is the sum total of what it has learned. And the knowl-
edge of an animal species is the collection of everything that all its mem-
bers have learned.
As for what we can get to know (i.e., what is possible for us to know),
there is no knowing. We only know that, whenever human beings have set
out to know something, they have succeeded up to a point; whereupon
their successors, if equally motivated, have succeeded in pushing the
frontiers of knowledge a bit further. There seem to be no biological limits
to learning, hence to knowledge, because neurons and neuronal systems,
far from being finite-state automata, can be in any of a nondenumerable
infinity of states. The limits on human knowability seem to be physical
(e.g., the finite speed oflight and the destruction of much of the cosmolog-
ical, geological, paleontological, and historical record). They are also
social: the scarcity of human and material research resources, often for
lack of interest in the support of research, sometimes even from interest
in its suppression. But, given a chance and the will, humankind seems to
be capable of finding out whatever there is to be found out, and even of
creating theories and stories without counterparts in the external world.
(For more on the limits on science see Bunge, 1978.)

9.2 Perception
Following an old tradition, we distinguish between sensation, or detec-
tion, and perception, or recognition ("interpretation") of the detected
stimulus. We assume that sensation can be automatic (i.e., preattentive
and unlearned), whereas perception involves attention and can be per-
fected with practice. Sensation and perception are different physiological
9.2. Perception 195

processes. At least in the higher vertebrate, sensation is the specific func-


tion or activity of a sensory system up to and including the primary
sensory region in the cortex. On the other hand, perception is "the activ-
ity of mediating processes directly excited but not fully controlled by
sensory input" and it involves further brain regions (Hebb, 1963).
True, the frog retina is reputed to contain "bug detectors," the new-
born chick is said to be frightened at the sight of a falcon, and we have
feature detectors that react to very particular stimuli. These cases would
seem to refute the hypothesis that perception involves some learning.
They do not, because the frog and the chicken are easily fooled by crude
decoys; it is likely that they would react in the same way to any
fast-moving objects. As for the feature detectors, they analyze the in-
coming stimulation, so they are incapable of perceiving any objects.
Perception requires synthesis. In particular, global percepts (e.g., the
perception of a face) and perceptual constancies (e.g., the recognition
of an object after rotation) seem to require the cooperative activity
of neurons, that is, the activity of neuron assemblies (Hebb, 1949;
Poppe I, 1977).
A number of experiments confirm the sensation-perception distinction
and help us understand their nature. Primates retain their visual sensa-
tions upon removal of their primary (striate) visual cortex, but they lose
temporarily their ability to perceive: They see familiar objects but cannot
recognize them. For example, a monkey after striate cortex removal can
see small specks of paper and food, but cannot tell which is which without
tasting them (i.e., without using cross-modal associations learned before
the operation). However, such loss of perceptual ability is not totally
irreversible. The animal can learn to discriminate between visual events
using such clues as spatial location and orientation. Human' 'blindsight,"
or vision even after destruction of the primary visual cortex, is similar.
Detection remains, though in a blurred way, but pattern recognition goes;

s, S2

FIGURE 9.2. Stimuli SI and S2 activate sensory systems SI and S2 respectively,


which in turn signal to subcortical systems LI and L 2. If Ll and L2 are contiguous
and are activated simultaneously a number of times, they end up becoming cou-
pled according to the Tanzi-Hebb hypothesis (i.e., they form a supersystem).
Consequently, the stimulation of either SI or S2 suffices to activate both LI and L 2;
cross-modal association occurs. If either Ll or L2 is destroyed, the association
ceases.
196 9. Higher Functions

still, patients can improve with training (Poppel, Held, & Frost, 1973;
Weiskrantz, 1980).
Cross-modal association was mentioned a moment ago. An example of
it is this: Once we have learned to recognize an object both visually and
tactually we can recognize it (i.e., perceive it correctly) either visually or
tactually. The same holds for other cross-modal associations. A hypothet-
ical explanation of this fact is that each sensory system projects to some
subcortical systems, which become linked upon repeated simultaneous
stimulation. The amygdala and the hippocampus are two such systems.
See Figure 9.2. This hypothesis explains also the fact that, if the amygdala
of a monkey is surgically destroyed, the animal loses the ability to recog-
nize by vision alone a familiar object examined by touch, or conversely
(Murray & Mishkin, 1985).
The role of attention in perception is being vigorously investigated on
both the molar and the neural levels, particularly in the case of vision.
(One important motivation and source of funds for this research is the
wish to come up with machines capable of pattern discrimination.) A
standard technique is to present a subject with an array of targets (e.g.,
letters) to be detected or discriminated, and measure the time it takes him
to perform either operation. It turns out that, in the case of detection, this
time is independent of the number of items, but increases roughly in a
linear fashion with that number in the case of discrimination (Sagi &
Julesz, 1986). A facile explanation of this fact is that, whereas detection
can be done in parallel (simultaneously), discrimination requires serial (or
one by one) search. A possible biological explanation of the serial-paral-
lel distinction will be mentioned in a while.
A run of simple yet remarkable psychophysical experiments on visual
scanning is the one performed by Treisman and her coworkers. In one
experiment the subject is shown an array of black X s and in the midst of it
a red T. This irregularity "pops out" at him: He detects it preattentively.
But if the subject is asked to spot the red T in the display, it takes him
much longer; moreover, the search time proves to be proportional to the
number of "distractors" (i.e., Xs). The two tasks are quite different; in
the first the subject has to detect a singularity in an otherwise regular
pattern, whereas in the other he has to examine all the items one by one
(Treisman, 1982).
These and related experiments have prompted the construction of the
so-called feature-integration theory of attention (or perhaps 'attention
theory of feature integration'). According to it,
Features are registered early, automatically, and in parallel across the visual field,
while objects are identified separately and only at a later stage, which requires
focused attention . . . . Thus focal attention provides the "glue" which inte-
grates the initially separable features into unitary objects. Once they have been
correctly registered, the compound objects continue to be perceived and stored as
such. However with memory decay or interference, the features may disintegrate
9.2. Perception 197

and "float freely" once more, or perhaps recombine to form "illusory conjunc-
tions." (Treisman & Gelade, 1980, p. 98)
Clearly, this theory, and the experimental evidence buttressing it, contra-
dict the Gestalt principle according to which all perception is a unitary or
global act. (Recall section 5.3.)
These findings cry for neurophysiological research because, after all,
sensation and perception are brain processes. One relevant finding is that
the excitability of parietal light-sensitive neurons in the monkey is greatly
increased when the animal fixes its gaze on a target (Mountcastle, Ander-
sen, & Motter, 1981). A possible explanation ofthis fact is that the frontal
lobes (and perhaps other regions as well) contain cell assemblies that
"prime" the light-sensitive neurons, so that they will record light signals
(Milner, 1957). A hypothetical neuronal mechanism capable of doing this
trick is shown in Figure 9.3.
As for the biological difference between serial and parallel visual scan-
ning, Evarts et al., (1984) have proposed the plausible neurophysiological
hypothesis shown in Figure 9.4. Although this conjecture concerns only
the gross anatomy of the two visual systems, it is a first step on the long
and winding road leading to an adequate understanding of vision.
Another step in the same direction is the recent finding that the brain
analyzes every visual object into two components: what (recognition) and
where (location), in addition to analyzing it into separate features such as
edges and colors. To put it metaphorically, when presented with a visual
stimulus we ask ourselves: What is it? and Where is it? (Presumably there

A
o

op op o
P

Preattentive Set Go Perception


phase

FIGURE 9.3. When an animal pays attention to stimuli of a certain kind, its atten-
tional center(s) (A) in the forebrain alert or prime the suitable detectors. The latter
fire when the stimulus is presented. For example, when asked to find a blue
triangle in a display to be presented during a very short time, the subject's feature
detectors for blue (B) and triangularity (n are jointly "set" or prepared by the
attentional center(s). The subject's perceiving the blue triangle is identical with
the joint firing of Band T, which activate the perceptual unit P. This activity is
recorded electrophysiologically with the help of electrodes inserted in P.
~

Primary Secondary Tertiary Primary Secondary Tertiary


visual visual r--- visual visual f---- visual t-- visual
cortex -- cortex cortex cortex cortex cortex

t 1 t ~
::t:
Lateral Lateral <§:
geniculate geniculate Thalamus
nucleus nucleus ~
~
t n

::s
en
Superior
Retina I-- colliculus
G L--.. _ _ _ _ _ ~

FIGURE 9.4. Possible neural mechanisms for the processing of visual "information": (a) serial and (b) parallel.
Adapted from Evarts, Shinoda, & Wise, (1984, p. 54).
9.2. Perception 199

is a what-when temporal analog.) Surprisingly, each ofthese questions is


answered by a distinct component of the visual system-yet another
victory for localizationism (section 7.5.) (Hence damage to one of the
systems need not impair the performance of the other.) In fact, whereas
visual identification is performed by an occipito-temporal system rooted
in the primary visual cortex, visual location is performed by an occipito-
parietal system originating in the same area (Ungerleider & Mishkin,
1982). The precise location and extension of the two systems were deter-
mined by means of the deoxyglucose technique, by which local glucose
consumption is recorded autoradiographically-a technique not found in
the toolbox of the prebiological psychologist.
So far we have been concerned with certain psychological phenomena
and their tentative neurophysiological explanation. We have emphasized
the difference between perception and sensation, but have abstained from
suggesting that this difference is due to the fact that, when intent on
recognizing external objects or discriminating among them, the animal
hypothesizes or computes anything-the way Helmholtz and Gregory had
assumed. On the other hand, the theories of perception popular among
computer fans hold that, in all biospecies, perception is a highly symbolic
process that operates with sophisticated concepts such as those of mathe-
matical function and differential operator. According to these theories the
mind literally computes visual images and much else. For example, the
detection ofthe change of reflectance at an edge would "use" the Laplac-
ian of a gaussian function; and the detection of motion would "use" the
time derivative of the convolution of that laplacian with an intensity func-
tion (Marr, 1982).
The computational view of perception is open to the following objec-
tions. In the first place it involves a confusion between the biological
process of perception and any of the possible theories about it. To be
sure, theorizing too is a biological process, but one of a higher order than
perception and one occurring in a different region of the brain. None of
the perceptual systems can accomplish any conceptual feats. Blurring the
distinction between perception and computation amounts to conflating
fact and theory. To state that an animal computes its visual or auditory
images is as absurd as claiming that the planets integrate the equations of
motion of mechanics as they move. It is one thing to perceive (or to move)
and another to model perception (or motion). Keeping this distinction is a
necessary condition for understanding that one and the same process can
be modeled in different ways.
A second objection to the computational view of perception is that to
subordinate perception to conception is to go against the grain of evolu-
tionary biology. In fact, most of the animals capable of perceiving are
incapable of conceiving hypotheses or of computing anything. Third,
computationalism does not even intimate how the nervous system per-
forms the alleged computations, particularly in the case of animals and
200 9. Higher Functions

people who are not known for their mathematical proficiency. Fourth,
people who have suffered severe damage to their cognitive neural sys-
tems, to the point that their speech makes no sense at all, can still per-
ceive correctly anything but words. In short, the computational view of
perception is wrongheaded. (For further objections see Fodor, 1983.) To
put it positively: We must keep the traditional distinction between per-
cepts and concepts, even while acknowledging that the latter may influ-
ence the former.
An adequate theory of perception should serve as a foil for a number of
special theories (i.e., models) of perception corresponding to the various
sensory modalities. (For the relation between a model and a general the-
ory underlying it, recall section 9.1.) Such a theory should also serve as a
basis for models of illusion, imagery, hallucination, and even dreaming. It
would be illusory to look for a unified or stretch theory explaining all the
processes in any such category. For example, whereas some illusions are
accounted for in terms of altered contexts, others are explained as gestalt
switches. (Actually these are only descriptions. The effect of context may
be explained as the effect of a disturbance caused on the central process
by the processes of perceiving the contextual items. The gestalt switch
may be explained as an effect of neuronal habituation or fatigue.) And the
fact that we sometimes see pattern (or "good form") even in the absence
of it may have to be attributed to the action of higher-order cognitive
processes.
Imagery may be described metaphorically as closed-circuit perception.
It is the specific activity of a perceptual system in the absence of external
stimulation. Thus, when having a visual image our visual cortex is active,
and when evoking a musical fragment Heschel's gyrus may become ac-
tive. In fact there is good evidence for the first conjunct and, more partic-
ularly, for the hypothesis that visual imagery is a function of the left visual
cortex (Farah, Gazzaniga, Holtzmann, & Kosslyn, 1985). However, be-
cause imagery is stimulus-independent, it should be ranked higher than
normal perception.
Dreaming is probably in the same category with imagery. It is well
known that we have a couple of objective dream indicators, one electro-
encephalographic, the other being rapid eye movement. On the other
hand, no objective indicator of dream content is available. However, in
principle it should be possible to design one based on the fine-structure
recording of the activity of the left visual cortex. As for the fact that we
are usually unable to remember our dreams, or even having dreamt, there
is no mystery about it and no need to invoke repression. We forget ordi-
nary dreams the same way we forget ordinary events during wakefulness,
namely as a result of the quick damping of most neuronal activity. In
other words, dream memory is in most cases short-term memory. Finally,
there is no good reason for attributing dreams a biological value (Hippo-
crates) or a psychological one (Freud). Suffice it to recall that the brain is
9.3. Conception 201

ever active, sometimes-as in the case of nightmares-far too active for


our well-being. The distorted picture of reality, as well as the terror, of a
nightmare can have no adaptive value. The idea that dreams serve a
useful function is just as far-fetched as the idea that sickness is good for
you. Both ideas are but remnants of teleology.
Finally, hallucination may be characterized as abnormal imagery, or as
morbid illusion. Hallucinations can be elicited experimentally by certain
drugs, by sensory deprivation, or by electrical stimulation of certain parts
of the brain. The effect of neuroleptic drugs is understandable for, by
changing the chemical composition of the intercellular fluid, the connec-
tivity of the various neuronal systems is altered, sometimes to the point
where they become temporarily disconnected from the memory systems.
Sensory deprivation may have the effect of disrupting the balance be-
tween the cortex and the rest of the brain, to the point where the former
becomes freed from the reality constraints set by sensory stimulation. As
for the hallucinations caused by electrical stimulation of the brain, they
can be understood as a disturbance of the normal pattern of activity of the
cortico-limbic system. Halgren (1982) found no correlation between the
location of the stimulating electrode in the limbic system and the category
of experience evoked. Consequently, he conjectures that the effect of
stimulation depends on the pattern of current activity in other areas of the
brain. Subjects with different propensities, expectations, and life histo-
ries, are bound to react differently to stimulations of homologous neu-
rons. One more nail in the coffin of S-R psychology.

9.3 Conception
How are concepts formed? Three main answers to this question have
been proposed. According to classical empiricism every concept is either
a sort of distillate from a collection of percepts, or the result of a combina-
tion of percepts or some oftheir components. On the other hand, rational-
ists hold that concepts are either inborn (nativism) or the products of free
creations of the human mind (constructivism). Finally, scientific realism
begins by distinguishing two kinds of concept: empirical, or having expe-
riential counterparts (e.g., "hot"), and transempirical, or lacking such
counterparts even though they may represent aspects of reality (e.g.,
"entropy"). And scientific realism goes on to hold that, whereas the
former are rooted to percepts, the latter are stimulus-free creations of the
brain. Finally, it admits that some concepts, whether empirical or tran-
sempirical, are capable of guiding or misguiding some perceptual or motor
processes. See Figure 9.5. (For a detailed epistemological and logical
discussion of concepts, see Bunge, 1983a.)
The simplest concepts are empirical categories, such as "tree" and
"tall." They are arrived at by overlooking individual differences among
particular percepts. Categorization of some sort or other occurs probably
202 9. Higher Functions

Transempirical
concepts

p~oc ------~ PD •• Empirical concepts

Percepts

(a) (b) (c)

FIGURE 9.5. (a) Empiricism: Every concept originates in percepts. (b) Rational-
ism: Concepts are self-generated and they make some percepts possible. (c) Sci-
entific realism: Whereas some concepts originate in percepts, others are created;
besides, some concepts generate others and still others guide perception. From
Bunge (l983a).

in all organisms, no matter how primitive. Thus a bacterium moving away


from noxious stimuli of various types lumps them all in the category "bad
for you." It is then capable of forming categories, though not as concepts
or mental images for the simple reason that it lacks a brain.
To categorize, which is to detect recurrences in the environment despite varia-
tions in local stimulus energies, must be so enormous an evolutionary advantage
that it may well be universal among living organisms. Seen in this light, categori-
zation is just object constancy, which is perhaps the fundamental constancy to-
ward which all other perceptual constancies converge. (Herrnstein, 1984)
In the higher animals, categorization, or object constancy, can reach
astounding heights. For example, homing pigeons can recognize objects
presented in unfamiliar orientations, scoring even better than college stu-
dents (Hollard & Delius, 1982). See Figure 9.6. This finding suggests
rather strongly that the pigeons have formed concepts of the stimulus
objects. Not surprisingly, similar experiments with monkeys have yielded
similar results.
It must be stressed that we can be sure that these transforms [in orientation, size,
etc.] had never been seen before by the animals, and therefore they necessarily
stimulated a fresh grouping of, for example, "orientation" neurons in the striate
cortex. And yet each virgin image was able to address the canonical or prototypi-
cal store with impressive efficiency. (Weiskrantz, 1985, p. 11)
How do we know that in the case of the higher vertebrate at least some
categorizations are mental processes? Two groups of data suggest it
rather strongly. One is our knowledge about the way schoolchildren learn
a number of abstract concepts, such as those of history and fairness,
which cannot be reached from percepts. Another is from the study of
neurological patients who have sustained damage in the neocortex of the
9.3. Conception 203

A C

•••
•••
00

,Br '.I. ·t. ·J •• e


~ 7]'" • • •

tI
II

1'1 ... a ••
..l ,.!t ".1 ""f • • G ,.0 0

FIGURE 9.6. Test of object constancy in pigeons. (A) Experimental apparatus. (B)
Visual forms used. (C) Examples of stimulus sets used for rotational invariance
test. Reproduced from Hollard & Delius (1982).

temporal lobe; they are severely impaired in identifying transforms or


"unusual views" of familiar objects (Warrington, 1982). That is, damage
to a higher neural center can cause the loss of object constancy and of the
corresponding concept.
A possible neural mechanism for the formation of empirical concepts is
shown in Figure 9.7, inspired by Rubel and Wiesel (1968). The simple
cells at the bottom of the pyramid would analyze the incoming stimula-
tion; the neuron assemblies in the middle level would put the results of
such analysis together; and the neuron assemblies at the top would form
the concepts. For example, a particular triangle would be decomposed by
the simple cells and reassembled by the medium-level systems, while the
joint activity of a number of such systems, everyone of them engaged in
perceiving or imagining a triangle of a particular shape and size, would
stimulate a conceptual neuron assembly. See Figure 9.7.
Anderson and his coworkers (Anderson, Silverstein, Ritz, & Jones,
1977; Knapp & Anderson, 1974) have applied the principles of the mathe-
maticallearning theory reviewed in section 9.1 to account for categoriza-
tion, be it from percepts or from concepts. Recall that the learning system
under consideration is composed of two initially independent neural sys-
204 9. Higher Functions

Conceptual neural system


o Concept

~,p~
11\" ".'" ,Y,,,m'/I\
Percepts

Simple cells Analysis -/~


1 1 1 1 1 1
Sensory stimulation

FIGURE 9.7. Possible neural mechanism explaining the formation of empirical


concepts from percepts.

terns, a and b, the state functions of which are called A and B respec-
tively. In keeping with the Tanzi-Hebb hypothesis, learning is assumed to
consist in the formation of bridges between a and b. As a consequence, B,
which had initially been independent of A, ends up by becoming B = CA,
where C is the connectivity matrix: Recall Formula [9.4].
To account for concept formation start by assuming that similar stimuli
cause similar responses, so that the corresponding vectors A and B be-
come correlated. When activity A occurs in system a, B occurs in b. And
when a new pattern of activity A' occurs in a, the pattern in b is B' =
CA = B A A If A and A are similar (i.e., nearly parallel), their inner
I I. I

product AA' is large, whence B' == B. That is, b has a strong activity
elicited by that of a. If on the other hand A and A' are very dissimilar,
their inner product is small, i.e., the vectors A and A' are nearly orthogo-
nal, so that B' == O-that is, h is hardly activated.
Anderson's model is thus one of a categorizer or generalizer. More-
over, it accounts for the fact that the concept extracted from the original
percepts or concepts need not resemble any of them. For example, the
concept of a human being has no counterpart in the percepts of Eric or
Silvia. Indeed, consider a set of correlated input vectors AI. A 2 , • • • ,
An, with mean M. Each of them can be written as the sum of M and a noise
vector Di representing the deviation of Ai from the mean M, i.e., Ai = M +
D i • When these n patterns occur in system a, the total system composed
of a and h learns or consolidates the connectivity matrix
C =B A = ~iB Ai = B ~i eM + Di) = n B M + B ~iDi' [9.5]
If the Di cancel on the average, the connectivity matrix reduces to C = n
B M. In words: The system behaves as if it had been exposed to a single
pattern M that it has never perceived or conceived of before. It has
created an abstract concept. (Caution: This is not the only way concepts
are formed. The abstract concepts in mathematics and science have no
discernible roots in perception.)
9.3. Conception 205

Humans are not the only animals capable of forming universals, or


abstract concepts. Homing pigeons seem to be born with the abstract
concepts of identity and difference (Lombardi, Fachinelli, & Delius,
1984). And, because they exhibit a marked preference for asymmetric
over symmetric figures, they can also be said to master an abstract con-
cept of symmetry (Delius & Nowak, 1982). Besides, they master such
empirical concepts as those of water, tree, pigeon, fish, and people.
If pigeons, why not apes? The observations of Asano and coworkers
(1982), at the famous Primate Research Institute in Kyoto, Japan, have
confirmed this suspicion and refuted the credo that apes can name only
items that they have been trained to request, such as peanuts and balls. In
fact, those workers succeeded in teaching three chimpanzees to match
five color names and eight object names to the corresponding stimuli,
which were independent of the rewards. For example, the animals learned
to match a facsimile of a shoe to a lexigram consisting of a horizontal S
inscribed in a square, and a patch of red to a diamond crossed by a
horizontal line. To be sure this is "only" association, but one that in-
volves nonrepresentational signs (i.e., symbols). In another experiment,
the chimpanzee Ai was trained to count up to 6 objects of at least 125
different types, some of which she had never encountered before (Matsu-
zawa, 1985). The two experiments show that apes can learn to handle
symbols. Although human beings seem to be the only symbol-making
animal, they are certainly not the only symbol-using one.
Finally, how about creativity? According to some magical worldviews,
human beings are incapable of creating anything: All their novel concepts
and actions would be inspired in them by such supernatural agencies as
the muses, God or, more often than not, the devil. Empiricists laugh at
such stories but do not admit originality either, except of the combinato-
rial kind. They allow humans only the ability to associate percepts or else
the concepts derived from those percepts. Moreover, they hold that such
combinations may occur in great numbers but always in limited ways. On
the other hand, emergentism, be it idealist or materialist, admits creativ-
ity. And emergentist materialism explains creativity as the formation of
radically new plastic neural systems.
The emergence of new neural systems may be triggered by external
stimulation or it may be the outcome of the spontaneous self-assembly (or
self-organization) of neurons. In turn, the latter occurrence may be trig-
gered by chance by the spontaneous firing of one or more neurons. The
spontaneous activity of neurons was discovered seven decades ago (Gra-
ham-Brown, 1914). Regrettably it came on the eve of World War I and at a
time when most of the best research was being done by reftexologists and
behaviorists, who were dominated by the Aristotelian principle causa
cessante cessat effectus. A paper titled "The Intrinsic Factors in the Act
of Progression in the Mammal" could hardly attract the attention of peo-
ple who believed in the omnipotence of extrinsic factors. Fortunately that
206 9. Higher Functions

finding was eventually rescued by Hebb (1949), who made it one of the
pillars of his biological theory of mind.
Because empiricists do not believe in creativity, rationalists take it for
granted as a gift of the immaterial mind, and intuitionists do not believe
that it can be studied scientifically, the subject is still largely in limbo.
Because there are no adequate theories of creativity, no valid measures of
it are available. For example, the widely used Kirton Adaptation-Innova-
tion Inventory (KAI) (Kirton, 1976) measures efficiency (the ability to
work within existing settings), conformity, and the tendency to depart
from consensus (i.e., originality), but it opposes the latter to the tendency
to be methodical or disciplined. Hence, although the KAI may well be
measuring deviance, it cannot measure creativity, which requires a dose
of self-imposed discipline. No novel idea can be carried through without
some hard work. Recall the dictum attributed to Buffon: "Genius is 10%
inspiration and 90% perspiration."
In concluding, let us say something about the highest, most powerful,
and least well understood of all mental processes, namely thinking. A
thought, such as "Fruit is good," is a system of ideas (i.e., images or
concepts). (Conceivably, feelings too might participate in thinking.)
Whereas some thoughts are imageless or "abstract," (i.e., purely concep-
tual), others are strings of images, and still others are sequences of images
and concepts. A fortiori, so are arguments, in particular deductions.
Craik (1943) characterized thinking as the manipulation of mental or
internal representations of the world. This characterization has become
fashionable in cognitive psychology, but it cannot be adopted for the
following reasons. First, it fails for logic, mathematics, and theology,
none of which model the external world. It also fails for thoughts about
thoughts, such as "That thought is true." Second, the characterization is
phenomenological or molar: It gives no hint that thinking is a brain pro-
cess. Third, it is metaphorical: The brain (or the mind) cannot "manipu-
late" anything because it does not have hands.
A thought may be identified with either the sequential or the simulta-
neous activation ofthe neuron assemblies corresponding to its component
images or concepts. Note the following points about this physiological
characterization of thinking. First, it does not require that we have one
thought at a time. It is possible to think two thoughts at the same time,
particularly if one of them is imageless and the other in images. Second,
this account does not limit the number of biospecies capable of thinking.
It may well be that every animal capable of forming images or concepts is
also capable of stringing some of them together. On the other hand, it
seems that animals other than humans are not proficient at reasoning (i.e.,
constructing arguments). Third, when thinking aloud, speaking, or writ-
ing, the order of the component neural activities may change, and it may
change again when translating the same thoughts into another language.
Moreover, some words, such as "is" in "Fruit is good," have no counter-
9.4. Cognition 207

parts in images or concepts. In this case "is good" is a single concept.


(This becomes obvious when symbolizing the sentence in the predicate
calculus.) Therefore language is not a completely reliable speculum
mentis.
The physiological account of thinking is open to the following objec-
tion: Thought appears to be far too swift to be carried out by strings of
neuronal systems, linked as these are by comparatively slow chemical
processes. We counter this objection by noting that (a) actually the speed
of thought is rather slow-say, on the order of a few' 'frames" (images or
concepts) per second; (b) when neurons are "primed" (i.e., in a state of
preparedness or set), an elementary event, such as the release or the
capture of a calcium ion, can trigger a quick process. However, it must be
admitted that the physiology of thinking is still in diapers.

9.4 Cognition
Cognition embraces perception, imagination, language, and conception
(including thinking). Cognition is of course the subject matter of cognitive
psychology. This discipline, often advertised as the dernier cri de La
mode, is actually the oldest branch of psychology. Indeed all philoso-
phers, from Socrates to Kant, were more intrigued by cognition than by
any other mental ability.
Cognitive psychology must not be mistaken for cognitivism, the fash-
ionable school that identifies psychology with cognitive psychology and
either overlooks behavior, motivation, emotion, and volition, or attempts
to treat them as cognitive (and particularly computational) processes.
From a scientific viewpoint cognitivism is thus both narrow and imperial-
istic. From a philosophical viewpoint it is an instance of radical rational-
ism as well as of mentalism or even animism.
Viewed biologically, cognition is any specific function discharged by
certain plastic subsystems of the higher vertebrate brain. Different brain
subsystems specialize in different cognitive tasks. (This explains why
certain local damages cause acalculia but not agraphia, and so on.) How-
ever, the converse is false: Certain cognitive tasks, even if specialized,
may recruit a number of subsystems, perhaps as support systems. Indeed,
the monitoring of blood flow in the various regions of the cerebral cortex
shows that, when a subject engages in hard intellectual work, all of the
regions consume about the same amount of blood: All systems go.
The fact that certain cognitive tasks engage the entire cortex or nearly
so does not prove that complex mental activities are distributed rather
than localized. It only shows that the activity of the specific "centers"
radiates to others and enlists their support, as well as that of systems
other than the cortex. Among the latter the limbic system stands out. In
particular, the perceptual and conceptual cortex interacts with the thala-
mus and hypothalamus via the amygdala. See Figure 9.8. This anatomical
208 9. Higher Functions

Perceptual and conceptual areas

000

Hypothalamus and Medial and midline


basal forebrain thalamus

FIGURE 9.8. Interactions between cognition and emotion. Inspired by Aggleton


and Mishkin (1985).

connection between the organs of cognition and those of affect explains


why perceptions, memories, and expectations can arouse emotions, and
why the latter can elicit, distort, or even inhibit some cognitive processes.
In sum, though different and localized in different brain regions, cognition
and emotion interact. It is mistaken therefore to study them in separation
from one another. Consequently, it is wrong to set up cognitive science
centers outside of psychology departments, as is fashionable nowadays.
The nervous system does not condone such autonomy of cognition any-
more than it condones the autonomy of overt behavior.
A cognitive process mayor may not leave a lasting trace (i.e., it mayor
may not be learned). If it does leave an engram one says that the animal
has learned something, or that it has acquired a bit of knowledge. Other-
wise cognition does not amount to knowledge. Thus a fleeting cognitive
process such as perceiving a scene, imagining an event, or forming
an intention, without remembering anything after a while, are cognitive
processes and, moreover, they involve knowledge investments, but
they do not enrich the animal's knowledge. All knowledge is an out-
come of a cognitive process involving learning. (See Bunge, 1983a
chap. 2.)
Our definition disqualifies innate knowledge as being a figment of the
philosopher's imagination. Whatever we know we have learned one way
or another. The definition makes it also clear that, pace Plato and his
followers (e.g., Popper, 1972), there is no knowledge without knowing (or
rather learning) subjects, for knowledge is the end point of a brain pro-
cess. Knowledge is then as personal as feeling. But this does not entail
that all knowledge is subjective and private. Some of it is objective, that
is, valid irrespective of the particular individuals who acquire it. And
some knowledge can be made public in various ways-for example,
orally or in print. In particular, genuine scientific knowledge is both objec-
tive and public. But much knowledge is subjective and remains private.
9.4. Cognition 209

There are of course many kinds of knowledge and several ways of


grouping the family of knowledge types into large categories. One parti-
tion is into sensorimotor (e.g., knowing how to type), perceptual (e.g.,
knowing how to look through a microscope), conceptual (e.g., knowing
how to do sums), and linguistic (e.g., knowing how to greet in Japanese).
Another division is that into conscious and nonconscious knowledge,
made obvious by the discovery that certain deep amnesics can learn new
skills without knowing that they have acquired them (for having lost
short-term memory). A third division is into know-how and know-that, or
skill-based and data-based knowledge respectively. A fourth is the divi-
sion into self-knowledge and other-knowledge.
There is steadily mounting evidence for the localizationist hypothesis
that different kinds of knowledge are "stored" in (learned by) different
brain regions. In particular, know-hows and know-thats are learned and
remembered by different subsystems of the brain. In fact, skills are re-
tained by humans and monkeys with intact medial temporal brain subsys-
tems ("structures") even if they have sustained severe amygdala or hip-
pocampus lesions. On the other hand, if the medial temporal subsystems
are damaged, the subjects do not learn or remember having learned cogni-
tive tasks. Thus the know-how/know-that distinction, initially introduced
by philosophers, "is honored by the nervous system" (Zola-Morgan,
Squire, & Mishkin, 1982).
Every animal endowed with a partly plastic nervous system is capable
of learning something about itself and about its environment. We shall
deal with self-knowledge in chapter 11, devoted to awareness and con-
sciousness. As for knowledge of the external world, it can be character-
ized as follows. An animal b can be said to have acquired some knowledge
of (some of) its environment e if h possesses a plastic neural system n such
that some events in e are mapped into events in n. (The correspondence
between external and internal events is a map proper if neighboring points
in the event space of e are mapped into neighboring points in the event
space of n, that is, if spatially and temporally contiguous points in the
environment e are represented by spatially and temporally contiguous
points in the space of events occuring in the nervous system n.) There are
many such maps of the external world: sensorimotor, perceptual (in par-
ticular haptic, visual, and auditory), and conceptual. The collection of
all such maps may be called the animal's atlas of the world external
to it.
An important part of our atlas of the external world concerns our con-
specifics and, in particular, their mental processes. But how do we know
that other people too have minds? Might not this be sheer wild, albeit
generous, conjecture? (By the same token, how do we know that com-
puters and robots do not have minds?) This problem has worried un-
counted philosophers. It looks puzzling only from a dualistic viewpoint,
for if minds are immaterial then they cannot be observed from the outside.
210 9. Higher Functions

From a biopsychological point of view, on the other hand, the problem of


other minds isjust as simple as that of other metabolisms. Let us explain.
Any biologist knows that all living beings metabolize, and that the
moment most metabolic processes stop, the organism dies. So, if we
record certain vital signs of an organism, we can be sure that it is alive,
even without having measured its metabolism. Likewise, all neurobiolo-
gists know that all normally performing humans have brains. They can
gather both indirect and direct evidence for the hypothesis that a particu-
lar human being is minding. They can watch a subject's behavior and
make use of a battery of behavioral indicators (such as facial expressions
and linguistic utterances). Or they can monitor and even elicit, deviate, or
stop some of a subject's mental processes with the help of sophisticated
electrophysiological and neurochemical techniques.
What is more, neurobiologists are likely to venture the evolutionary
hypothesis that all normal humans are capable of minding, as they have
common ancestors, and hence they possess very similar brains. For the
same reason neurobiologists will not hesitate to attribute minds to other
primates, as well as to other higher vertebrates. But, of course, they will
become more and more reticent as distance increases in the descent tree.
On the other hand, the neurobiologist will abstain from attributing a mind
to a computer or to a robot. Indeed, the most cursory inspection of the
composition and organization of any machine, however sophisticated,
would exhibit such tremendous dissimilarities with human neuroanatomy
and neurophysiology, that the scientist would have no more reason to
attribute it a mind than for assuming that brains are composed of silicon
chips and run on electricity.
The remainder of this section will be devoted to a jumble of problems
about cognition: pattern seeking, preconception, model making, problem
solving, and intelligence. The first problem: The human being seems to
be, both perceptually and conceptually, a pattern seeker and maker. We
are forever seeking regularities or constancies, whether associations or
causal connections, be they natural (laws) or artificial (rules). For exam-
ple, young children, innocent of standard grammars, partly make up their
own grammatical rules as they acquire a language; in particular, they
regularize all irregular verbs. (Language acquisition is thus a combination
oflearning and invention.) In other words, we tend to overlook irregulari-
ties, imperfections, and even coincidences and exceptions. This propen-
sity is so strong that most people find it hard to believe that anything
accidental might happen. In particular, the psychoanalyst Jung made
much of "synchronicity" or coincidence, and there might not be magic or
parapsychology without such resistance to belief in coincidence or acci-
dent.
The Gestalt school made pattern seeking into a law in the case of
perception. As for conception, the neurologist Luria (1975, p. 339) writes
about the "law of disregard of negative information": "Facts that fit a
9.4. Cognition 211

preconceived hypothesis attract attention, are singled out, and are re-
membered; facts that are contrary to it are disregarded, treated as 'excep-
tional,' and forgotten. " Even scientists are likely to behave in this man-
ner-for better or worse.
This tendency to seek patterns everywhere is double-edged. On the one
hand it favors our finding genuine regularities, but on the other it prevents
us from noticing departures from regularities-as any proofreader knows.
It thus leads us now to truth, now to error. Therefore Popper's prescrip-
tion (Popper, 1959)-to always attempt to refute hypothesized regulari-
ties-is psychologically unrealistic and methodologically far too con-
straining. The normal scientific attitude is to start by looking for
confirming cases, to try and accomodate the early exceptions by ad hoc
hypotheses, and to give up the central hypothesis only in extremis. Only
the stubborn disregard for repeated exceptions is foolish or worse.
A preconception is a more or less tacit idea that had been learned
earlier and that may be kept even after it has been shown to be wrong. For
example, most children continue to believe that "multiplication makes
bigger" even after having learned to multiply by numbers smaller than 1.
And most adults continue to hold the ancient impetus hypothesis, accord-
ing to which a body slows down, and eventually stops moving altogether,
as its momentum or fuel becomes exhausted. The coexistence of incorrect
naive ideas with correct formal ones has been investigated experimentally
in the cases of arithmetical operations (Fishbein, Deri, Sainati Nelo, &
Sciolis Marino, 1985) and of the laws of motion (McCloskey, Caramazza,
& Green, 1980). One may speculate that such coexistence is nothing but
the persistence of the early intuitive cell assemblies alongside the later
ones. Unlearning (i.e., dismantling cell assemblies), is harder than form-
ing new ones. We pay for memory with preconception.
Cognitivists have made much of mental models, holding in particular
that to understand an item is to model it (Johnson-Laird, 1983). This view
has a grain of truth. Indeed, modeling X helps understand X. However,
the view is still very coarse. For one thing, the notion of a mental model is
vague-or, better, the expression designates a number of concepts,
among them those of mental picture, analog, and theory (Bunge, 1973c).
Because the processes of mental picture formation, analogizing, and theo-
rizing are so very different from one another, it is unlikely that a single
theory will fit them all. Second, if the model is either an analog or a black
box, it won't explain anything. Only a conceptual model (or theory) de-
scribing some mechanism can explain and, therefore, produce some un-
derstanding. Third, the very concept of understanding is complex, if only
because there are degrees or levels of understanding. Fourth, in addition
to a more precise description of the processes of understanding and mod-
eling, we need to build and test neural models of them.
Another specialty of cognitivists is problem solving, which behaviorists
had neglected, and the Gestalt workers had accounted for in terms of
212 9. Higher Functions

sudden insight or global intuition. Insight has fallen out of fashion since
the cognitivist upheaval, probably because computers cannot possibly be
attributed any. According to cognitivism the problem solver does nothing
but "process information"-and so, presumably, do the problem poser,
the conjecturer, the critic, the evaluator, and everyone else as well. In
this approach knowledge is a long-term store, and problem solving con-
sists in searching such store for relevant items. Moreover, the process is
regarded as a computation according to definite (though of course mostly
unknown) algorithms (i.e., explicit rules for symbol manipulation). No
room is made for originality other than novelty of the combinatorial kind.
Computationalism is supposed to hold for computers as well as hu-
mans. Thus Simon (1979) claims that BACON is both a program for
building theories out of data, and a model of theory formation in science.
The catch is that someone has to tell the machine which variables to
examine. Thus, if instructed to find the law linking the current to the
voltage in a DC metallic circuit, the computer is likely to come up with the
correct linear function, namely Ohm's law-provided the programmer
has fed it a set of correct current-voltage pairs. Therefore the credit goes
to the programmer and to the experimental physicist who supplied the
data; the computer has not discovered Ohm's law-or any other.
The computationalist view of problem solving surely holds for com-
puters working on well-defined problems for which well-defined methods
are known. However, someone has got to invent such rules to begin with.
And rule invention is not a rule-directed process-or, at least, nobody has
produced any rules for inventing rules. We handle original problems in
original ways, mixing reason with intuition. (See e.g., Bunge, 1962.) Only
some optimistic seventeenth-century philosophers, as well as some com-
puter fans, have been naive enough to believe that it must be possible to
invent an ars inueniendi.
Experimental research on problem solving has failed to confirm the
computational view of the human intellect. In particular, it has found that
real people do not behave quite rationally in handling the hypotheses they
form in the course of their problem-solving attempts. For example, in
principle the following four strategies are possible with regard to any
given hypothesis h relative to some set of data taken to be true:

h confirmed =? keep h h confirmed =? change h


(win-stay) (win-shift)
h refuted =? keep h h refuted =? change h
(lose-stay) (lose-shift)

Obviously, a rational subject will stick to the diagonal. As a matter of fact


a large percentage of people very often adopt the off-diagonal strategy
(Matthews & Patton, 1975). Likewise, most people do not obey the mo-
dus ponens (p, p =? q :. q), which is the basic rule of deductive inference
9.4. Cognition 213

(Wason & Johnson-Laird, 1972). And the great mqjority of preliterate


people do not dare derive any conclusions from a set of sufficient prem-
ises. They stick to the empiricist case-by-case strategy and shun abstrac-
tion altogether. For example, Luria (1976) in Central Asia, and Scribner
and Cole (1981) in Liberia, found that illiterate peasants presented with
the premises" All A are B, and" a is an A," will refuse to conclude that a
is a B unless they are acquainted with a. Computation indeed!
The just-quoted findings, together with a number of studies in develop-
mental psychology, have shown that reason is learned, not inborn-even
though it is undeniable that having the right genetic stuff is necessary for
learning to reason correctly. Now, reason, or conceptual intelligence, can
be of different kinds: the ease with which one grasps new material, makes
deductions, calculates, discovers problems, "sees" patterns, criticizes,
"jumps to conclusions" (i.e., forms hypotheses), and so forth. Most of us
are not good at all of these tasks, but specialize in some of them. Overall
conceptual intelligence is the prerogative of genius; intelligence of a single
kind is the mark of the idiot savant; and overall lack of intelligence is that
of the severely mentally retarded. Given the variety of kinds of concep-
tual intelligence, it is likely that there are several distinct intelligence
systems in the brain. This hypothesis explains the persistence of intelli-
gent behavior of some kinds, as well as the loss of others, when certain
parts of the brain are put out of action.
Intelligence testing is, of course, an attempt to measure intelligence of
the conceptual (rather than of the motor, perceptual, or social) kind. It
has become both an industry and a favorite target of methodological and
ideological criticism. Whereas some of these criticisms are valid, others
are not. Among the invalid criticisms of intelligence testing we single out
the following: (a) that intelligence, being an attribute of the immaterial
mind, is unmeasurable-wrong, for presupposing a myth; (b) that intelli-
gence, being multifactorial (i.e., being composed of a number of abilities),
cannot be measured by a single number-wrong, for, if intelligence can be
represented as a vector, then it has a definite magnitude as a whole; and
(c) that intelligence testing has been used to justify racial and class dis-
crimination-unfair because, even though the charge is true, the misuse
of an instrument does not prove it to be worthless or immoral.
The main valid criticisms of traditional or mainstream intelligence test-
ing are these. First, it does not rest on any accepted theory (or even a set
of exact definitions), whence it is a purely empirical operation. Because,
unlike height or skin color, intelligence is not a directly observable prop-
erty, it can only be measured with the help of indicator hypotheses having
both some theoretical and some empirical basis; it is in the same method-
ological boat with physical properties such as mass and charge. (Recall
section 4.3.) Second, intelligence testing is not reliable: We all know of
dull individuals who have been assigned a high IQ, and of very creative
ones who got a mediocre IQ. Third, it is not reasonable to try to evaluate
214 9. Higher Functions

intelligence on the strength of a few hours of questioning on problems that


the subject may not be interested in, rather than on the strength of one's
performance on a long-term project for which one feels motivated.
Fourth, most intelligence tests measure the ability to grasp and retain
information; they do not measure creativity. (For further criticisms see
Garcia, 1981; Hoffmann, 1962; Sternberg, 1985.)
The latter criticism points indirectly to a fifth weakness of traditional
intelligence testing, namely its lack of neurophysiological roots. Indeed,
patients who have sustained extensive excisions from the frontal lobes
can do well on traditional intelligence tests. (Remember, for example, the
astounding case of H.M.) On the other hand, the same patients do poorly
on tests of "divergent thinking," which it is hoped will measure (alas,
only empirically) creativity-qualities such as suggesting possible uncon-
ventional uses of a given thing, or posing problems of a certain kind given
a set of data, or cutting corners in a process of inference. As one might
expect, patients who have sustained excisions from the frontal lobes do
poorly on such tests because divergent thinking involves planning and
deciding, which are specific mental functions of those brain regions. This
is in fact an outcome of experiments done on both humans and monkeys
who, after frontal-lobe lesions, were instructed to perform certain tasks
involving the organization of their own behavior. (See e.g., Petrides &
Milner, 1982.)
We conclude this section by noting that the information-processing
variety of cognitive psychology (i.e., cognitivism), has made no remark-
able contributions to our understanding of cognition. Worse, by isolating
cognition from affect and motor control, it cannot even begin to explain
such basic phenomena as exploratory behavior and the coupling of per-
ception and movement. (Recall that the unmotivated animal does not
explore, and the immobile human eye does not see.)
Why then the phenomenal current popularity of cognitivism? One obvi-
ous reason is that it handles many genuine and important problems that
behaviorism had ignored. However, American psychologists often forget
that such problematics, though neglected in the United States between
the two world wars, had been handled by such distinguished European
cognitive psychologists as Piaget, Vygotsky, Rignano, Claparecte, and
Bartlett, not to mention the whole Gestalt school. The main reason for the
popularity of cognitivism seems to be a different one, namely that, like
Santa Claus, it has got something for everyone.
In fact, the proposed reduction of all cognitive processes to computa-
tion appeals to simplicity lovers as well as to those who wish to be spared
the study of neuroscience. It appeals to some materialists because it
draws no distinction between animal and machine; to some idealists be-
cause it detaches mind from matter; to some rationalists for stressing
computation; to some empiricists for its emphasis on data processing; to
some psychologists because it sounds very technical and novel; to some
9.5. Intention 215

Artificial Intelligence (AI) experts because it simplifies psychology for


them. Finally, it appeals to some linguists and philosophers for being
basically simple and nonexperimental, as well as contrary to both behav-
iorism and biopsychology. Obviously, only a simplistic view could possi-
bly command such a large measure of consensus across disciplines and
schools.

9.5 Intention
According to the idealist tradition, only humans (and perhaps also deities)
have intentions and the will to carry them out; animals are machine-like.
This view has persisted until recently. For example, both Vygotsky and
Lewin regarded voluntary activity as a product of the historico-cultural
evolution of behavior. And both Eccles (e.g., 1982) and Libet (e.g., 1985)
have used intention and voluntary movement as instances of the action of
the immaterial mind on the nervous system. This idealist view began to
evaporate, or at least to become obsolete, the moment the first results
from the neurophysiology of voluntary movement started to come in. (See
e.g., Evarts et ai., 1984; Goldberg, 1985.)
The first inkling that the will has a definite neural seat, mainly in the
frontal cortex, came when lobotomized patients were studied in the mid-
1930s. It was found that they had all but lost the capacity to make plans
and decisions-their willpower had been excised with a lancet. Later
studies, on monkeys and humans, showed the existence of a large number
of neurons that "are not activated by sensory stimuli, but discharge at
high rates when the animal projects his arm or manipUlates with his hand
within the immediate extrapersonal space to obtain an object he de-
sires" (Mountcastle, Lynch, Georgopoulos, Sakata, & Acuna, 1975,
p. 904). Moreover, long before (about 800 msec) the actual voluntary
movement starts, a special change in the potential-the readiness
potential-recorded from the scalp appears. Intention is thus a brain
process.
We may characterize a voluntary movement as a movement initiated
within a higher brain center. It involves set or preparation, and its reac-
tion time is longer than that of the corresponding reflex, if any. Inciden-
tally, voluntary movement involves reflex movement rather than being its
opposite. For, to put it metaphorically, the will keeps some reflexes under
control and organizes them. Thus running involves several automatic
stabilizing movements. And voluntary movements of other types are pre-
ceded by sensorimotor or conceptual learning, as is the case with the
precise movements of the craftsman. Here we may speak of "cognitive
steering" rather than of "downward causation" from immaterial mind to
body. This steering is an action of one part of the brain upon another.
Free will has been endlessly discussed in the philosophical and theolog-
216 9. Higher Functions

icalliterature. Until recently psychologists have either ignored or denied


it. For example, Mandler and Kessen (1974, p. 341) hold that "man is as
free as a falling leaf," although they share James' belief that the false
belief in free will may be useful, and they are interested in studying the
psychogenesis of that belief. On the other hand, Hebb (e.g., 1980), who
can hardly be regarded as a "softy," has admitted that free will is real and
moreover a biological phenomenon, hence one that can be studied scien-
tifically. But such study requires some prior conceptual clarification, a
task in which the philosopher may be useful.
To begin with we need a definition of the concept. We shall stipulate
that free will is volition with a free choice of goal, with or without fore-
sight of the possible outcome. In other words, we propose to call a behav-
ioral or mental process free if, far from being either independent of
antecedent conditions (i.e., indeterminate) or fully controlled by
sensory stimulation, it is innerly directed and, more particularly, it is
under the control of conceptual processes. If preferred, we stipulate
that an animal acts of its own free will if, and only if, (a) its act is volun-
tary (rather than either indeterminate or determined by external coer-
cion), and (b) it has free choice of its goal: That is, it is under no
programmed or external compulsion to attain the chosen goal (Bunge,
1980).
The philosophical literature is littered with confusions regarding free
will. Two of them are the alleged identities "determinism = predictabil-
ity" and "free will = indeterminacy." Actually the concept of determi-
nacy is an ontological category, whereas that of predictability is an episte-
mological one (see Bunge, 1979b). Hence, in principle we can have the
one without the other. For example, even though a process may be per-
fectly determinate (i.e., lawful and subject to antecedent conditions), we
may know it only imperfectly, and therefore may not be in a position to
predict it. Most physical processes are of this kind. Likewise, the concept
of free will is an ontological (and ethical) category, so predictability does
not count against it, and unpredictability cannot be taken as a test or
criterion offree will. If we know a person reasonably well, we may be able
to predict that, whenever he finds himself confronted with a problem of
kind A, he will choose of his own free will (i.e., irrespective of external
compulsions) to perform actions of type B. Free will may then be both
perfectly determinate and predictable.
The possibility offree will is a presupposition of planning, one of the so-
called executive functions of the brain. Other executive functions are the
regulation and checking of behavioral or mental processes. Cognitivists
hold that these functions are "encoded" in highly specialized "sche-
mata" or "routine programs," the "neural embodiments" of which they
regard as unimportant and therefore do not care to specify. (Recall sec-
tions 5.4 and 9.4.) They think of such programs as being activated or
9.5. Intention 217

Initial position Goal position

FIGURE 9.9. The Tower of London test. In this case a minimum offour moves is
required. See Shallice (1982).

"released" by external stimuli. (Initiative plays no role in this view.)


Moreover, cognitivists hold that the various "programs" run indepen-
dently from one another. This seems indeed to be the case; thus one can
plan the day's activity while preparing breakfast. But all this is mere
description in computerese. One should be more curious and inquire into
the neural mechanisms of the executive functions.
The most telling data about these functions are coming from neuropsy-
chology. Experimental work on chimpanzees, and clinical work on hu-
mans, suggest that the left frontal lobe is dominant (Milner, 1982). The
kind of test capable of revealing deficits in initiative and organization
caused by lesions in that region is not the one where the subject is asked
to reproduce or copy a task designed by the experimenter. Rather, the
subject is asked to organize and execute a sequence of responses in view
of some goal: He or she is challenged to show initiative and planning
activity. The Tower of London test has been employed to this end (Shal-
lice, 1982). Three beads (one red, one green and one blue) have to be
moved from a starting configuration to a target position in a minimum
number of moves. See Figure 9.9. Patients with unilateral lesions, of
various etiologies, were compared with normal subjects. Those suffering
from lesions in the frontal lobe showed a marked deficit on the task. If, in
addition, the subjects are asked to repeat continually 'ABCDEFG' to
themselves while doing the test, it is found that the performance is not
significantly impaired (Shallice, 1982). This result refutes Luria's well-
known hypothesis that speech, in particular inner speech, regulates the
performance of executive functions.
To conclude, volition, until recently banned from scientific psychology,
is back. Initiative, planning, and other executive functions are being stud-
ied experimentally in humans and animals. Moreover, they are being
deliberately tampered with by surgical and chemical means, in an attempt
to disclose their neural mechanisms. Even intention, formerly regarded as
nonmaterial, has been brought within the ken of experimental biopsychol-
ogy. Indeed, it can be measured as the firing frequency of certain neurons,
namely those that do the set or preparation for a task. Therefore it is no
longer possible to regard intention as the mark of immateriality.
218 9. Higher Functions

9.6 Summing Up
Mental processes, formerly the preserve of mentalism, are now being
studied biologically as well. Though young, the biology of mind has come
up with some results that would have gladdened the materialist philoso-
phers of ancient Greece as well as the medical schools of Hippocrates and
Galen. For one thing, it has confirmed the psychoneural identity hypothe-
sis. More specifically, it has corroborated the localizationist hypothesis,
that every mental process is the specific function of some subsystem of
the brain. This has made it possible to alter many mental processes-to
the point of starting or stopping them-by electrical, chemical, or surgical
means.
However, though impressive, the findings of the young biology of mind
are still rather few and spotty. The main result of these studies is perhaps
the growing realization that even apparently simple mental processes,
such as correctly identifying a visual stimulus (as different from locating
it), are so complicated that their description in traditional molar terms, or
in fashionable information-processing or computational ones, does them
no justice. (After all, there are nearly 20 distinct visual "areas" in the
primate neocortex, and everyone of them seems to analyze the visual
field in its own way-ways that are only now being slowly unveiled.) This
complexity guarantees that, as long as psychologists continue to be curi-
ous and to resist the temptations of simplicism, they will not run out of
research problems. And keeping curiosity alive requires keeping dogma-
in particular psychoneural dualism-at bay.
This concludes our examination of the young biology of behavior and
mind. The next part of the book will be concerned with the social matrix
of human behavior and mind, as well as with the applications of psychol-
ogy to the deliberate modification of behavior and mind.
v

The Social Aspect


CHAPTER 10

The Social Matrix of Behavior

Humans and their behavior exist in a social context. Without a nervous


system there is no human behavior, and there is none without society
either. It has been said that an isolated chimpanzee is not a real chimpan-
zee, and the same can be said of human beings.
Through social interaction human beings acquire the attitudes, values,
goals, and patterns of behavior that define them as members of the spe-
cies. They also acquire prejudices and maladaptive behaviors. Society
surrounds all humans from birth, imposing its norms and pressuring them
to act in particular ways. The enormous differences that can be observed
between one culture and another are proof of the pervasive influence of
society on behavior.
Because of the importance of the social matrix in explanations of hu-
man behavior, we hold that a critical analysis of the foundations of con-
temporary psychology requires an examination of the notions of culture,
social class, socialization, and other similar topics. This also seems the
appropriate place to consider the problem of classifying psychology as a
natural or a social science.
The nature of relations between the individual and society have often
been the object of speculation. "Crowd psychology" and "psychology of
the masses" come to mind as examples. Analogies have been made be-
tween organisms and animal or human societies. Questions have been
raised about the possible existence of a collective mind and about the
appropriate unit of analysis: whether it is the cell, the organ, the organ-
ism, the small group, the collectivity, or the culture as a whole. The
notion of a collective consciousness has intrigued many philosophers,
sociologists, and psychologists during the last century, although it is
surely as much of a pseudo-problem as has ever been found in the devel-
opment of science.
Problems such as the generality of psychological laws, the adequacy of
methods for studying social phenomena, the influence of politics and
ideology, the historical nature of human behavior, the differences be-
tween cultures, socialization, subjective responses to the human-made
222 10. The Social Matrix of Behavior

part of the environment, the appropriate units of analysis, language, the


influence of culture on psychological processes, and the existence or not
of universal conceptual categories, are all part of this domain of study.
It is surely !>~ssible to construct explanations of human behavior in-
cluding the contribution of genetic endowment, biochemical constitution,
behavior, culture and their interactions. The relative weight given to each
cOffiponent of the explanation ." ~ll be a function of the behavior explained
and the results of research.
The study of humans' social matrix has had a long history and has
produced many opposing theories. Most of these are quite soft and are
hypotheses that hardly deserve being called theories. Herodotus, Thuci-
dides, Aristotle, and other Greeks wrestled with these problems, as did
Ibn Khaldun and, more recently, Montesquieu, Herder, Vico, Le Bon,
and Wundt. Wundt attempted to find some order in the mass of data
gathered by anthropologists, historians, and linguists, and ended up em-
phasizing the ways in which thought is conditioned by language, myth,
and customs. Later on came studies of culture and personality that were
very much influenced by psychoanalysis, by cultural relativism, and by
national character, which do not rate very highly as scientific investiga-
tions. Linguistics became very influential, even when it seriously consid-
ered the Sapir-Whorf hypothesis that we see, hear, and experience the
world as we do because the language of our community furnishes us with
certain categories for interpreting the world. Consequently it was con-
cluded that observers would not have the same picture of the universe
unless their linguistic backgrounds were similar. Anthropologists like
Boas, Mead, Benedict, and Kluckhohn shed light on these mostly psycho-
logical problems.
Investigation in social psychology and particularly in transcultural psy-
chology has taken great strides. There is still a long way to go, but we are
beginning to understand the social matrix of human behavior. The jigsaw
puzzle of society, culture, politics, human relations, and the mythical
"collective consciousness" is beginning to make sense.
In the present volume we have emphasized physiological psychology,
and given the neurosciences the place they deserve in the search for an
adequate understanding of human behavior. Mind, like digestion, makes
no sense without physiology. Nevertheless, at the level of the behavior of
organisms psychology functions like the other behavioral sciences-eco-
nomics, anthropology, sociology, linguistics-all of which would not ex-
ist if there were no biology, but whose laws and hypotheses refer not to
physiological but to behavioral events.
Behavior is what organisms dO or say. This definition goes further than
Watson's classical definition involving muscular contractions and glandu-
lar secretions. Psychology is the scientific study of behavior and its rela-
tion to the environment, in organisms endowed with a nervous system
enabling them to learn. Psychology is not the only science that studies
10.1. Psychology: Natural Science or Social Science? 223

behavior, but, as Wundt emphasized more than a century ago, it is the


basic science of behavior.
A behavior may be described as a functional relation with the environ-
ment. It is possible to study social motivation (for achievement, affilia-
tion, or whatever) without saying that it is based on the hypothalamus and
cerebral cortex. It is possible to investigate the relation between leaders
and followers without hypothesizing that there are organic changes that
explain such behavior. Love is much more than the secretion of sexual
hormones. That human beings make plans for the future and "mental
maps" of their environment does not necessarily need to be explained at
the level of the temporal or occipital cortex. "Brainless" psychology is a
thing of the past, but physiological reductionism is to be avoided as well.
We grant the neurosciences the place they deserve but make it clear that
psychology studies the behavior, in a broad sense, of organisms rather
than only their physiology.
This problem is especially relevant to the present chapter. Social be-
havior has received much attention lately and is often considered the
basis of all human activity. Talk of the nervous, endocrine, or immune
systems does not imply that one can suppose that moral judgement is
located in a specific nook or cranny of the eNS, or that socializing a child
means to create in him or her a specific engram. Patterns of behavior are
what is formed. Although this does not exist without an organic basis
(inorganic things have no mind), the object of psychology is what organ-
isms do or say. In this sense, psychology does not have to be either a
reductionist or a "brainless" science.

10.1 Psychology: Natural Science or Social Science?


The place of psychology in the classification of the sciences has always
been a bone of contention among philosophers. The psychologist's object
of study-the behavior of organisms and their relation to the environ-
ment-doubtless includes biological as well as social factors. The ques-
tion is whether this implies that psychology belongs to the natural or the
so-called social sciences.
First, let us review the differences between natural and social sciences,
in the tradition of Dilthey and Windelband, as nomothetic and idiographic
disciplines, respectively. Nomothetic disciplines-the paradigm case is
physics-study general laws and repeatable events. Idiographic disci-
plines, on the other hand-the paradigm case is history-refer to non-
reproducible facts and individual events. For Dilthey psychology was an
idiographic discipline like history rather than a nomothetic one like phys-
ics. What Wundt demonstrated was that psychology has general, univer-
sal laws just as physics does, and hence that it is more like physics than
like history. This dichotomizing view, due to Dilthey and Windelband and
224 10. The Social Matrix of Behavior

more recently defended by the Frankfurt school, holds that there is a gulf
between the history- and meaning-centered disciplines (social sciences)
such as psychology, anthropology, economics, and the like, and the na-
ture-centered disciplines (natural sciences) such as physics, chemistry,
biology, and so forth. This is tantamount to assuming that human and
animal behavior are not a part of nature, or that the meaning of data and
theories in physics, chemistry, and similar disciplines, is not studied! The
central assumption is one of a radical distinction between humans and
nature, and this dualism is as unfortunate as the mind-body dualism,
which has done so much harm to psychology.
A unified view of science-one that goes beyond positivism and physi-
calism, beyond Popper and Kuhn-is urgently needed. All sciences share
an interest in understanding natural, real-world phenomena. This holds
for physics as well as sociology and any other discipline that uses the
scientific method, whether it deals with physical, biological, or social
phenomena. We construe psychology as a natural science, one very close
to biology, which investigates the behavior of organisms. It can thus be
grouped with other sciences, such as economics, anthropology, and soci-
ology, which also study behavior, although they emphasize particular
kinds of behavior. The concept of such a group of behavioral sciences is
relatively new. It does not imply, however, that other sciences do not
study behavior (e.g., that physics does not study the behavior of matter
and chemistry that of molecules). Clearly, to deny this would be to deny
the existence of these disciplines. Thus, in a broad sense all sciences are
behavioral sciences, just as they are all natural sciences because they
study nature.
Psychology, then, is a behavioral science. It is also a social science, as
well as a natural science and, in particular, a biological science. Histori-
cally, however, the special nature of the problems it deals with-the
"soul," "psyche," or "mind," and behavior-has made psychology es-
pecially puzzling. Physics is clearly a natural science. So is biology, when
it avoids vitalism. Sociology and economics are just as clearly social or
behavioral sciences. However, psychology, with one foot in nature and
the other in society, has been especially difficult to classify using tradi-
tional distinctions. Given the way science has evolved, these distinctions
rooted in eighteenth-century thought are clearly of no rel,Nance to psy-
chology near the beginning of the twenty-first century.

10.2 Culture
For anthropologists, the concept of culture is as important, all-encom-
passing, and difficult to conceptualize as is the concept of behavior for
psychologists. In this respect it is also analogous to the concept of matter
for physicists and the concept oflife for biologists. Indeed, these concepts
10.2. Culture 225

are so important that views of each of them form the conceptual basis for
entire disciplines.
One definition of culture is "the part of nature made by man." It in-
cludes objective culture, that is, human-made objects, and also what
Triandis, V. Vassiliou, G. Vassiliou, Tanaka, & Shanmugam (1972) call
"subjective culture." It encompasses architecture, highways, art, and
science, all of which are part of objective culture. Subjective culture
includes the values and attitudes that give meaning to human actions and
determine notions of ethics and esthetics, as well as attitudes toward the
universe, the people in it, and the great metaphysical problems that we
can neither solve nor ignore. Culture includes both these objective and
subjective products of human action. Both are important for psychology.
Culture is the great social matrix within which we are born, we grow,
and we die. It gives meaning to human action, and we transmit it to our
biological and spiritual descendants (our children and our students). It has
many philosophical, political, and practical implications: It tells us what is
good and bad; how to live and die; how to talk, dress, and love; things to
eat and when to eat them; how to express happiness and sadness; what to
consider desirable and what to detest.
The intriguing thing is that culture is so close that we almost never
notice its presence. As the forest does not let us see the trees, so culture
surrounds us completely. Apparently insects, ants for example, are un-
able to see a whole human or elephant. Similarly, humans cannot grasp all
of culture, which is tacit and need not be studied or questioned. So it is
that one of the most educational experiences someone can have is to visit
a different culture and live with its inhabitants-for example, a year
spent in Saudi Arabia for an Englishman, or one spent in Malawi for an
American.
There are nations in which one need only run for a few kilometers to go
from one culture to another. Strangely enough, people never seem to do
this, and for the most part live in their own cultural millieu as if it were a
jail or a reserve. Living in another culture is important because it demon-
strates clearly the relativity of many of the values and practices we take
most for granted. What are universal and eternal truths in one culture can
be seen as quaint, paroquial beliefs in another. Not everyone rises at 7:00
to have eggs and coffee for breakfast, drives to the office, and returns
home at 5:00 to watch television. There are cultures where insects and
rodents are eaten and where people do not bathe, where a man can legally
have several wives, and a few where a woman can legally have several
husbands. Housing and clothing change from culture to culture, as does
what people expect from life, what they do when a relative dies, and what
they consider a fair salary. The experience of living in a different culture
makes us see our own more objectively and realistically.
Culture thus bears directly on psychological research because what is
normal and abnormal, acceptable and unacceptable in behavior depend on
226 10. The Social Matrix of Behavior

the culture, age, and social class of the subjects studied. Segall, Camp-
bell, and Herskovits (1966), for example, found that even perception
changes with culture. They discovered that visual illusions-phenomena
often studied by psychologists-are not present in "non-carpentered"
cultures. The study was designed to verify whether or not cultural differ-
ences could be of sufficient magnitude to influence perceptual tendencies.
A total of 1,878 persons from 13 non-Western (mainly African) and 3
Western groups were shown 50 drawings, each an example of one of five
geometrical illusions: the Miiller-Lyer, the Sander Parallelogram, two
forms of the Horizontal-Vertical illusion, and a simple perspective
drawing.
Their results supported the hypothesis that a "carpentered world" and
"foreshortening" made the perception of these stimuli relevant. Accord-
ing to the authors, perception depends in part upon ecological validity.
The carpentered world hypothesis says that the human visual system is
generally functional across the species, but that specific conditions affect
how the system is utilized in different ecologies. Such ecology-internal
tendencies may thus be compared across different ecologies. People who
live in carpentered, right-angled environments learn quite early in their
lives to compensate for the "reality" of visual illusions. The learned
habits of visual inference make the typical Westerner probabilistically
more susceptible to the Miiller-Lyer illusion, which has as an eco-cul-
tural analog the frequent arrays of right-angled buildings and boxes. Simi-
larly, the cultural analog of the Horizontal-Vertical illusion is the flat
cornfield. The inhabitants of such terrain consider the vertical line longer,
because an adjustment has to be made for the two-dimensional depiction
of a three-dimensional reality.
These results were quite surprising because perception has always been
viewed as a basic biological function independent of learning. This exam-
ple illustrates the pervasive influence of cultural learning on behavior.
In cross-cultural psychology the researcher begins by identifying some
major themes that differ across cultures, and then explores the anteced-
ents and consequents of those themes. For instance, the concept of time
seems to entail different behaviors, values, and attitudes toward work,
leisure, money, and so forth, as it varies over the cultures of North
America, Latin America, the Far East, Europe, Africa, and Asia. If we
could discover a clear contrast between some of these cultures, which
implies that cultures with similar concepts of time would have other
things in common and would be different from cultures with different
concepts of time, it would be possible to discover what correlates with
given concepts of time (e.g., traditionalism vs. modernism, internal vs.
external control, etc.). We could also discover the main antecedents, for
instance in child-rearing practices, and the main consequences of these
notions with respect, for example, to work, money, leisure, the future,
the past, and so on.
10.3. Social Classes 227

Cross-cultural investigation considers broad cultural differences so that


studies done in various cultural regions can be linked. Usually the minute
details of these differences are downplayed in order to emphasize the
broad patterns of differences necessary for a theory of culture that can be
used by anthropology and many other sciences. This broad strategy and a
worldwide perspective can later be used to identify similar patterns of
differences within cultural areas.
Several decades ago, few could have imagined that human culture is so
varied and heterogenous. Today we know that cultural unity is not a
reality in our world, and that behavior changes accordingly. The laws of
learning, however, are universal and transcend cultural boundaries. It is a
fact, for instance, that organisms learn in accordance with the principle of
reinforcement. What changes from one group to another is the content of
learning, as can be demonstrated by comparing some of the different
social groups that coexist today.

10.3 Social Classes


Social classes have always existed. Some hold that in the future there will
be none, and that one of the goals of a "just" and "egalitarian" society is
to make class differences disappear. Socialist societies insist that they
have already reached this state, although most sociologists maintain that
there are clear class differences that are far from disappearing. Perhaps
the differences are not as marked as in capitalist societies, but they exist.
The existence of social classes implies behavior patterns that differenti-
ate members of one from members of another, and these have important
psychological implications. Middle-class persons behave differently from
members of the lower class. Language also differs across groups, even if
they belong to the same culture. Sports, work habits, likes and dislikes,
sexual behavior all change with social class. Without a doubt it is a very
powerful element of human behavior.
Social stratification is a system of hierarchical arrangement within a
society of status, prestige, resources, privileges, and power. Social
classes are the relatively homogenous groups that share particular levels
of status and access to resources. Each social class tends to develop its
own norms and ideologies.
It is important to point out that an individual can simultaneously have
different amounts of status, and for this reason it is meaningless to talk of
status as a single-valued variable. In general, status is the prestige the
individual has in a given social system. When roles change, status also
changes. Each social role has associated responsibilities, duties, and
rights. Most social roles have reciprocals (father-child, teacher-student)
and the society is made up of these interrelated roles, which are necessary
for the functioning of the social system. Socialization, as we shall ob-
serve, thus involves training children in the roles and norms of society.
228 10. The Social Matrix of Behavior

Much effort has been directed toward measuring socioeconomic status


and its psychological correlates. Socioeconomic status is a continuous
variable that consists of the weighted scores of income, education, and
occupational prestige. Individuals in higher strata of a system have higher
status and prestige as well as greater access to the rewards the society has
to offer.
Differences among social classes have been studied by psychologists
and found to be very important. Attitudes, community involvement, val-
ues, and even family interactions vary with social stratum. Violence is
more common in lower social classes. Personal happiness is associated
with education (as well as with personality factors). High-status persons
tend to speak more and to be listened to more often.
One complex problem in this domain is the interaction of socioeco-
nomic status and intelligence. A positive correlation has been found be-
tween the two that is stronger for adults than for children. In the contem-
porary controversy over the relative influences of genetic and
environmental factors in the origin of intelligence, social class has played
an important role. Perhaps middle- and upper-class people are more intel-
ligent because intelligent people tend to move up the social ladder and less
intelligent ones tend to move down.
In any case, socioeconomic status is an important index of the social
matrix. An upper-class person in Paris has much more in common with an
upper-class person in Los Angeles or Tokyo than with a lower-class per-
son in the same city. Values, norms of conduct, lifestyles, and philoso-
phies of life are very different across classes. Perhaps in the future there
will be no social classes, but the differential prestige associated with
different occupations, differential access to power, and different social
roles will probably remain.

10.4 Socialization
The process of socialization is that of the acquisition of the values, atti-
tudes, and behavior patterns that are characteristic of the culture within
which the individual is born. It is the process of becoming a part of that
culture, but can be more broadly construed as the process oftransforming
the individual into a human being. It might thus be called "humanization"
rather than socialization.
Individuals of our species are born very "prematurely" and the child
leaves the maternal womb only to enter the social womb. It is not possible
for the newborn to survive without the help of the society into which it is
born. Infants would die of hunger, exposure, or disease if it were not for
the social matrix that receives them. This matrix consists of one's mother
and the other members of one's family as well as the cultural context
10.4. Socialization 229

including medicine, nutritional habits, and measures for disease preven-


tion. It is a wide net of support which facilitates the newborn's survival,
feeding, and growing but in return imposes a series of demands.
Different cultures have different types of norms. Though there is a
common substrate which is a function of our human nature, there are
many components of these norms that are learned. As the nervous system
matures, myelinization facilitates and often even enables complex behav-
iors. Without an adequate level of maturity, there is no learning. Beyond
this biological threshold, however, the rest of learning is culturally medi-
ated (Ardila, 1979a). Although the laws of learnirig are universal, the
content of learning is determined by culture. The child learns what the
culture requires it to learn, and this includes certain values, notions of
good and bad, and the like.
Anthropologists often hold that there are no superior or inferior cul-
tures but that all are equivalent in the sense that they accomplish analo-
gous functions and satisfy similar needs: German culture of the 1930s is
no better than the African cultures of today; the Greek culture of the Age
of Pericles is no superior to those of the Amazon. There are no backward
or advanced cultures; all are equivalent. There are no measures we can
use to objectively classify one culture as superior or inferior to another.
This assumption of the equivalence of cultures is central to the battle
against racism and cultural hegemony.
Culture influences the child even before it is born, by way of current
concepts of nutrition, family planning, techniques and ceremonies of
childbirth, and so on. Culture dictates whether parents should or should
not plan their families. The child is born with scientific help or that of
ancestral traditions, and in cultures where children of one sex are prefer-
red, its survival depends in part on the existence of traditions of neglect-
ing or killing newborns of the other.
After the child's birth, the influence of culture is even greater. Routines
of sleeping and feeding are imposed on the infant. Perhaps its needs are
satisfied immediately, or it is seen as necessary to delay gratification of
primary impulses. The child is exposed to the elements or protected from
them, is isolated from infections or kept in a normal environment to
acquire defenses against them.
Socialization is quite complex. The attitudes and behaviors that the
society has traditionally held to be correct and which have withstood the
test of time are those that are transmitted, even if they do not always fit
the current reality. Many things arc learned directly, others by imitation.
In many cases both parents, or just one (usually the mother), or other
members of the family are in charge of direct teaching. The enormous
variability of cultural differences that anthropologists have documented
demonstrates the flexibility of human behavior and the importance of
cultural learning.
230 10. The Social Matrix of Behavior

Many things are learned by imitation. In some cultures, the girl learns
she must look after her younger brothers and sisters, and the boy learns
he must go hunting with his father; the girl acquires customs related to
agriculture and cattle breeding, and the boy those related to war and
hunting. In Western cultures, the boy usually learns to be competitive and
the girl to be supportive; the boy learns to think that it is better to be an
engineer than a nurse, and the girl learns the opposite.
Sex roles have changed with the passage of time. Especially apparent
are the changes of the last 30 years. Women as engineers and men as
nurses have demonstrated that nothing exists in human nature that predis-
poses humans for one kind of work rather than another. The concepts of
masculinity and femininity have changed considerably and androgeny has
appeared as an important alternative to them. We are going through a time
of great social change, and the roles of man and woman in society seem to
have changed most.
In socialization we find direct teaching and imitation which provide for
the child a set of norms of conduct. If we add the fact that one's nervous
system can only process a limited amount of information and that the
child accepts more readily whatever comes from people who provide
reinforcement, it is clear that infancy has a lasting effect on behavior.
Young children do not know any alternative, they just accept what hap-
pens around them without much objection and integrate it into their per-
sonal world. As time passes, they become members of that culture and
transmit their acquisitions to the next generation. In this way culture
propagates itself.
Most socialization takes place within the family. Of course other social
agencies, such as education, advertising, and television are also influen-
tial. But it seems that the family is the principal agent of the child's
socialization, as has been the case for many centuries.
It is important to note that the family has undergone radical changes.
The extended family, so important until a few decades ago, scarcely
exists. Although in the Third World it is still believed that the extended
family has a great deal of input in socialization, in reality it is losing
ground everywhere. It is being replaced by the nuclear family of father,
mother, and children, and today one talks of alternatives to the traditional
family (Ardila, 1980b). In the future, the family will continue to exist but
alongside communes and other alternatives that will play the same social-
izing role as the extended family of yesterday and the nuclear family of
today. It is not true that the extended family has died or disintegrated, but
society has tried to find alternatives to it. This is a very important advance
that will surely continue to evolve and from which all the participants will
benefit, women in particular. Furthermore, there is no return to the ex-
tended family, just as there is no return to being cave dwellers: We must
look ahead.
10.5. Cultural Homogenization 231

10.5 Cultural Homogenization

Many people believe that "strange" cultures will end up disappearing,


that the planet tends toward cultural homogeneity, and that the values of
the Western world will prevail, even in Africa, the Amazon, China, and
the Soviet Union-it is just a matter of time. This seems to be the case in
North America and Latin America. The American nations have a com-
mon history; they were all colonies of Europe and gained their indepen-
dence at around the same time. Then came a difficult period of stabiliza-
tion, in many cases civil wars, until peace and social order could be
established. Today, however, there are enormous differences between
North America and Latin America, and it is interesting to consider how
these groups of nations are evolving, whether in the direction of greater
unity or greater diversity. Most experts on Latir, America hold that the
trend is one of inevitably greater homogenization, that the differences
between the nations of the Western hemisphere are getting smaller and
smaller.
If this is also the trend in Europe, it is much more difficult to discern.
Nevertheless, social changes in Western Europe, (e.g., in the EEC) have
been in the direction of reducing differences. Life in Britain or France,
Denmark or Germany, is not very different. Many of the differences
between Spain and Italy, even between Portugal and France, seem to be
close to disappearing. Many distinctions will end up being just ones of
"folklore," and will be maintained as an affirmation of the cultural indi-
viduality of a country, just as kings have been maintained in government
but without any power. Kings, folkloric dances, and national character
will surely come to be considered a little ridiculous and a little vulnerable,
things of the past and not of the future.
African and Asiatic countries, however, are quite different from the
rest of the world and far from homogenous. Islamic culture may be com-
mon to many nations but they are still quite different from each other. The
cultures of sub-Saharan Africa have many things in common, though the
dissimilarities are probably even greater. There have been large differ-
ences between China and Japan for centuries, but it looks as if they are
evolving in a similar direction.
In all this it is clear that certain elements of Western culture have
tended to prevail. They are the ones we associate with science, its appli-
cations to improvements in human welfare, the concept of efficiency, and
attitudes toward the use of time. The concept of efficiency, like those of
money and time, is of course culture-specific. However, it seems clear
that Western values pertaining to these things have become more widely
accepted. It is unlikely that cultural homogenization of the world will be
possible in the near future, and it is certain that it is not desirable that it
232 10. The Social Matrix of Behavior

happen at all. The contributions of each culture must be maintained and


will surely help the world to be a better place, and help the people in it to
live each according to their own values.
The social matrix to which we have referred in this chapter is omnipres-
ent in human life. Culture is the human-made part, and no doubt the most
important part, of our environment. We have built buildings and theories,
but we have not made mountains or radically changed the seas. For the
people of today and tomorrow, human works will be more important than
the nature we have inherited and which will probably still be here when
we are gone.

10.6 Summing Up
We are social as well as biological beings. Our nervous system and biolog-
ical structure enable us to learn very much and to modify our behavior to
control its consequences. In this sense, the laws of learning can be said to
be universal. On the other hand, the content of such learning processes is
determined by culture. We learn through socialization what the culture
stipulates should be transmitted from one generation to the next. Children
are led to acquire a series of values, attitudes, and patterns of behavior
that make them members of their society. Socialization, then, is a pro-
cess of humanization.
Culture surrounds us completely. It influences our behavior and our
membership in a social class, which in turn entails the social roles we
have to fill and confers status, prestige, and power. The different cultures
of the planet have changed with the passage of time. It is unlikely that
they will become one in the near future. In any case, the interaction
among them is greater and greater. Surely no single culture or single
language will emerge, although the differences between cultures are ever
fewer.
Psychological phenomena are intimately connected to social structure.
Behavior depends to a great extent on the cultural environment in which
we function. In this sense, a study of the conceptual and methodological
foundations of contemporary psychology cannot be complete without due
consideration of social structure. Although all animals are social to some
extent, the human being is the social animal par excellence.
CHAPTER 11

Consciousness

Consciousness, the pride of the righteous and the curse of the sinner,
has baffled countless thinkers over millennia, and is still regarded by
many as an unknown. Thus Johnson-Laird (1983, p. 448) notes: "No one
really knows what consciousness is, what it does, or what function it
serves." It stood at the very center of traditional psychology until it was
banished by behaviorism and reflexology. However, even at the peak of
these two movements the concept of consciousness only went under-
ground. In fact, it was used when distinguishing subliminal from con-
scious perception, or an alert animal from an anesthesized or sleeping
one.
The concept of consciousness is now making a vigorous comeback, and
it is increasingly being admitted that it is a legitimate concern of science.
In fact, a number of investigators have begun to explore it as well as its
dual, namely nonconsciousness. For example, they study nonconscious
learning and recognition as well as nonconscious vision ("blindsight")
and nonconscious recall ("memoryless memory"). (See e.g., Davidson &
Davidson, 1980; Dimond, 1976; Doty, 1975; Eccles, 1966; Edelman &
Mountcastle, 1978; Fernandez-Guardiola, 1979; Griffin, 1984; Hilgard,
1977, 1980; Humphrey, 1983; Ingvar, 1979; LeDoux, Wilson, & Gaz-
zaniga, 1979; Libet, 1965, 1978, 1985; Mandler, 1984; Mishkin, 1982; Oat-
ley, 1980; Poppel, 1985; Schacter, 1985; Shallice, 1972; Tranel & Dama-
sio, 1985; Tulving, 1985a; Underwood & Stevens, 1980; Weiskrantz,
1985.)
The problem of consciousness, like any other tough scientific problem,
has two components: one conceptual, the other empirical. The former
consists in defining the concept (or concepts) of consciousness, either
explicitly or by way of a system of postulates, and in devising models of
conscious processes, as well as of mental processes that fail to emerge on
a conscious level. The empirical problems of consciousness consist in
devising reliable indicators of conscious processes, in finding the so-
called seat(s) or organ(s) of consciousness, and in determining how the
level of consciousness changes in the course of individual development,
234 11. Consciousness

as well as the way it alters as a consequence of changes in internal and


environmental variables.
The two sets of questions, the conceptual and the empirical, are inti-
mately related. In fact it is pointless to try to localize consciousness
unless one has a reasonably clear idea of what it is. And it is foolish to try
to model it unless one has secured a psychological and neurophysiological
data base. The two components of the problem of consciousness may
then be regarded as so many projections of one and the same problem.
In this chapter we shall propose some definitions, explore a few hypoth-
eses, and review some of the experimental material relevant to our defini-
tions and hypotheses.

11.1 Distinctions
To motivate the definitions and hypotheses that will be proposed later on
we shall start by drawing some distinctions. First, we distinguish con-
sciousness from reactivity or sensitivity. All things, whether alive or not,
are sensitive to some physical or chemical agents, though none responds
to all. If we were to identify consciousness with reactivity or sensitivity
we would have to adopt animism or panpsychism, thus giving up the more
or less tacit ontology of modern science.
A second concept to be analyzed is that of awareness. Any animal
capable of identifying or discriminating some (internal or external) stim-
uli, or some of its own actions, can be said to be aware ofthem provided it
can do something to control either the sources of stimulation or its own
reaction to them-not so if it cannot help responding on cue. For exam-
ple, the gazelle that approaches a watering hole in sight ofa pride of lions,
and the rat that accepts an electric shock in exchange for the chance to eat
or to explore a new maze, can be imputed awareness. In short, a test or
indicator of awareness would be the ability to learn new behavior patterns
incompatible with inherited or previously learned ones.
An animal aware of what it is feeling or doing can be said to be self-
aware. It not only walks and feels hungry but also notices that it is
walking and feeling hungry-as suggested by the way it goes about solv-
ing the problems it encounters along the way. On the other hand, certain
neurological patients are confused as to the origin of some of their own
feelings and doings; they are not fully self-aware. (Example: "neglect" or
the failure to recognize a part of one's own body.) Nor are normal adults
self-aware all of the time; we often manage to temporarily forget hunger
or even pain, and we perform many actions automatically.
An animal aware of what it is perceiving or thinking may be said to be
conscious even if momentarily oblivious of some of its own feelings and
doings, or not manifestly responding to some external stimuli that usually
elicit its reaction. The goose that rolls an imaginary egg with the ventral
11.2. Definitions 235

part of its beak is not conscious: Its movements are regulated by a sort of
"motor tape" in its nervous system, and it can do nothing to help moving
that way. On the other hand, the pigeon that looks intently at a rotated
figure to see whether it is the same as the original, in expectation of a
reward, may be said to be conscious: It is monitoring and "manipulating"
some of its own mental states and movements.
Finally, an animal that is occasionally conscious, that sometimes re-
flects upon its own perceptions or thoughts (concurrent or past), and does
not attribute them to something or somebody else, can be said to be self-
conscious. On the other hand, an animal that attributes its own percep-
tions or thoughts to an external object fails to be self-conscious; so is
the person who "hears voices," imputes his dreams to spirits, or claims
to communicate with the dead. Likewise, an individual immersed in a
motor or intellectual task who does not pause to reflect upon what she is
doing or thinking is not self-conscious. She is herself without being con-
scious of her self.
Let us next hone the five concepts introduced so far.

11.2 Definitions
The terms 'awareness' and 'consciousness' are ambiguous (i.e., they des-
ignate several different though related concepts). The least demanding of
these concepts is that of sensitivity of reactivity to some stimuli. It may be
chancterized by the following:
Definition 1. Let b denote a thing (living or nonliving) and X an action
on b or on a part of b, and originating either outside or in a part of b. Then
b is X-sensitive (or X-responsive) if, and only if, b reacts to X (i.e., if X
causes or triggers a change in the state of b), either always or with a
certain probability.
Photosensitivity, chemical sensitivity, and the ability to respond to
social stimuli are examples of specific sensitivity or reactivity. Obviously,
the proverbial pinprick is not an adequate test of consciousness, for even
the lowest grade of consciousness requires far more than the capacity to
react to external physical or chemical stimuli.
Next comes the concept of awareness, which refers only to animals of
certain species. We define it as follows:
Definition 2. If b is an animal, b is aware of (or notices) change X
(internal or external to b) if, and only if, b feels (senses) X-otherwise b is
unaware of X.
Awareness requires neither more nor less than neurosensors of some
kind. (Hence plants and animals lacking neurosensors cannot be aware of
anything. Not even the sea urchin can be aware of anything, for it lacks
sense organs. A fortiori, machines cannot attain awareness, although, if
236 11. Consciousness

equipped with suitable "sensors," such as photocells, they can react to


certain stimuli.)
Now, an animal may be aware of its surroundings but not of what it is
feeling or doing. If it is aware of what it feels and does, it can be said to be
self-aware. Therefore we also need
Definition 3. If b is an animal, b is self-aware (or has self-awareness) if,
and only if, b is aware of some of its inner changes and actions.
To be self-aware is to be aware of oneself as something different from
everything else. A self-aware animal notices, however dimly, that it is the
subject of its own feelings and doings. Self-awareness is so much taken
for granted that we tend to forget that we are not aware of ourselves when
absent-minded, and that it can be a hindrance when working, which re-
quires other-awareness.
Note that self-awareness does not require thinking about one's own
perceptions or conceptions. Satisfaction of this additional condition quali-
fies as consciousness. Therefore we stipulate
Definition 4. If b is an animal, b is conscious of perception or thought X
in b itself if, and only if, b thinks of X-otherwise b is not conscious of X.
According to this convention, an animal can only be conscious of some
of its own higher mental processes: Not just feeling, sensing, and doing,
but also thinking of what it perceives or thinks. (To be sure, thought need
not be abstract or even verbalizable; it can be in images, as when we
perform a mathematical calculation imagining that we are writing on a
blackboard.) An animal conscious of mental process X in itself undergoes
(either in parallel or in quick succession) two different mental processes:
X (the object mental process or content of its consciousness), and think-
ing about X (i.e., being conscious of X). The object of X can be a percep-
tion (e.g., of a hot pan), a memory (e.g., of a tasty sausage), a theorem, or
what have you.
Note the difference between consciousness and awareness. The ani-
mals of certain species can become aware of certain stimuli, and many are
capable of attention, but they cannot be conscious of anything unless they
can think. Conversely, a person lost in daydreaming or in deep, produc-
tive thought may be unaware of his or her surroundings. Consequently,
the concepts of consciousness and awareness are mutually independent.
This being so, they should not be confused. And the hybrid 'conscious
awareness' should be avoided.
All consciousness is consciousness of something. This something is
called the content or object of consciousness. (Consciousness without a
content, as in the "nirvanic" state attained in Zen "meditation," is not
consciousness at all; it is merely a state of mindlessness.) Hence we
propose
Definition 5. The content (or object) of a conscious state is the object
being perceived or thought about while in that state.
11.2. Definitions 237

Being conscious of a mental process in oneself is to be in a certain


mental state-which, according to physiological psychology, is the same
as the brain being in a certain state (or rather undergoing processes of a
certain kind). Hence consciousness, which in traditional psychology (in-
cluding psychoanalysis) is conceived of as an entity, is better regarded as
a collection of brain states. For this reason we adopt
Definition 6. The consciousness of animal b is the set of all the states of
the brain of b in which b is conscious of some perception or thought in b
itself.
Just as there is no consciousness as an entity, so there is no entity that
can rightly be labeled The Unconscious. Instead, there are simply some
mental processes that remain nonconscious or preconscious, even though
they can occasionally be manifested behaviorally and thus become the
subject of scientific investigation. More on this in section 11.5.
Just as self-awareness is one rung higher than awareness, so self-con-
sciousness is one step higher than consciousness. A subject is self-con-
scious only if he has consciousness of his own perceptions and thoughts
as occurring in himself. At first sight the term 'self-consciousness' is a
pleonasm. However, there is solid clinical evidence that subjects in cer-
tain pathological conditions are confused about the source of some of
their own mental experiences and even actions. Therefore we need
Definition 7. An animal is self-conscious if, and only if, it knows who
and what it is.
Now, in order for one to know who and what one is, one must have
some recollection of one's past: We are what we have become, and we
know what we have learned. On the other hand the animal need not be
able to extrapolate its own life into the future: It may not be capable of
imagining or planning its next move. Thus, the primate with a frontal
lobotomy appears to be self-conscious from moment to moment, but "the
stream of happenings is not segmented and so runs together in a present
which is forever, without past or future. The organism becomes com-
pletely a monitor at the mercy of his momentary states, instead of an actor
on them" (Pribram, 1971, p. 348). These data call for the following dis-
tinctions:
Definition 8. A self-conscious individual is
(a) antero-self-conscious if and only if he recalls correctly some of his
recent past;
(b) pro-self-conscious if and only if he can imagine (even wrongly) his
own future, and
(c) fully se(f-conscious if and only if he is both antero- and pro-self-con-
scious.
Note that our definitions do not contain the concepts of attention, inten-
tion, intentionality, the ability to report on one's own mental state, the
238 11. Consciousness

ability to form representations or models of the world, or information


processing. The reasons for these exclusions are as follows.
An animal can be aware of a part of its surroundings even while not
being attentive. Conversely, it may be prepared or "set" to perceive
something and yet fail to detect the event it is expecting. In short, atten-
tion is neither necessary nor sufficient for awareness.
On the other hand, intention does seem to call for awareness, though
not necessarily consciousness. For example, we may be intent on
performing a certain task in an automatic fashion. This requires focus-
ing on the task itself rather than on our "stream of consciousness."
For this reason the attempt to define consciousness in terms of inten-
tion or purpose is misguided. It may be just a vestige of the time
when psychology, whether academic (as in the case of MacDougall)
or popular (as in that of Freud), was dominated by animism and tele-
ology.
Brentano (1874) regarded "intentionality", which he characterized as
"the reference to something as an object," as the mark of mental phe-
nomena, which he placed in "inner consciousness." But he did not define
the latter and he overlooked preconscious mental processes, which many
before him (e.g., Hume) had taken for granted. Moreover his use of
"intentionality" as different from "intention" and identical with "refer-
ring" is misleading, for most mental processes are unintentional. Let us
keep "intention" as a psychological concept and "reference" as a seman-
tic one.
As for the ability to report on one's mental processes, we cannot use it
to define either "awareness" or "consciousness," for two reasons: First,
because it does not apply in an obvious way to animals other than talking
humans-whence that definition would block research into the question
of animal consciousness. Second, because the ability to report on one's
own mental process is sufficient but not necessary for consciousness. In
fact, a fully conscious subject may be so anxious or angry as to be incapa-
ble of informing others about his or her own mental processes. The ability
to report on one's own mental states cannot be used as a definiens of
"consciousness," although it can be used as an uncertain indicator of
conscious processes-and it is so used by clinical psychologists, psychia-
trists, and neurologists.
Nor can consciousness be equated with the capacity to form represen-
tations or models of the world-pace the opinion of some distinguished
students of the matter. Such equation is erroneous for two reasons. First,
it implies that photographic cameras are endowed with consciousness.
Second, it is quite possible that all animals possessing a nervous system
have the ability to form models of their environment, for without them
they could not "navigate" in it. (See von Uexkiill, 1921.) Consciousness
is a very special kind of knowledge that seems to be the prerogative only
of highly evolved mammals.
11.3. Applications 239

Finally, we have made no use of the notion of information processing,


for it is vague and far too comprehensive: Recall section 5.4. At best, the
information-processing approach suggests rather nondescript black boxes
and "flowchart metaphors" (Miller, 1980) that, with luck, might sketch
some of the connections among subsystems of the nervous system. (Ex-
ample: the ingenious but suspiciously simple diagram proposed by Shal-
lice, 1972, to represent consciousness.) At worst, the information-pro-
cessing approach conflates all psychological phenomena, thus depriving
the conscious ones of their pride of place.

11.3 Applications
Let us now put the preceding definitions to work in redescribing some
mental abilities and deficits.
Animal awareness. All animals can be aware of some external and
internal stimuli, and some may become self-aware at an early age. Fur-
ther, it may be conjectured that all mammals and birds can be in conscious
states, however dimly and infrequently, but we do not know for sure
because no physiological indicators of consciousness have so far been
devised.
Development of consciousness. Humans seem to become aware of
themselves (i.e., self-aware), by age 2. By age 6 they usually have become
conscious (i.e., aware of what they perceive and think). In fact, a normal
child of 6 will answer unhesitatingly whether or not he or she is thinking of
something definite, such as a scene from a film or a raid on the cookie jar.
Infantile amnesia. Although early experiences contribute decisively to
forming our personalities, we hardly recall them. Most of our early learn-
ing is preconscious: Consciousness develops only later on and gradually.
Another way of putting this is to say that the infant brain lacks the mecha-
nism of episodic memory. In fact it is so primitive that it cannot possibly
repress anything.
Subliminal perception. The well-known findings on subliminal percep-
tion (e.g., Dixon, 1971) confirm the hypothesis that consciousness comes
in degrees: nil, dim, normal, and heightened. 'Subliminal perception' is
just another name for nonconscious perception, a process in which we are
aware but not conscious of what we sense. This can be generalized: We
can learn (acquire knowledge) without being conscious, or even aware, of
doing so.
Blindsight. Patients who have sustained damage to their primary visual
cortex experience a blind spot or area (scotoma), or even total blind-
ness-or so they report. Actually they may have considerable residual
vision without being conscious of it, whence they cannot describe the
objects they actually see with their "second sight." This w~s demon-
strated by shining light on those "blind spots" and asking the· ubjects to
240 11. Consciousness

guess what was being shown them. Most of the time they guessed cor-
rectly (Poppel, Held, & Frost, 1973; Weiskrantz, 1980). These then are
cases of unconscious vision, and they suggest the existence of a second
visual system. (See Stoerig, Hubner, & Poppet, 1985.) "Blindtouch"
(Paillard, Michel, & Stelmach, 1983) is parallel. One may conjecture that
there are similar phenomena in the remaining modalities, so that one may
speak in general of nonconscious perception. (The difference with sublim-
inal perception is that in the latter there need be no definite anatomical
lesion, so that it must involve different neural systems.)
Prosopagnosia. A stroke, a tumor, or some other insult to the inferior
medial occipito-temporal cortices and their connections can cause a loss
of the ability to recognize familiar faces, animals, automobiles, and the
like. This impairment of object recognition is called 'prosopagnosia.'
However, when shown a picture of a close relative or friend, a proso-
pagnosic generates a far larger skin conductance response than when
shown an unfamiliar face (Tranel & Damasio, 1985). This objective test
shows that the patient does identify the face without being conscious: this
is another case of nonconscious perception. This is one of many findings
suggesting that the process offace recognition, and perception in general,
is a multiple-stage process, only the last of which consists in the monitor-
ing or readout of the preceding ones (A. R. Damasio, G. W. Damasio, &
Van Hoesen, 1982; Treisman & Gelade, 1980).
Loss of short term memory. Because every mental process takes some
time, however short (usually at least 10 msec), we must ask whether
memory is necessary for self-consciousness (i.e., for the concurrent mon-
itoring of some mental processes). One could not possibly reflect on what
he had been thinking only a short while ago unless he could recall that
thought. Individuals who have lost their short term memory, like the
famous H.M. so thoroughly studied by Brenda Milner (1959), though
possibly conscious, are not fully self-conscious even while chatting intelli-
gently with their examiners. Likewise, the consciousness ofN.N., stud-
ied by Schacter and Tulving, is severely impaired although he can pro-
duce a pretty good definition of "consciousness" (Tulving, 1985a).
Something similar holds for patients in advanced stages of Korsakoffs
"psychosis" or Alzheimer's disease; they are not fully self-conscious
because their memory span is so short that they cannot reflect upon their
own mental activity. These patients have also lost much of their explicit
or declarative knowledge (vs. know-how), including knowledge of them-
selves-they hardly know who they are.
Declarative (explicit) versus procedural (tacit) knowledge. Amnesic pa-
tients are worst at performing tasks requiring explicit knowledge (or
know-that), and best at those requiring tacit or procedural knowledge (or
know-how), such as driving and speaking. Moreover "conscious recollec-
tion is neither useful nor necessary" for the performance of a number of
skilled tasks (Schacter, 1985). In amnesic patients the correct response
11.3. Applications 241

"is produced more or less automatically by the cue, i.e., it is produced


acognitively" (Weiskrantz, 1982). The patient is only dimly conscious.
Dreaming. Whether asleep or awake, the dreamer is usually unaware of
her surroundings as well as of her own feelings and actions, if any; that is,
ordinarily she also fails to be self-aware. However, the dreaming subject
may sense some sensory stimuli or visceral signals, in which case the
content of her dream may take a sharp turn or even stop. Moreover, the
dreamer is occasionally conscious of her own thought (imagery) process
(i.e., she can be self-conscious). In fact, she may reflect upon her own
ongoing dreaming, to the point of wondering whether she is asleep or
awake. Dreaming, in short, is not a case of total lack of awareness and
consciousness but rather one of abnormal and altered awareness or con-
sciousness.
Orgasm. This state too may be regarded as an altered state of aware-
ness and consciousness (Davidson, 1980). During orgasm human subjects
"lose touch with reality" as well as control over themselves. To some
extent they even lose their sense of identity; they feel at one with their
partner. (Apparently this experience is essentially the same in the two
sexes.)
Altered states of awareness and consciousness. Cerebral injuries and
drug overdoses can cause a variety of abnormal states of awareness and
consciousness, from mere dimness (obnubilation) to confusion and from
delirium to coma. In cases of tissue damage, the degree of abnormality is
roughly proportional to the amount of brain tissue destroyed (Plum &
Posner, 1980)-which suggests literally measuring the degree of con-
sciousness with a scale.
Experimentally induced alterations of awareness and consciousness.
Certain drugs, as well as the electrical stimulation of certain brain regions
(e.g., the cortex and the limbic system) can distort a subject's mental
processes and, a fortiori, his own thoughts about them. For example, the
subject can suffer visual or auditory hallucinations, and think "intruding
thoughts" over which he has no command. He can be self-conscious and,
a fortiori, conscious, but he is not his usual self. Both his consciousness
and his self-consciousness are distorted. If preferred, his thinking about
his own perceptions and conceptions has become distorted.
Voluntary extinction of consciousness. Consciousness can be voluntar-
ily dimmed to the point of extinction, by falling asleep or by means of
certain "techniques" that boil down to the suppression of thought. The
various "meditation" practices may be characterized as so many proce-
dures for the self-regulation of consciousness. (In particular, the "nirva-
nic state," which physiologically is hardly distinguishable from dreamless
sleep, is attained through stopping the stream of consciousness alto-
gether.) These practices are part of certain cults and have given rise to a
flood of escapist, parascientific, and mystic literature. (See e.g., Ornstein,
1973.)
242 11. Consciousness

Split-brain. The consciousness of normal subjects is unitary; that of


split-brain patients is divided (Gazzaniga, 1967). In other words, cutting
the brain in two results in two minds and, in particular, in double con-
sciousness. Not surprisingly, the ordinary split-brain subject is confused
about what is happening to her while she experien.::es two simultaneous
streams of consciousness. She is notfully self-consc:ious, and she now has
two selves: she has become they, and "they" have a hard time coping
with one another, because they cannot communicate directly.
Causal efficacy of conscious processes. According to epiphenomenal-
ism (section 1.2) consciousness has no causal efficacy because the mind is
"secreted" by the brain. Accordingly, self-consciousness would only be a
sort of dashboard of the brain monitoring some brain processes. Ordinary
experience refutes this passive view of consciousness. In fact, conscious-
ness can steer behavior, whence it is a sort of driving wheel in addition to
being a dashboard. For example, we can learn to keep our eyes open for
eye drops, and to master the gag reflex while swallowing a probe. We can
thus learn to control some reflex actions, not only voluntary movement.
In section 7.2 we reviewed the experiments of Melvill Jones, Berthoz, &
Segal (1984) showing that we can weaken certain reflexes through the
conscious performance of certain mental operations. From a biological
viewpoint this is one more example of the action of one part of the central
nervous system upon another.
We shall wind up this section by reviewing a few nonexamples of con-
sciousness that sometimes pass for instances of it.
Divided attention. Sometimes we succeed in dividing our attention
among two or more tasks at the same time, as when arguing with some-
body while walking and pushing a stroller. Divided attention is sometimes
called 'divided consciousness' (e.g., Hilgard, 1977), but this seems an
unnecessary complication deriving from a lack of clear definitions. In
order to perform the just-mentioned complex task all we need is to be
aware of our interlocutor's speech and of our own actions. We need not
even be conscious of the way we plan our replies. In fact, as Hilgard
(1977, p. 2) himself points out, the planning of the answer "may continue
without any awareness of it at all. When that appears to be the case, the
concealed part of the total ongoing thought and action may be described
as dissociated from the conscious experience of the person." So, this is
not a case of divided consciousness after all, but one of divided mind;
there are several par~ el processes going on-everyone of them
in a different brain su;· ystem-only one of which may (but need not)
surface.
Dual personality. Some mental patients have two different personali-
ties-though not at the same time. Dual personality does not support the
hypothesis that consciousness is multiple rather than unitary. It only
illustrates the changeability of the self, which traditional psychology re-
garded as constant.
11.4. Hypotheses 243

Hypnosis. A subject under hypnosis is said to be capable of performing


certain tasks "without awareness of what he is doing." However, all we
know is that, once out of hypnosis, the subjects have no clear recollection
of what they did, felt, or thought while in it. Accordingly, hypnosis might
not be a case of loss of consciousness, or even awareness, but one of
temporary loss of episodic memory in the sense of Tulving (1983). The
mere possibility that this alternative account of hypnosis is true suggests
the need for honing the key concepts of awareness and consciousness
before proceeding with experiments on hypnosis.
Machine consciousness. Some programmable machines are occasion-
ally attributed not just minds but also conscious states. This holds, in
particular, for computers guided by programs containing executive sub-
programs, the function of which is to monitor and control the perfor-
mance of the programmed task. Actually this is only an analog of con-
sciousness. Any artifact equipped with measuring instruments that
monitor its own states could be said to have an artificial consciousness.
The literal reading of such metaphors is inconsistent with the naturalistic
thesis that mentality is a biological function that can only be imitated (and
even so partially) by machines (recall section 5.4).
Let the preceding examples and nonexamples suffice to motivate and
justify the conjectures to be proposed anon.

11.4 Hypotheses
We now propose a few general and rather mild hypotheses concerning
awareness and consciousness. We start with a restatement of the psycho-
neural identity thesis:
Hypothesis 1. All mental processes, whether or not conscious, are
brain processes and, more precisely, processes occurring in plastic re-
gions of the brain.
Hypothesis 2. Awareness and consciousness of any kind are specific
(exclusive and nonroutine) activities of distinct subsystems of the central
nervous system of a highly evolved animal.
This hypothesis is suggested by neurological data such as the following.
Damage to the parietal lobes can cause loss of body awareness but not
necessarily loss of consciousness proper. In particular, a parietal lobe
patient may be able to reflect on his own perceptions and thoughts so long
as these do not refer to the part of his body that he does not acknowledge
as his own. (The dual of neglect, namely phantom limb pain, is sometimes
successfully treated by surgery on a parietal lobe.)
Hypothesis 3. All conscious processes occur in the neocortex in con-
junction with some subcortical systems, particularly the thalamus, the
hippocampus, or the amygdala.
This conjecture contains the popular view that conscious processes
occur, at least partly, in the frontal lobes (e.g., Ingvar, 1979; Luria, 1973).
244 11. Consciousness

But it also states, contrary to the standard view, that the cortex does not
suffice for consciousness. One piece of evidence for our conjecture is the
finding that full-blown amnesia requires damage to at least part of the
limbic system (Aggleton & Mishkin, 1983; Mishkin, Spiegler, Saunders,
& Malamut, 1982; Saunders et aI., 1984). Because dense amnesia is in-
compatible with self-consciousness (section 1.3), it may be conjectured
that an intact limbic system is required for full self-consciousness-
though not for plain consciousness, let alone for mere awareness. We may
speculate that conscious processes are identical with interactions be-
tween a cortical system and the thalamic and limbic systems. And, given
the variety of cortical regions, we may also speculate that there are as
many neural (cortico-thalamic and cortico-limbic) consciousness systems
as kinds of conscious process-motor, perceptual, cognitive, and so
forth. These speculations account for the great variety and shiftiness of
conscious experience.
Hypothesis 4. A conscious process is one whereby one part of the brain
(call it C) records or controls perceptual or cognitive processes that occur
in another part (call it N) of the same brain.
In other words, consciousness requires two distinct but connected neu-
ral systems: See Figure 11.1. It follows that, if these systems are discon-
nected, whether temporarily or permanently, the corresponding con-
scious experience is interrupted. This would explain the momentary loss
of consciousness in deep sleep or as a result of a concussion. It would also
help explain blindsight, loss of episodic memory, nonconscious learning,
and the like. (Recall section 11.2.) All these would be disconnection syn-
dromes, hence basically similar to some aphasias, agnosias, and apraxias
(Geschwind, 1965), as well as to some amnesias (Warrington &
Weiskrantz, 1982).
Hypothesis 5. Consciousness comes in degrees:
(a) For every kind of mental process there is a threshold below which the
monitoring system C is not activated.
(b) The intensity of conscious process c equals the intensity of the specific
(nonroutine) activity of the neural system C that monitors the activity
of a plastic neural system N (other than C) that has afferences to C.
(Hence the definition: An animal is conscious of process n going on in N
if and only if N acts upon C.)
This hypothesis-a refinement of that of James (1890)-accounts for
subliminal experience as well as for the dimming of awareness and con-
sciousness (e.g., when falling asleep) and their heightening (e.g., when
performing a difficult task for the first time).
Hypothesis 6. Consciousness surfaces and submerges:
(a) Unlearned behavior may become conscious, and
(b) Learned behavior, if initially conscious, may become nonconscious
(automatic).
11.4. Hypotheses 245

FIGURE 11.1. A possible mechanism of consciousness. Hypothetical neuron as-


sembly C records the activity of neural system N. Subject is aware of activity in
N, or in muscles innervated by N,just in case N stimulates C, or C controls N. In
the first case, C monitors N (dashboard metaphor); in the second case C exerts a
causal action on N (steering wheel metaphor). Adapted from Bunge (1980).

This hypothesis summarizes our knowledge of the gradual emergence


of consciousness during the learning process, as well as of the automatiza-
tion of motor, perceptual, and conceptual tasks the original learning of
which took a conscious effort.
Hypothesis 7. Consciousness is causally efficient:
(a) All conscious processes have some effect on other brain processes,
some of which may in turn have motor outcomes.
(b) Some conscious processes can control certain reflexes, sometimes to
the point of temporary extinction.
This hypothesis implies the rejection of epiphenomenalism (recall sec-
tion 11.3) as well as the admission of "mind power" -though conceived
of as the action of one bit of matter upon another, rather than as the
victory of the immaterial spirit over passive matter.
Hypothesis 8. Awareness and consciousness develop gradually:
(a) The newborn can be aware but not conscious.
(b) Awareness becomes finer, and consciousness emerges gradually, as
the animal matures and learns.
(c) The development of self-consciousness is accelerated by interactions
with conspecifics.
This hypothesis condenses some of the findings of developmental psy-
chology concerning the process whereby the young child gradually learns
to draw the line between herself and the rest of the world (i.e., the process
of formation of the self).
Hypothesis 9. Consciousness and self-consciousness are products of
evolution:
(a)Consciousness emerged naturally at some stage in the evolution of
some higher vertebrates.
246 11. Consciousness

(b)Self-consciousness emerged at some stage in the evolution of the


social behavior of some higher vertebrates.

This hypothesis sums up the little we know, or rather suspect, about


mental evolution. The first part of it is at odds with supernaturalism. The
second part harmonizes with the conjectures of Vygotsky (1978), Luria
(1976), Humphrey (1983), and a few others about the social origin and
function of the higher mental functions. Self-consciousness is part of our
knowledge of ourselves, and it is particularly useful in social behavior. In
fact, my knowledge of others, and my ability to get along with them or to
modify their behavior, derives to a large extent from an analogy with
myself: I model others after myself, and in this way I can empathize with
others and foresee (or believe to foresee) some of their behavior. Self-
consciousness also is, to a large extent, a product of social intercourse: I
am the more self-conscious, the more intensely I fear or hope that my
behavior will be watched and judged by my fellow humans. Thus con-
sciousness, and particularly self-consciousness, are both effect and cause
of social behavior. This is why the present chapter follows immediately
that on the social matrix, rather than chapter 9 on the higher functions.
Should it be objected that there is no empirical evidence for Hypothesis
9, we would reply that there is as much circumstantial evidence for it as
for any other evolutionary hypothesis not supported by fossils or by
genetics. Besides, the first part of the hypothesis jibes with evolutionary
biology and the naturalistic world view inherent in modern science, and
its second part encapsulates some of the findings of social psychology and
cognitive ethology. However, there is no denying that Hypothesis 9 is
rather sketchy, and that the same holds for Hypothesis 8 on individual
development. The reason for this imprecision is that our knowledge of
evolutionary and developmental processes is still embryonic. (Remember
that the scientific study of mental processes in babies started only very
recently, and that evolutionary psychology is hardly more than a research
program.) It is necessary to speculate on the precise stage of development
in a given species, or the precise evolutionary stage, at which conscious-
ness and self-consciousness emerge, while remembering that any such
speculation must be consistent with the bulk of our antecedent knowl-
edge, and that it should eventually be supported or undermined in a more
direct way by empirical data.
Hypothesis 10. Self-consciousness and language are coeval and they
have coevolved alongside sociality.
This hypothesis is suggested in part by our daily experience with silent
(internal) speech. It may be conjectured that the primitive system of
animal communication that eventually evolved into human language al-
lowed hominids to internalize certain aspects of their social behavior. As
their system of communication was perfected, it became more than a
means of social intercourse, namely a tool of self-analysis and a carrier of
11.5. Experimental Tests 247

prefabricated chunks of thought that could be summoned and combined


almost at will. Eventually our ancestors became able to talk to them-
selves, i.e., to internalize conversations, some of which must have re-
ferred to their own perceptions and thoughts.
Thus, self-consciousness and language may have co evolved by a feed-
forward mechanism. This implies of course that our hominid ancestors
enjoyed and suffered a sort of dim self-consciousness. Note that our
coevolution hypothesis differs from the speculation that all mental pro-
cesses are a byproduct of speech (Luria, 1973; Vygotsky, 1978). It is also
at variance with the converse conjecture, that language is a product of
(immaterial) mind (Popper & Eccles, 1977).
The preceding 10 hypotheses might serve as heuristic tools for theory
construction and experimental design. Whatever may survive ofthem will
eventually have to be cast in mathematical terms as part of some neuro-
physiological and sociopsychological theories of the higher mental func-
tions.

11.5 Experimental Tests


The first problem faced by the experimentalist wishing to test any clearly
formulated hypothesis about an unobservable property or process, such
as a mental one, is that of devising faithful indicators or objectifiers of the
property or process in question. The second problem is that of quantitat-
ing at least the measurable manifestations of the property or process
under investigation. These two problems are intertwined, and the poorer
the relevant conceptual frameworks and theoretical models, the harder
the problems are. This is precisely the problem with experiments on
awareness or consciousness at the present time; they are difficult to de-
sign and even more difficult to interpret, because our ideas on awareness
and consciousness are still hazy. The situation is reminiscent of that of
physics in the early days of electricity and magnetism.
The two problems-those of finding reliable indicators and of quantitat-
ing-are often compounded by the operationist philosophy still prevailing
in psychology and social science. According to this philosophy all con-
cepts are to be "defined" by empirical operations (in particular measure-
ments), and unobservable properties and events are to be identified with
their basic observable manifestations (indicators). This is why awareness
has occasionally been "defined" by its objectifiers, such as key pressing
and verbal reporting (e.g., Yates, 1985). For the same reason, the verbal
system (or speech "center") has been regarded as the seat or organ of
consciousness: because it is the only neural system capable of reporting
on our mental states (e.g., LeDoux et al., 1979). But this is like "defining"
(not just measuring) the severity of a common cold as the frequency of
248 II. Consciousness

sneezing, or attention as the percentage of correct identifications of cer-


tain symbols mixed with distractors.
A well-known experiment purporting to demonstrate consciousness in
rats (Beninger, Kendall, & Vanderwolf, 1974) illustrates these difficulties.
The rats were conditioned to associate a certain natural behavior, such as
face washing, with pressing a lever that delivered reinforcement. Every
time the animal engaged in behavior B and pressed the right lever L it was
rewarded with a food pellet. The authors interpreted this instance of
operant conditioning as similar to the verbal reports made by humans
about their own internal states and their actions, e.g., "I have a tooth
ache," or "I am thinking of the concept of thought." However, because
the particular lever associated with the given behavior was chosen arbi-
trarily by the experimenter, the learned "discrimination" (actually pair-
ing) need not exhibit rat consciousness. There is no knowing whether,
when pressing the correct lever L every time it performed behavior B, the
animal was telling the experimenter: "Watch, I am doing B." We only
know that the rat learned to pair Band L. The experiment tells us nothing
about the animal's mental processes, because neither B nor the pressing
of L are independently known to involve mental processes, even though
we may be inclined to believe that they do. The experiment does not
involve an action, such as pressing a second lever, that could be inter-
preted as the rat's "running commentary" upon what it was doing with
the first lever. (See Weiskrantz, 1985.)
Another, even better known set of experiments concerns the behavior
of animals looking at mirrors. The tacit hypothesis is that an animal that
gives signs of recognition of itself is self-aware, whereas if it does not it
cannot be attributed self-awareness. It seems that, whereas chimpanzees
and a few other animals do recognize their own mirror images, dogs and
other animals do not. But this is no proof of self-awareness; in some cases
it may only indicate that recognition in a mirror takes learning. As a
matter of fact, this is exactly what Gallup (1977) found in a study of young
wild-born chimpanzees. It took them two or three days to use the mirror
to groom themselves. This suggests that, before the experiment, they only
lacked a visual self-image; it does not prove that they lacked self-aware-
ness. It would be interesting to repeat the experiment with tribespeople
who had never before seen their faces reflected in a mirror. Maybe they,
too, would fail at first to recognize the reflected faces as their own.
Several severe difficulties face the design of experiments aiming at
testing the hypothesis that animals of a given species, or at a given stage
of their development, can be in conscious states. One of them is the lack
of consensus on what consciousness is, as well as on what the best physi-
ological and behavioral indicators of it might be. Another difficulty, to
speak anthropomorphically, is to make the animal "see the point of the
experiment" by "asking him questions" that he is likely to understand
11.5. Experimental Tests 249

and that will motivate him to make an effort to answer (Weiskrantz, 1977).
The same holds, mutatis mutandis, for human infants. Many failures to
demonstrate certain cognitive abilities in human babies and in certain
animals turned out to be failures of the experimenter's imagination in
designing the proper experiments.
What holds for consciousness holds, a fortiori, for its dual. Consider,
for example, Freud's conjecture that all slips of the tongue are pranks of
the Unconscious. Motley (1985) found that human subjects made anxious
by a (phony) threat of electric shock are likely to commit more spooner-
isms (e.g., 'pobody is nerfect' for 'nobody is perfect') than are relaxed
subjects. But presumably, anxious individuals are prone to make more
mistakes of all kinds-motor, perceptual, conceptual, linguistic, and
combinations thereof-not just speech errors. The same is true of the
occurrence of slips of the tongue with a seeming sexual content (e.g.,
'bared shoulders' for 'shared boulders ') in the presence of a provocatively
dressed person of the opposite sex. Such errors can be interpreted as
either confirming Freud's conjecture or as indicating a general cognitive
impairment caused by an emotional disturbance. Presumably, if a threat-
ening tiger were substituted for an attractive person of the opposite sex,
subjects would occasionally say 'dapper tiger' for 'happy trigger,' or
would utter nonsense expressions or even curses. As well, it isjust possi-
ble that the alleged sexual content is mainly in the experimenter's mind;
after all, this is what he is looking for. Maybe he should be the subject of a
second experiment.
A good experimental design to test Freud's hypothesis on slips of the
tongue should include not only the eliciting of spoonerisms but also of
motor, perceptual, and conceptual errors made by anxious, excited,
bored, absent-minded, or tired subjects. But even this control would be
insufficient, for every scientific hypothesis should cohere with the bulk of
antecedent knowledge (Bunge, 1967a, 1967b). In particular, Freud's hy-
pothesis should be shown to harmonize with our general (albeit largely
nonscientific) knowledge of human error-which is plainly not the case
with Freud's fantasy. After all, there is such thing as plain error, in
particular the misarticulation of phoneme sequences. And this might be
explained in terms of unusual interneuronal connections. (See e.g., Dell,
1985.)
In short, Motley's experiments do not involve all of the necessary
controls. Consequently, his results do not confirm even the watered-down
version of Freud's hypothesis, that at least some slips ofthe tongue are to
be blamed on the unconscious. Nor do they refute the alternative hypoth-
eses that stress impairs all of our "faculties," and that a number of speech
errors (including some with an ostensive sexual content) are merely
results of more or less random departures from the normal neural connec-
tions and activation sequences.
250 II. Consciousness

11.6 Summing Up
We have found it necessary to distinguish a number of concepts, such as
those of sensitivity, awareness, consciousness, and self-consciousness,
that are often conflated in the literature. (Tulving, 1985a, is an exception.)
We have tacitly assumed not just that we are dealing with different de-
grees of a single capacity (like, say, visual acuity), but that they are
qualitatively different phenomena. Therefore it is likely that different
kinds of awareness and consciousness are imputable to different neural
systems. For example, awareness of a visceral process is likely to be an
activity in a neural system distinct from that corresponding to conscious-
ness of a thought. This hypothesis is neither idle nor wild: It might be
testable with the help of biopsychological techniques.
Our definitions have helped us formulate some more or less speculative
hypotheses. They should help construct many more (e.g., about the delib-
erate construction of cognitive strategies, and about the automatic appli-
cation of the latter to the solution of routine problems). Regrettably, there
is a dearth of well-formulated and empirically well-confirmed, or at least
testable, hypotheses and theories about awareness and consciousness.
Consequently, there are not enough well-designed experiments on these
phenomena. No wonder, for there is no good experimental design and no
correct interpretation of experimental data in a conceptual vacuum-
even less in a conceptual muddle.
The behaviorists' neglect of consciousness and their mistrust of theo-
ries, as well as the naive reductionism of reflexology, are partly to be
blamed for this lamentable situation. However, other schools are just as
responsible for it. In particular, the intuitionists and holists have blocked
theorizing about consciousness by claiming that we do not need a theory
about it for experiencing it directly. But of course we also directly experi-
ence pain and pleasure, which is no excuse for denying that we need
theories about them. We need theories in addition to descriptions because
we want explanations in addition to descriptions and in place of mys-
teries. There is also a practical rationale for our wishing to have theories,
namely that only scientific theories, in conjunction with data, can yield
the predictions needed in applied psychology to prevent or treat behav-
ioral and mental disorders. However, the practical applications of psy-
chology deserve a separate chapter.
CHAPTER 12

Psychotechnology

Psychologists have always given great importance to the application of


knowledge acquired through basic research. Psychology has always been
a profession as well as a science, with work using applied psychology,
psychotechnology, applied behavioral analysis, and other approaches.
Almost all scientific disciplines have separated their pure and applied
functions. Thus we have physics and engineering, biology and medicine,
and so forth. In psychology the same person is often a researcher and a
practitioner at the same time, as if someone were a physicist and engineer
simultaneously. In the early part of this century, Wundt's pupil Titchener
insisted on separating psychology, which was a basic laboratory science,
from psychotechnology, which was the application of knowledge so
gained to education, clinical practice, society, justice, industry, and the
like.
Over time, many psychologists have come to reject Titchener's posi-
tion, insisting on practicing as both scientists and professionals. Even the
models for training psychologists take this position. The American Psy-
chological Association has stipulated that training must include scientific
and professional experience, so that a PhD from an accredited program
will have received training in both. A PhD in clinical psychology is thus
expected to know some mathematics and computer science, along with
physiology, experimental psychology, research methods, perception,
learning, social psychology, and also to have training in diagnosis and
therapeutic methods, and to serve an entire year of internship. This was
required to make the psychologist both scientist and clinician, and to
make psychology both research and practice.
In 1920 the International Association of Applied Psychology was
founded by E. Claparede (1875-1940) in Switzerland, in spite of the resis-
tance of Wundt, Titchener, and other pioneers. This led to the legalization
of applied psychological practice, which had the support of Watson and,
later on, Skinner and the other radical behaviorists.
In the beginning, the applications of psychology were limited to rela-
tively simple matters such as the use of tests in education and the selec-
252 12. Psychotechnology

tion of personnel. If the school psychologist indicated that the student had
more aptitude for a certain profession, it was necessary to follow up this
prediction and correlate it with the student's degree of success. Applica-
tions were also found in industry (e.g., assessing a worker's suitability for
a given job), clinical practice, and in rehabilitating delinquents.
Later on, the scope of applications of psychology was greatly enlarged,
especially with Skinner's (1938, 1953, 1971) approach to the experimental
analysis of behavior and applied behavioral analysis. Applications were
undertaken for everything from the treatment of mental retardation to the
design of cultures. This new applied psychology, or psychotechnology,
led to many more opportunities for psychologists, and the field became
one of great possibilities and ambitions.
There have been many criticisms of applied psychology and psycho-
technology. Some question the ethics of manipulating other humans;
others question the morality of the unclear goals sometimes involved.
Something that has never been questioned, however, is the effectiveness
of the techniques involved. We will consider the modern forms of the
main areas of psychotechnology.

12.1 Clinical Psychology and Psychiatry


Abnormal behavior has traditionally been a "no-man's land" or, even
worse, public property. Witches, charlatans, and psychoanalysts have
struck it rich in this undeniably important area. Anthropologists, sociolo-
gists, and clinical psychologists have demonstrated the enormous difficul-
ties involved in labeling a behavior as normal or not. Without a clear
notion of normality, there can be no well-founded clinical psychology or
scientific psychiatry and psychopathology.
To understand normality, we need first recognize that the concept of
mental health is broader than just "absence of disease." The psychology
of mental health emphasizes human growth, relations with physical and
social surroundings, and programs of primary prevention, which seem to
be more important than simple therapy. It is possible that in the future the
psychology of health, with its broad social and community orientation
(see Matarazzo, Miller, Weiss, & Herd, 1984; Stone, Adler, & Cohen,
1978), will replace much of what is known today as psychiatry and clinical
psychology.
In the area of abnormal behavior there are many obscure theories and
lots of contradictory facts. Biochemists and psychoanalysts cannot agree
about the nature of schizophrenia. Existential and behavioral psycholo-
gists use different methods and techniques for the treatment of depres-
sion. For some psychologists, alcoholism and drug abuse are diseases but
for others they are deviant behaviors. The behavioral techniques for treat-
ing insomnia are very different from the biochemical ones. Some psychol-
ogists consider adolescence a period of inevitable difficulties resulting
12.1. Clinical Psychology and Psychiatry 253

from the presence of sexual hormones in the blood, whereas for others it
is only a change in social roles that should not cause any trauma.
Notions of good and bad, normal and abnormal, well adapted or mal-
adapted change with time and culture. Efforts such as the American Psy-
chiatric Association's (1980) Diagnostic and Statistical Manual of Mental
Disorders (DSM III) are laudable because they require the imposition of
some order. In this volume, mental diseases are defined and classified,
providing a standard. However, there were DSM II and DSM I before it,
and DSM IV can be expected in the future. The changes in these editions
show just how relative the procedures for evaluation and diagnosis are,
and how fragile are the concepts of normality and abnormality on which
they are based.
Three criteria are tacitly or implicitly accepted by psychologists in
deciding what is normal:
Statistical criterion: "Normal" refers to what most people do nowa-
days, in a given culture and social group. It was normal to be a homosex-
ual in ancient Greece but not in the Middle Ages. It is normal to drink
alcohol at social gatherings but not when one is alone. This criterion is
widely used, both in tests of great validity and reliability (such as the
MMPI) and those of dubious or no validity (such as the Rorschach and
other projective tests). It is clear that construing normal behavior as what
most people do poses many problems, because this implies that highly
creative people (e.g., Da Vinci, Van Gogh, Einstein) are abnormal.
Axiological criterion: "Normal" refers to behavior in accordance with
the ideals of a given culture in which it occurs. In general, these ideal
patterns of behavior are not distributed evenly in the population, but this
is still an alternative criterion. In Western society, this would involve
being success-oriented, competetive, motivated to reach one's goals; val-
uing family without letting it interfere with personal success; having a
narrow area of expertise; having stable relations with one's mate, and so
forth. Again, this criterion presents many difficulties, in particular the
relativity of the values of different groups and the impossibility of objec-
tive comparisons across groups.
Clinical criterion: "Normal" people are those who feel well with them-
selves and with their social group. It is obvious that nobody feels well
with everyone and it is oflittle importance if this is the case. No one feels
well all the time, but the idea is that if one usually feels well with oneself
and with the important people in one's environment, one is clinically
normal. Obviously an artist worried about some aspect of her work or a
scientist puzzled by a particular problem will not feel well with herself.
On the other hand, a drug addict, an alcoholic, or someone mentally
handicapped may feel quite well with his situation and would then be
clinically normal, given this criterion.
These criteria seem to cover what has been said about normality and
abnormality. Social philosophies, psychiatry texts, and counterculture
254 12. Psychotechnology

movements are all covered. Einstein is normal by the axiological criterion


and perhaps by the clinical criterion, though not by the statistical. The
bank worker who keeps on with his work, loves his wife but goes out with
another woman once in a while, and gets drunk once a week is normal by
the statistical criterion. The institutionalized psychotic, with whom it
makes no sense to speak of "happiness" even though she feels satisfied
with her life because she knows no other alternative, is normal from the
clinical point of view. These examples illustrate the difficulty of the prob-
lem, how contradictory the criteria are, and the ambiguity of the concept
of mental health.
When specialists have taken the trouble to discuss what a psychologi-
cally healthy person is, they have defined such person by exclusion.
Subjects who do not suffer from any of the diseases listed in the DSM III
and who live according to the rules of their culture without too much
stress would be normal. Some hold that a normal person is one who
develops his or her full potential, which again is as ambiguous as it is
limited by the particular culture and difficult to explain.
Psychoanalysts attempted to provide a notion of normality by saying
that we are all abnormal, which is an outright contradiction (we all must
be normal, adjusted to a normal curve, and people with problems should
be the exception). Behavioral genetics proposed another solution: Mental
disease is linked to recessive and not dominant genes. Thus, persons
without genes for mental disease are, by definition, normal. In sum, there
have been many proposed solutions to this problem but none have had
much success.
Clinical psychology deals with problems of the diagnosis, treatment,
prevention, and research related to behaviors that deviate from some
norm. First we must agree on the nature ofthe norm, keeping in mind that
it is relative and culture specific. Then we must decide how to diagnose
normality and abnormality-with projective tests or statistical or behav-
ioral evaluation (see Ardila, 1985). This in turn raises the question of
which therapeutic system to use.
The practice of therapy has probably engendered more controversy
than any other set of questions in psychology. There have been five major
revolutions in psychological therapy:
1. Pinel's liberation of the mentally ill from incarceration
2. Freud and his notion of the "unconscious"
3. Shock therapy
4. Psychopharmacology
5. Behavior therapy
Each is associated with a different time in history and a different world-
view. Not all of these approaches can be correct. We cannot see any use
for psychoanalysis in the treatment of mental illness; its many shortcom-
ings have been discussed in detail elsewhere (e.g., Ardila, 1980a; Bunge,
12.2. Educational Psychology 255

1967a). We are cautious about shock therapy, but hopeful about the use-
fulness of (4) and (5) in treating deviation and unhappiness.
There are many systems of psychological therapy. Some years ago it
was estimated that there were 20 such systems, which seemed excessive.
Today the estimate lies closer to 200, which indicates more confusion
than healthy diversity. We have therapeutic systems based on psychoa-
nalysis, behavioral psychology, humanistic psychology, psychopharma-
cology, biological psychology, and on any mixture of them. There are too
many of them, each with precious little evidence to support it. Some are
new, others old; some from Eastern religions, others from Western labo-
ratory science. For some the goal is adaptation ofthe patient to his or her
environment; for others it is the adaptation of the environment to the
patient. Some promise mental health, others inner peace or personal ful-
fillment.
In this melange, the techniques used also vary widely. They range from
several years of verbal therapy on a psychoanalyst's couch to a few
seconds of shock therapy. In between are therapies in which the therapist
says nothing or only "Aha," and those where the patient is given thera-
peutic homework (behavior therapy). Some techniques come from Zen
and Japanese philosophies, and some even involve keeping the patient in
a bath for several days without seeing anyone. Another technique is sen-
sory deprivation, which grew out of the interesting work done by Hebb
and his colleagues at McGill University. In some therapies the therapist is
active; in others passive. "Acting out" situations is the preferred tech-
nique in some; in others it is expressly prohibited. It is clearly incorrect to
say that all therapeutic systems derive from the work of Freud, Rogers, or
Skinner. The only thing that can be said for sure is that they do come from
extremely varied philosophies or worldviews.
In sum, clinical psychology and psychiatry are disciplines that are re-
markable for their internal contradictions and heterogeneity. Not all of
the systems are compatible, nor do they have the same goals. The grand
synthesis some are hoping for in psychology is clearly still far away.

12.2 Educational Psychology


The second most important branch of psychotechnology, after clinical
psychology, is educational psychology, although it is probably older. It
includes such broad problems as the adaptation of students to the educa-
tional system, the adaptation of the educational system to students, and
the fit between the educational system and society. Education goes from
cradle to grave including along the way preschool, primary and high
school, and for some, college and post-university experiences, and it
includes normal as well as special populations such as the blind, deaf, and
mentally retarded.
256 12. Psychotechnology

One of the basic problems in education is its conservative character.


Education is in the business of transmitting information, in particular the
gains of civilization, from one generation to the next. Consequently, this
must be done in the way most faithful to the subject matter to be transmit-
ted, with no changes and especially no additions or embellishments. This
conservative nature of the educational process is the root of its many
problems.
The philosophy of education is closely related to one's context of
action, so it is very close to social philosophy, as well. In a competetive
culture education helps people obtain the characteristics and skills they
need to survive and succeed. In a cooperative culture it helps people to
cooperate. In some cases knowledge is emphasized, in others awareness
of reality including society and its foundations. Education is closely
linked to both sociology and philosophy, so it cannot avoid either.
Psychologists are very interested in education, and the interest has
been mutual. Psychological knowledge has been used at all levels and in
all areas of education, but especially with more difficult popUlations: the
handicapped, preschoolers, and the learning disabled. Applied behavioral
analysis has been particularly useful with special populations (see Adam-
son & Adamson, 1979, for learning disabilities).
Psychological technology in education, or educational technology, has
a very long history. First it was based on Binet and Simon's tests for
selecting children who should continue their studies. Later on the field
was refined, greatly enlarged, and was applied to work in addition to
testing. Psychologists helped teachers to perform more valid evaluations,
to deal with learning difficulties, and to develop programs and materials.
Testing has many limitations and even more critics. Right now it does
not enjoy great prestige, despite its solid mathematical basis, its use of
scientific methods, and its enormously widespread application. In the first
place, tests measure constructs that are often very difficult to operationa-
lize, such as intelligence and personality. When they refer to more clear-
cut matters such as achievement, attitudes, and aptitudes, there is no
problem. But difficulties arise when tests concentrate on broader psycho-
logical constructs. Because of their ignorance of philosophy, most mod-
em psychologists have not been able to solve these problems. If we are
measuring intelligence, first we have to define it, to know what it is we are
attempting to measure. Theories of intelligence are very numerous and
have not yielded solutions to the basic problems-the solutions on which
to base any consistent, systematic knowledge. As for personality testing,
the situation is even worse. Workers on measurement and evaluation
have often tried to circumvent the problems by saying that intelligence is
what intelligence tests measure, as Guilford said of his own creativity
tests.
In the context of measurement and educational psychotechnology it is
important to remember the case of Cyril Burt (1883-1971). The most
12.3. Industrial and Organizational Psychology 257

important psychologist in Great Britain for several decades, he focused


his work on child development and on statistics. His work on intelligence,
factor analysis, child guidance, mental retardation, and delinquency
gained him worldwide respect. But after his death it was discovered that
Burt had falsified his data to support his theories of race and social class.
Fraud is the worst sin in science and Burt commited it to further his point
of view. Most of the subjects in his studies of twins had never existed and
his statistical correlations were wrong. Nobody knows how many things
Burt falsified, but he certainly did so to prove his theories of the superior-
ity of the Anglo-Saxon race and British culture.
Burt emphasized the genetic contributions to intelligence, rather than
the role of environment. For him, a high percetage of the variance in
intelligence is due to inheritance; intelligence is inherited just as eye and
hair color is. Burt's conclusions were widely accepted and disseminated
in textbooks in many languages. Because it was later revealed that Burt
was just someone with strong prejudices and no respect for science, we
do not know to what extent intelligence is inherited, how much early
stimulation can increase someone's intellectual potential, or whether
there is some interaction of genetic and environmental factors. In sum,
Burt's studies of intelligence, its measurement, and its genetic origins
were discarded after discovering that Professor Burt was a charlatan.
This rather sad and well-known case shows how human weakness plays
an important role even with famous scientists. However, it also shows
that science is self-correcting.

12.3 Industrial and Organizational Psychology


Industrial applications of psychotechnologies are much more recent than
educational or clinical ones. In spite of some additional difficulties they
face, these applications have reached a high level of development.
In general, industrial psychology involves the selection of employees
for an organization: bosses, secretaries, workers, artists, athletes,
models, and so forth. Initially, the techniques were limited to unstruc-
tured interviews and standardized tests, but today other methods are
used, such as role-playing, and training in specific areas. Psychologists
have been quite successful in this area.
One of the most interesting questions industrial psychology deals with
is worker motivation: raising their morale and increasing their involve-
ment with the enterprise. All administrators are interested in motivating
employees to feel vitally involved with the goals of the enterprise, to work
harder for it, and to defend it as if it were their own.
Do people work for money? Curiously, the answer seems to be nega-
tive. In many cases people do not work for money but for other rewards,
for example, affiliation, recognition, or personal satisfaction. An artist
258 12. Psychotechnology

paints pictures not for money but for prestige. A scientist may spend a
lifetime researching a complex problem in physics just for the sake of
solving it, as if it were a riddle, and not for money. The motivations of
scientists and artists are very complex (see Mahoney, 1976, for discussion
of scientists). Affirming that people work only for money is clearly false,
but it would be just as false to say that money is not a motivating factor.
Industrial and organizational psychology have shown that money is not
the central motivating factor in work, although there are large cultural
differences in the importance ascribed to having and earning money.
Besides the motivations for working, psycho technology has investi-
gated the social and physical characteristics of the workplace, for exam-
ple, with respect to lighting, presentation of material, noise, and the like.
It is no longer devoted to the time-and-motion studies like those done by
Taylor, but investigates the effects on efficiency and productivity of noise
level, ambient colors, and amount of time between breaks.
The social dimensions of work have received much attention in the last
few decades, and a substantial body of work has appeared (see Varela,
1971, 1977). Without a doubt the findings on human communication, in-
formation networks, and the creation of small groups apply to work and to
the creation of psychotechnologies for the workplace.

12.4 Designing Cultures


One of the most ambitious goals of psychotechnology is designing whole
human environments: institutions, communities, cities, countries, even
the whole planet. These ambitions have much to do with utopian litera-
ture such as Huxley's Brave New World (1932) and Orwell's 1984 (1949).
More particularly, psychological designs have been inspired especially by
Skinner's Walden Two (1948), which has been taken very seriously and
has inspired many communes in Europe and the Americas.
In attempting to design a culture, one tacitly assumes that human con-
duct is subject to laws-an assumption of any scientific psychology-and
is controllable. It is also assumed that if we design the social and physical
environment in the appropriate way, we will obtain the behavioral results
we desire: that it is possible to make people happy, cooperative, produc-
tive, efficient, and nonagressive; that children can learn to read and write,
add and subtract without any effort, and later on acquire the knowledge
they need to be creative, to produce new and valuable ideas; that the
social environment can be arranged so that crime, indigence, and delin-
quency will not occur; that work can be a pleasure rather than a chore.
This program is tremendously, even excessively, ambitious. In fact, it
is one of the dreams of the forgers of utopias of all ages, including reli-
gious communities and followers of Lenin and Trotsky. They are all uto-
12.4. Designing Cultures 259

pian programs for changing people and society to reach traditional ideals
that have not been attainable otherwise.
The difference between the psychological method of designing cultures
(ft la Skinner), and the others is that it is based on the precepts of science,
in particular experimental behavioral analysis: Species and individual dif-
ferences yield their importance to the laws of animal and human learning;
reinforcement and its scheduling are the basic principles, and the rate of
responding the basic unit of measure.
There are differences among psychological utopias; for example, the
problems faced in Walden Two are different from those that arise in
Walden Tres (Ardila, 1979b) because in the latter we are working with a
whole country, so political, historical, and socioeconomic factors take on
more importance. The goal may be the same (i.e., designing a culture
based on the principles of operant psychology, showing how to proceed,
and some of the obstacles that will appear), but when we go from a few
hundred people in Skinner's Walden Two to several million in Walden
Tres the situation changes drastically and the problems are of a very
different magnitude. When Thoreau wrote Walden (1854), he probably
never thought it would inspire a Walden Two and later a Walden Tres
more than a hundred years later.
The design of cultures is the most ambitious goal of psychotechnology.
It is not limited to changing the processes associated with mental health,
education, work and productivity, but seeks to change the entire society.
It insists on modifying people, sometimes tremendously, without regard
for their innate capacities or limitations. Though this may not seem very
realistic, psychological utopias have been taken very seriously. Several
communes have been organized on the principles of Skinner's psychol-
ogy, including Twin Oaks in Virginia and Los Horcones in Mexico. All of
these communes used Walden Two as a guide to building a perfect soci-
ety; they exist and have been relatively successful. The goal in Walden
Two is the same as that for other utopian projects, but this perfect society
is based on psychology and its concrete, realistic principles. In Beyond
Freedom and Dignity, Skinner (1971) wrote:

The application of the physical and biological sciences alone will not solve our
problems because the solutions lie in another field. Better contraceptives will
control population only if people use them. New weapons may offset new de-
fenses and vice versa, but a nuclear holocaust can be prevented only if the condi-
tions under which nations make war can be changed. New methods of agriculture
and medicine will not help if they are not practiced, and housing is a matter not
only of buildings and cities but of how people live. Overcrowding can be corrected
only by inducing people not to crowd, and the environment will continue to
deteriorate until polluting practices are abandoned. . . . in short, we need to
make vast changes in human behavior. . . . what we need is a technology of
behavior." (pp. 2-3)
260 12. Psychotechnology

12.5 The Goals of Psychotechnology


One problem with the enormous philosophical and political implications
of psychotechnology is to define the goals of change. What should we
change, why, and in what direction? Which are the implicit and explicit
values we are using when we say that it is necessary to make education
more efficient, or when we use a behavior modification program to make
workers more productive, or when we use psychological methods to
change a criminal's behavior?
In other words, what or whom does psychology serve? The mainte-
nance of the status quo? Advances toward the humanitarian ideals of our
civilization? The best interests of the individual? Psychology, being a
science, serves no interest but that of the quest of knowledge for its own
sake. Basic science is different in this respect from applied science and
technology (Bunge, 1983b), and psychotechnology serves a series of val-
ues and implicit assumptions.
Psychotechnology can be criticized for often being naive and simplistic.
It seeks solutions for concrete, objective problems of the here and now,
much as any other kind of engineering. Varela's social technology, which
has received much attention from psychologists, is an attempt to integrate
the findings of different areas of psychology, especially social psychol-
ogy, and apply them to the solution of practical problems, particularly in
organizational settings. It is one example of how basic psychology can be
helpful in solving practical difficulties and thus give rise to a psychotech-
nology. The harsh reality, however, is that many psychological facts and
theories are mutually incompatible, are rooted in different philosophical
positions, and have different goals. The time for finding points of conver-
gence between them has not yet arrived. Social technologists insist that
their discipline is more similar to engineering than to physics, in the sense
that it seeks to solve practical problems rather than to find scientific
explanations that are coherent and internally consistent. Practical prob-
lems are urgent, and we cannot wait until laboratory research solves all
their theoretical and methodological underpinnings to begin to work on
them. If we wait for that, we will never do any applied science or technol-
ogy. So, workers in every technology need to improvise (see Walden
Tres) , because practical problems are usually too important and urgent. In
fact the goals of psychotechnology are not the same as those of science.
The gains of one cannot be evaluated using the criteria of the other.
Psychotechnology has many limitations and is open to many criticisms.
This can be seen, for example, in the enormous number of therapeutic
systems, most of which are worthless. Their proponents cannot even
agree on what mental health, normality or abnormality are, about the
effectiveness of a given treatment, or even the parameters that should be
measured. This is similar to the situation in physics a few centuries ago,
12.6. Summing Up 261

with the difference that in pre-Newtonian physics there were not quite so
many conflicting points of view.
In spite of its problems, psychotechnology has grown rapidly and made
much progress since the founding of the International Association of Ap-
plied Psychology in 1920. Its fields of application have been broadened to
encompass ecological and economic problems, publicity and advertising,
and the stimulation of creativity. Fifty years ago no one would have
thought that psychology was going to design prosthetic surroundings for
old people or select and train astronauts to go to the moon. The relations
between cancer and mental health were not known, nor was it anticipated
that war and peace would become a subject for the application of psychol-
ogy (see Ardila, 1986, for the psychological impact of nuclear war). One
field to see enormous growth was that of the relations between psychol-
ogy and society.
The background problems of psychotechnology are still very compli-
cated and difficult. They are not psychological but political, structural,
philosophical, and conceptual problems. One tries to understand the di-
rections in which individuals and societies want to develop. In this en-
deavor, psychology must make use of the other disciplines, especially
philosophy-its birthplace-but also physics and the other sciences that
psychologists have tried to ignore.
Recently the "hard" sciences such as physics and chemistry have
shown increased interest in their philosophical foundations, and the
"soft" sciences such as economics and psychology have become more
interested in their historical roots. It is as important for the hard sciences
to find their philosophical underpinnings as it is for the soft sciences to
find their historical roots. Surely in the near future both groups will under-
stand that both history and philosophy are fundamental. When this hap-
pens, psychologists will recognize that a philosophical analysis of psy-
chology and psychotechnology has much to offer to the understanding of
humans and their behavior.
Moreover, it is not enough just to understand how humans act. As
Marx said, philosophy has beed dedicated for a long time to understand-
ing the world; to change it is the important thing now. This has always
been the goal of psychotechnology.

12.6 Summing Up
Psychology differs from other behavioral sciences in the great interest it
has shown in its applications. It holds that science and its applications go
hand in hand. Whereas in most other disciplines the person who produces
basic knowledge and the person who applies it are different, in psychol-
ogy the same person often does both. Psychology has insisted on being
both a science and a profession.
262 12. Psychotechnology

Applied psychology, or psychotechnology, has diverse roots, and it


extends into numerous fields, among them the treatment of mental dis-
eases, the selection of students using Binet's tests, applications in indus-
try, applied behavioral analysis, and the design of cultures. These are
fields that are growing rapidly and becoming ever more diversified. The
large majority of these applied fields are internally contradictory, using
different assumptions. The integration of their findings seems very far off.
There is a great gap between basic and applied research in psychology.
Perhaps not even 10% of laboratory findings have been applied. Maybe
many of these findings never will be applied at all, as is common in other
sciences. However, in order to have a valid science, it is not necessary to
have a valid body of applications.
VI

Conclusion
CHAPTER 13

Concluding Remarks

The reader who has survived this far may feel somewhat puzzled. For
instance, he or she may wonder how to reconcile reduction (of the mental
to the neural) with emergence (of mental functions out of nonmental
ones); or how reductionism could possibly promote the integration of the
various branches of psychology, which are so far largely dismembered; or
why there is the insistence that current psychology is poor in theories,
hence also in explanations, and that mature science does not include
metaphors except as heuristic props.
In this final chapter we shall attempt to solve some of these puzzles. We
shall also propose a diagnosis of current psychology and shall venture an
optimistic prognosis provided certain current trends are reinforced
whereas others are weakened. Finally, we shall summarize some of the
philosophical implications of current psychological research and shall at-
tack once again the divorce between philosophy and science, which paral-
lels and helps to keep that between psychology and biology.

13.1 Reduction
Throughout this book we have espoused and exemplified the psycho-
neural identity hypothesis, that all mental processes are neural processes
of a special kind (section 1.3). This is a reductionist thesis, for it identifies
two classes of facts that, from alternative viewpoints, are mutually dis-
joint. The thesis is in the same boat with the theses that light is electro-
magnetic radiation, and human history the evolution of human societies.
All these theses exemplify ontological reductionism-ontological in that
they concern things, properties, or processes rather than our knowledge
of them.
The methodological status of any such identity thesis depends upon the
stage in the historical evolution of the branch of knowledge in which the
thesis occurs. In fact, identity theses usually begin as hypotheses (corrigi-
ble assumptions). But, if confirmed and embedded in well-corroborated
theories, they end up as definitions (conventions in the form of identities).
266 13. Concluding Remarks

Thus, in modern physics light is defined as electromagnetic radiation of


wavelengths that are between 3.800 and 7.600 A. As a consequence,
optics, formerly an independent science, has become a chapter of electro-
magnetism. Still, most optical processes can be described, though not
explained, in purely optical terms-that is, by taking the concept of light
as basic (primitive) rather than derived (defined). But these are epistemo-
logical matters.
Should the psychoneural identity hypothesis become generally ac-
cepted and incorporated into a solid general theory of mind, it would
occur as a definition in the latter. A possible candidate for such post is the
identity: "Mental process = Specific process occurring in a multi-
neuronal plastic system" (Bunge, 1980). This is an instance of a reductive
definition. Further reductive definitions in line with the preceding ones
are: "Learning is the strengthening of synaptic connections in a group of
neurons," "vision is the specific function of the visual system (including
the visual cortex)," and "mental disorders are either maladaptive learned
behavior patterns or brain dysfunctions."
What we have called 'reductive definitions' are often called bridge
formulas, for bridging two formerly separate theories or disciplines-in
our case psychology and neuroscience. However, the term 'bridge for-
mula' is too weak, for there can be bridging without reduction. For exam-
ple, the formula for the force that an electromagnetic field exerts on an
electrically charged particle links mechanics to electromagnetic theory
without reducing either of them to the other. Likewise a psychoneural
dualist might propose bridge formulas, such as "consciousness can move
neurons," which are far from suggesting the reduction ofthe mental to the
neurophysiological.
We shall distinguish two kinds of ontological reductionism: radical or
leveling, and moderate or emergentist. Radical reductionism denies,
whereas moderate reductionism admits, that wholes or systems may have
properties that their parts or components lack. (Recall section 3.2.) Radi-
calor leveling reductionism is the same as microreductionism: It holds
that a system can have no other properties than those of its constituents.
(This thesis leads inexorably to the chain reduction: social science to
biology to chemistry to physics.)
On the other hand, emergentist (or systemic) reductionism affirms that
systems possess (emergent) properties that their components lack. For
example, density, viscosity, transparency, and conductivity are molar
properties of a liquid body, which its component atoms or molecules do
not possess. Likewise the abilities to remember, learn, perceive, or think
are properties of systems composed of many neurons, not of single neu-
rons; like stability and temperature, they are emergent properties, not
resultant ones such as energy. Given the staggering amount of empirical
evidence for the existence of emergent properties-particularly of multi-
neuron systems-we reject radical reductionism and the attendant mi-
13.1. Reduction 267

croreductionism, and espouse emergent or moderate reductionism in-


stead.
So far we have dealt only with what may be called the structural aspect
of emergence (i.e., the occurrence of new systemic properties at anyone
time). But such properties appear at certain points in the history of things,
for example, as the result of chemical reactions or of social interactions.
In particular, the behavioral and mental abilities emerge in the course of
the development of the individual animal or in the evolution of its biopo-
pulation. This is the dynamic aspect of emergence, stressed by such di-
verse evolutionary emergentists as C. Darwin, F. Engels, H. Bergson, C.
Lloyd Morgan, A. N. Whitehead, and R. W. Sellars.
Our view of behavior and mind covers both the structural and the
dynamic aspects of emergence. It states not only that all complex behav-
ior is controlled by multineuron systems and that all mental events are
changes in the couplings that hold such systems together; it also affirms
that behavior and mind are outcomes of evolution (in the case of popula-
tions) or of development (in the case of individuals). On the other hand,
psychoneural dualism, if consistent, has no serious use for the evolution
and development of behavior and mind. In particular, when deriving from
theology it is bound to deny the relevance of evolution and development,
for theology holds that the ready-made soul or spirit is infused into the
fetus by the deity (e.g., Eccles, 1980, p. 240).
The preceding considerations on ontological reduction and emergence
are not idle metaphysical pastimes: They are relevant to the strategy of
scientific research. Indeed, if mind is not a biological property, then it
cannot be studied with the help of biological ideas and procedures. And if
mind is biological but not emergent, then we should be able to discover it
in the single neuron-failing which we might conclude that it is not biolog-
ical after all. This again is not idle speculation. In an epoch-making book
Eccles (1953) sketched the program for the development of neurosci-
ence over the next three decades, by stating that the understanding
of the nervous system would come from a thorough understanding of
the individual neuron. This microreductionist strategy paid off hand-
somely-but only up to a certain point, because it failed to disclose
the secrets of behavior and mind. Eccles himself fell into the trap: Not
finding the mind in the single neuron, he concluded that it must be
nonmaterial.
We must then resist extreme reductionism with regard to knowledge
because, although it is likely to further research for a while, it is bound to
mislead us in the face of emergence. If there are qualitative nova de re
then there ought to be nova de dicto; that is, emergent things, properties,
and processes must be represented by new concepts, hypotheses, or theo-
ries. If we acknowledge this maxim we can understand why every science
is at the same time very special and closely attached to other sciences:
why it is specific yet not independent.
268 13. Concluding Remarks

For instance, geology, though based on physics and chemistry, contrib-


utes its own particular concepts (e.g., that of crustal plate), theories (e.g.,
that of plate tectonics), and methods (e.g., seismological recording). But
at the same time geology explains its facts in physical and chemical terms.
Thus geology combines epistemic emergence (the formation of new ideas)
with epistemic reduction (via definitions and explanations). This peculiar
combination of emergence and reduction in the domain of knowledge may
be called moderate epistemic reductionism.
Let us apply the preceding ideas to the psychology-biology relation.
There are essentially two views concerning this relation: autonomism and
reductionism. Methodological autonomism is the classical view that psy-
chology is an independent science that owes nothing to any of the other
sciences. This view has been defended not only by mentalists such as
Maine de Biran (1823-1824) and Freud (1929) but also by behaviorists
such as Suppes (1975). The latter's reason for holding that psychology is
as fundamental a science as physics is that "the most important psycho-
logical theories are to a large degree independent of physiology and
biology" (p. 270). Supposing this to be true, it would only suggest
that so far psychology has not been anchored to biology-but the
future need not copy the past. However, the contention is not true:
There are a few theories, even mathematical ones, in physiological
psychology. The mere existence of the latter refutes psychological
autonomism.
As for the anti-autonomist thesis, it comes in two strengths. The strong
reductionist thesis, or biologism in the field of psychology, states that
psychology must eventually become as much a branch of biology as ge-
netics or cellular biology: that it will be in no need of specific concepts,
hypotheses, or methods. Obviously, this thesis does not hold for present-
day psychology, which handles a large number of concepts, such as those
of dreaming and thinking, that are alien to biology, strictly speaking, even
though they are in the process of being biologized. So, radical epistemic
reductionism is still only a program. The question is whether this program
can be implemented. Let us see.
Of course, the thrust of biopsychology is the reduction of all psycholog-
ical concepts to neurobiological ones. However, this does not entail that
all psychological hypotheses are likely to go the same way. The reason for
holding that some of them will not be reduced to biological propositions is
that much of psychology happens to study social animals. And biology
does not supply such sociological concepts as those of culture, crowding,
and antisocial behavior, which are needed to account for some aspects of
behavior and mind. In this regard psychology is quite different from ge-
netics or cellular biology, which are at the very center of biology. One
may hope only that psychology will move toward the periphery of biology
and, more particularly, toward the intersection of biology and social sci-
ence. Let us illustrate this point with an example.
13.1. Reduction 269

Consider this proposition: "Extreme crowding increases the level of


adrenocortical hormones, which augments stress, which in turn favors
antisocial behavior." This generalization, well known to social psycholo-
gists mindful of physiology, contains the concept of antisocial behavior.
This concept is irreducible to biology-and, moreover, it is culture-
bound. Of course, it can be hoped that the mechanism whereby crowding
alters social behavior will be unveiled with the help of neuroendocrino-
logy. But, because it involves deviant social behavior, its study calls for
the cooperation of social scientists. In this regard the psychologist is in
the same position as the family physician; the latter too must take the
social circumstances of his or her patients into account. In short, even
though behavior and mentation are biological phenomena, they are often
socially conditioned, whence they cannot be studied by biology alone.
For this reason biological reductionism can only go so far in psychology.
It is more advisable to adopt biosociological reductionism instead.
Favoring the inclusion of psychology in the intersection of biology (in
particular neuroscience) and social science (in particular sociology) does
not amount to fostering the disappearance of psychology, but only the end
of its pretense of independence. The full inclusion of psychology in the
intersection of biology and social science can only strenghten it. There are
instructive historical precedents. For instance, astronomy and meteorol-
ogy were autonomous disciplines for centuries and eventually reached a
point where they could not advance any farther because they were di-
vorced from physics, which alone could supply the mechanisms underly-
ing the astronomical and meteorological phenomena. The two sciences
expanded tremendously upon becoming chapters of physics: astronomy
in the seventeenth century and meteorology in the nineteenth.
Nor does favoring a reductionist research strategy entail recommending
microreductionism. The analysis of a system into components, though
necessary, is never sufficient to understand a system. One also needs to
look into the environment as well as into the interactions between the
system components and between these and the environmental items. (Re-
call section 3.2, in particular Formula [3.3]). In other words, when study-
ing a system we should employ two mutually complementary strategies:
same-level (or horizontal) and inter-level (or vertical) study.
A same-level study of a system regards it as a whole and attempts to
discover its global or molar behavior. An inter-level study, on the other
hand, inquires into the relations among items on various levels of organi-
zation (e.g., cellular, organismic. and social). In turn, an inter-level study
may be either of two kinds: top-down or bottom-up. A top-down study of
a system analyzes it into its components at one or more levels. On the
other hand, a bottom-up study attempts to reconstitute the whole from its
parts and their interactions. No doubt, all three strategies are used in
contemporary science, and particularly, in psychology. For example, the
studies of behavior and cognition in the prebiological tradition are same-
270 13. Concluding Remarks

level studies; in particular, most investigations in psychophysics and in


cognitive psychology are of the same-level type. However, an analysis of
the visual system (or any other subsystem of the nervous system) into
components and processes is of the top-down type, and the attempts to
explain memory, learning, and creativity in terms of changes in synaptic
connections are of the bottom-up kind.
In sum, with regard to the problems of behavior and mind we favor a
combination of moderate ontological reductionism with moderate episte-
mological reductionism. (In other matters-for example, the relations
between living matter and its nonliving components, or between society
and its members-we favor a combination of ontological emergentism
with moderate epistemological reductionism: See Bunge, 1977b, 1985a.)
That combination accounts for emergence without mystery, as well as for
reduction without the exaggerations of microreduction. It also facilitates
the much needed integration of the various disciplines dealing with behav-
ior and mind (of which more in the next section).
Our thesis about the reducibility of psychology does not entail that all
psychologists should stop plying the special tools of their trade to become
neurophysiologists or sociologists. It only entails that they should stop
thinking in nonbiological and nonsociological terms-for example, re-
garding the mental as disembodied, and wishing to account for brain
function and social behavior in terms of symbols rather than the other
way round.

13.2 Integration
Things cannot always be explained by analysis or reduction only; quite
often they can only be explained by placing them in a wider context. In
turn, the consideration of such a wider context may require bringing
together or consolidating results obtained in two or more fields of re-
search. More often than not, a multidisciplinary study will achieve the
desired goal, but occasionally a more intimate relationship proves neces-
sary, and the merger of theories or even disciplines may result. Table 13.1
lists some revolutionary mergers. Some of them have made it possible to
study properties, events, and processes on a given level in terms of lower-
level laws.
The integration or synthesis of approaches, data, hypotheses, theories,
methods, and sometimes even entire fields of research is needed for sev-
eral reasons: First, because there are no perfectly isolated things except
for the universe as a whole; second, because every property is lawfully
related to some other properties; and third, because every thing is a
system or a component of one or more systems. Thus, just as the variety
of reality and the limitations of the human intellect require a multitude of
disciplines, so the integration and advancement of the latter are necessi-
13.2. Integration 271

TABLE 13.1 Examples of mergers of previously separate research fields.


Original fields Merger
Logic, mathematics Mathematical logic
Algebra, geometry Analytic geometry
Mechanics, gravitation theory Celestial mechanics
Mechanics, thermodynamics Statistical mechanics
Physics, chemistry Physical chemistry
Classical genetics, biochemistry Molecular biology
Darwinian theory of evolution, classical genetics Synthetic theory of evolution
Synthetic theory of evolution, ecology Evolutionary population ecology
Psychology, neuroscience Physiological psychology
Economics, sociology Economic sociology
Sociology, history Social history

tated by the unity and complexity of reality. Just as reduction promotes


depth, so integration precludes narrowness.
We may say that a theory or field of research T is a merger of the
theories or fields of research T, and T2 if, and only if, the following
conditions are satisfied. First, T, and T2 share some referents as well as
some concepts denoting such common referents. (For example, psychol-
ogy and neuroscience are about animals, and they share, among others,
the concepts of animal and of excitation.) Second, there is a set G of glue
or bridge formulas relating some concepts of T, to concepts of T2• (For
example, "Habituation results from inhibition. ") And third, the bridge or
glue formulas in G are sufficiently well confirmed.
The first condition excludes theories or fields that are totally alien to
one another, such as astronomy and personology-astrology notwith-
standing. The second condition draws attention to the links needed to
form a conceptual system out of two formerly separate theories or fields
of research. And the third condition is added because, in principle, there
are infinitely many bridge or glue formulas. Only those matching the
available empirical evidence will effectively hold together the original
theories or fields. (However, at the programmatic stage no such evidence
may be available, and the Gs function as hypotheses prodding research.)
Let us see how all this applies to psychology. Far from ignoring the
genuine findings of the various chapters of classical (prebiological) psy-
chology, the biopsychologist will make full use of them and will attempt to
unify them, thus overcoming the barriers that have slowed down or even
blocked research into behavior and mind. An example will suggest why
and how this must be done (Bunge, 1986). Suppose someone writes on a
blackboard a sentence with a cognitive content. Because the sentence
conveys an item of knowledge, its production and understanding fall
within the purview of cognitive psychology. But, because it is a sentence,
linguistics too, in particular psycholinguistics, is competent to study it.
Nor are these the only disciplines interested in the matter. Because writ-
ing a sentence is, among other things, a motor act, it also comes under the
272 13. Concluding Remarks

heading of the study of behavior. Moreover, the act of writing involves


visual and haptic perceptions, so that it is of interest to the psychology of
perception (in particular psychophysics) as well. But because the sen-
tence may have served to communicate something to someone, it also
falls under the sway of social psychology. Because it would not have been
written without some motivation, it is of interest to neuroendocrinology
as well. Had there been anything wrong with the spelling or the reading of
the sentence, neurologists and educators would have pounced on it. Last
but not least, thinking up the idea and writing it down involve brain
processes-particularly in Wernicke's "area," the sensory and motor
strips, and the visual cortex-so that physiological psychologists would
stake a claim on it.
This example suggests that there are no autonomous branches in scien-
tific psychology. The divisions of psychology into different fields (e.g.,
according to the various classical faculties) is utterly artificial-and so is
the currently fashionable detachment of cognitive psychology from the
other branches of the science of behavior and mind. (Recall sections 2.3.
and 9.4.) In actual fact, writing or reading a sentence is a unitary process
with a number of different though mutually connected aspects. The divi-
sion is not in the process itself but in the eyes of its beholder. The process
of forming and writing a sentence is a biosocial one and, although it is
legitimate-nay indispensable-to distinguish its various aspects and to
emphasize them one at a time, such aspects should not be detached from

~ S

A A
E N N

@
1
E I E I
I

~
FIGURE 13.1. Interdisciplinary fragmentation not honored by anatomy or physiol-
ogy. Example: Interactions among the three regulatory body systems (N = Ner-
vous, E = Endocrine, I = Immune) and between them and the rest of the body
(RB). Both the explanation of the workings of the healthy organism and the
treatment of its dysfunctions require an integration of the various disciplines.
Effective medicine is systemic (though not holistic). And systemic medicine is
based on integrated (not dismembered) medicine. In turn, the latter is incomplete
unless the organism is treated as embedded in its social matrix S.
13.2. Integration 273

one another. In general, the ticket is this: Distinguish but do not sepa-
rate-and unite but do not conflate. See Figure 13.1.
To isolate any chapter of psychology, for example, cognitive psychol-
ogy, from the rest of psychology, as well as from neuroscience, is as bad a
research strategy as to isolate the study of clouds from the rest of physics.
Meteorology became a science the day it was transformed from the study
of "meteors" into the physics of the atmosphere. Likewise, psychology
will turn into a mature science only if it is conceived of and cultivated as
the biological and sociological study of behavior and mind. Analysis, and
the accompanying division of labor, is effective only when accompanied
or followed by synthesis and the concomitant cooperation among the
relevant disciplines. (For the need to combine analysis with synthesis see
Ardila, 1987.) It is one thing to emphasize now this, now that aspect of
psychology, and another to reify the artificial boundaries between its
branches. Specialization should be tempered with integration.
It might be claimed that the recent constitution of "cognitive science,"
as the merger of cognitive psychology, linguistics, and artificial intelli-
gence, effects the desired integration. We submit that it is the wrong
synthesis, for it excludes the other branches of psychology and it ignores
both neuroscience and social science, whereas it includes a branch of
technology. (Three cousins do not constitute a family.) Cognitive science
effects also the wrong reduction, for it conceives of every bit of behavior
and mentation as a case of information processing or "computation" on
certain inputs or representations. (Recall section 5.4.)
The correct and badly needed synthesis is the merger of all the
branches of psychology on the basis of neuroscience, together with devel-
opmental and evolutionary biology, and in tandem with social science.
This is the correct synthesis because behavior and mentation happen to be
biological processes occurring in animals living in societies. For this rea-
son we placed mature psychology in the intersection of biology and social
science: Recall section 13.1.
The absorption of psychology by biology and social science does not
eliminate the former as a special or distinctive science, that is, one with its
peculiar problematics, methodics, and concepts; it only puts an end to the
alleged independence of our science. Indeed, scientific psychology, just
like its proto scientific precursor, will continue to study problems of its
own, such as those of learning and thinking. But it will study them as
neurophysiological processes, presumably consisting of synaptogenesis
and the formation of new neural systems ("rewiring") occurring under
the influence of other body systems (in particular the endocrine one) as
well as under the influence of external stimuli (in particular social ones).
In short, psychology will lose its autonomy but not its specificity. It will
cease to be the anomalous discipline to become a member of the tightly
knit system of scientific knowledge. In this regard its evolution will be
similar to that of chemistry, biology, and history.
274 13. Concluding Remarks

So far we have dealt with the reducibility of psychology to other sci-


ences. What about the converse reducibility, of other sciences to psychol-
ogy? A consistent individualist or atomist (as opposed to both a holist and
a systemist) will hold that all of the social sciences, from anthropology
and sociology to economics, political science, and history, are "ulti-
mately" reducible to psychology. This thesis is implicit in behaviorism,
which calls all social sciences 'behavioral sciences' and places them on
the same footing with psychology. But of course the thesis is much older:
It can be found, for example, in classical and neoclassical economics. In
particular, the latter holds that the basic law of human behavior, which is
supposed to explain all social facts, is that all rational beings endeavor to
maximize their subjective values or utilities. (For an eloquent defense of
this thesis see Romans, 1974.)
There are several problems with the individualist thesis. (See e.g.,
Bunge, 1979c, 1985b.) A first problem is that the restriction of the "basic
law of human behavior" to rationals renders it irrefutable (i.e., insensitive
to adverse empirical information). Indeed, whenever an individual fails to
maximize his utilities he or she is declared to be nonrational: so much the
worse for facts. A second problem is that the social sciences do not refer
to individuals but to social groups, as a consequence of which they con-
tain supraindividual categories. Suffice it to mention those of means of
production, technology, capital, balance of payments, international rela-
tions, and political system. None of these categories seems to be definable
in terms of individual psychology. On the contrary, as we saw in chapter
10, some facets of individual behavior can only be explained in terms of
the membership of every individual in various social groups. In conclu-
sion, though based on psychology, the social sciences go beyond it. Logi-
cally and methodologically the social science-psychology relation is simi-
lar to the chemistry-physics and the biology-chemistry relations. In all
three cases novel concepts, hypotheses, and methods emerge that ac-
count for the emergence of novel entities (i.e., entities characterized by
emergent properties). (For details see Bunge, 1985a, 1985b.)

13.3 Explanation
All young sciences are predominantly descriptive: They are poor in hy-
potheses and, a fortiori, in theories. For this reason their descriptions are
coarse and superficial. (Try to describe something you have observed but
about which you have not the faintest idea of what it is or what makes it
tick. The outcome is likely to resemble children's descriptions of complex
systems.) For the same reason the young sciences are seldom capable of
supplying adequate explanations and predictions of the facts they de-
scribe. For example, we still lack an adequate explanation of color vision,
and we can seldom predict the performance of a person, although there is
13.3. Explanation 275

no dearth of descriptions of vision and behavior. What we need is better


theories of vision, personality, and so forth. There is no adequate expla-
nation or prediction without adequate theory.
However, sometimes theoretical poverty is extolled as a virtue, in
obeisance to the credo of early positivism. Thus Skinner (1969, p. xi):
Hypotheses are resorted to
only because the investigator has turned his attention to inaccessible events-
some of them fictitious, others irrelevant .... Behavior is one of those subject
matters which do not call for hypothetico-deductive methods. Both behavior itself
and most of the variables of which it is the function are usually conspicuous.

It is methodologically suicidal and psychologically naive to believe


that human beings, no matter how carefully theory-aversion conditioned
they may become, will ever cease asking whys and framing conjectures
and theories to answer them. The point is not to avoid asking or answer-
ing whys but to avoid pseudoexplanations, and to propose adequate theo-
ries capable offraming correct explanations. Regrettably, the psychologi-
cal literature is replete with pseudoexplanations. Let us review them
quickly, before examining the types of genuine psychological explanation
(Bunge, 1985g).
A first rather common type of pseudoexplanation may be called tauto-
logical for consisting in accounting for mental facts in terms of mental
faculties. For example, we are sometimes told that we remember because
we are endowed with memory, or that we can talk because we are born
with the faculte de lang age. This type of pseudoexplanation is rampant in
mentalism. Obviously, it is no explanation at all. In fact, nothing is being
explained by stating that A does B because A has the ability of doing B, or
because A was born to do B, or because A knows from birth how to do B.
Moreover, a logical fallacy is involved therein, because "A" entails "A is
possible" but not conversely: Possibility does not ensure actuality. In
particular, ability or competence does not guarantee performance.
Another common type of pseudoexplanation is teleological. It consists
in positing goals or purposes whether or not there is any evidence for
them. For example, Freud affirmed that neurotic symptoms arise in order
to avoid anxiety. Invocations to purposiveness are devoid of explanatory
power; nothing follows from the assertion that A does B in order to attain
C. This is not to deny the existence of intention and purposive behavior:
See section 9.5. It is just that purpose cries for explanation. For example,
it is true that we look in order to see; but this is a fact to be explained-
perhaps by finding that the activation of the prefrontal lobes "primes" the
visual cortex, thus facilitating the functioning ofthe latter. In short, teleo-
logical accounts are sometimes permissible but they are not sufficient.
They must be supplemented with causal, probabilistic, or evolutionary
explanations. Therefore, recommending that psychology get rid of effi-
cient causes and embrace final causes, as some philosophers (e.g., Tay-
lor, 1964) have done, is an invitation for it to turn the clock back.
276 13. Concluding Remarks

A third common type of pseudoexplanation is mentalist, or the attribu-


tion of behavioral or mental states to further mental states. Examples: A
dreamed of B because A secretly longed for B (Freud); perceiving and
thinking are computing-and so is computing (computationalism). Trying
to explain the mental by the mental is as hopeless as trying to explain
behavior in terms of conspicuous variables. Genuine explanation involves
reference to some concrete (and often not directly observable) mecha-
nism. More on this in a while.
A fourth type of psychological pseudoexplanation is metaphorical, or
by analogy with physical or social processes, or with machines. Exam-
ples: Consciousness is like a stream; memory is encoded information; the
brain works like a computer. An account of behavioral or mental pro-
cesses in analogical terms is just a fancy redescription. It does not derive
(deduce) the fact to be explained from law statements and data, and it
does not involve any neural mechanism. It only creates the illusion of
understanding through familiarity.
This is not to deny the heuristic value of some analogies. Some progress
in psychology has been inspired by analogies from physics, chemistry,
and technology. (See e.g., Marshall, 1977.) Metaphors can suggest scien-
tific hypotheses, and sometimes they summarize and bring home other-
wise abstruse ideas. For example, animism was summed up by Plato by
stating that the mind is to the body as the pilot to the ship; and the
psychoneural identity hypothesis was summarized by Vttal (1978) in the
apt metaphor: "The mind is to the brain as rotation is to the wheel."
But the point is that analogies are not theories, whence analogical re-
descriptions have no explanatory power. For all their seductive power
they have no deductive structure-that is, nothing follows logically from
them. For this reason the scientific status of a discipline is directly propor-
tional to the number of confirmed theories it posseses, and inversely
proportional to the number of analogies currently used in it. Viewed in
this light, psychology is not doing too well, for it uses literally hundreds of
metaphors: animistic, spatial, hydraulic, electrical, chemical, technologi-
cal, and so on. (See e.g., Gentner & Grudin, 1985.)
So much for pseudoexplanation in psychology. Let us now make a
quick review of explanation types having a scientific potential, starting
with genetic explanations. A genetic explanation of a behavioral or men-
tal trait A is proposed when it is conjectured that A is inherited, not
learned. More precisely, the hypothesis is that a gene (or more likely a
group of linked genes) exists that controls the morphogenesis of a neural
system B, the specific function of which is A. Only the identification of
such a gene can provide a direct confirmation of the hypothesis. (Regret-
tably, thus far most cases of confirmation of this type have concerned
abnormalities.) The indirect way of confirming a genetic hypothesis is to
show the insensitivity of the corresponding trait to environmental varia-
tions. In any event, genetic explanations are speculative as long as the
13.3. Explanation 277

corresponding genetic hypotheses have not been confirmed one way or


the other. Unfortunately this elementary methodological remark has often
been forgotten in the spirited debates about the nature-nurture question.
(Recall section 6.2 and 7.3.)
A second kind of possible scientific explanation in psychology is the
developmental type. An account of the emergence of skills in terms of
both the maturation of the nervous system and the occurrence of suitable
environmental stimuli will, if correct, be a genuine developmental expla-
nation. On the other hand, Freud's attribution of personality traits to
early toilet training, and of neuroses to libido repression during infancy,
are developmental conjectures without an empirical basis. As for Piaget's
account of the appearance of skills in terms of development stages, it may
well be true but it is not an explanation: It is a data base. Thus, saying that
Johnny cannot reason correctly because he is still at the stage of concrete
operations (7-11 years) exemplifies the tautology "If X is impossible,
then X is not the case." In short, pure (brainless) developmental (or
genetic) psychology, though doubtless important, explains nothing. Be-
cause human development is a biosocial process of neural reorganization
intertwined with socialization, only developmental biopsychology jointly
with social psychology holds promise of scientific explanations of the
developmental kind.
Next come environmental explanations (i.e., accounts in terms of sen-
sory stimulation or some other exogenous factor). No behavioral or men-
tal fact can be satisfactorily explained without the help of some premises
(hypotheses or data) concerning the subject's environment, if only be-
cause the latter is necessary to keep the organism alive. However, envi-
ronmental agents affect behavior or mind only if they have a significant
impact on the nervous system. Hence no purely environmental account
can constitute more than part of an explanation: Recall section 6.2.
Evolutionary explanations point to the selective advantage or disadvan-
tage of a behavioral or mental trait. For example, biologically inclined
psychologists and linguists agree that our cognitive and linguistic skills
have evolved from some more primitive capacities. And some (e.g., Oak-
ley, 1983) have suggested the reasonable but so far untested hypothesis
that the new learning capacities emerged with the formation of new neuro-
nal modules. However, evolutionary psychology is still little more than a
research project. Therefore, evolutionary explanations must be handled
with care. For example, it has been claimed that the capacity to feel pain
evolved because of its survival value: It allows us to identify noxious
stimuli and thus to avoid them. However, a high pain threshold under
duress (e.g., in childbirth or in battle) is also supposed to confer selective
advantage. Moral: As long as evolutionary psychology continues to be
poor in laws, trends, and data, we should regard evolutionary explana-
tions as speculations. Still, we should continue to propose and investigate
them.
278 13. Concluding Remarks

Finally, physiological explanations are explanations in terms of physio-


logical (in particular neurophysiological and neuroendocrinological)
terms. In particular, a physiological explanation of a behavioral fact (trait,
event, process) A is proposed when it is hypothesized that A is controlled
by some neural system B (which may in turn be modulated by some
endocrine system C). And a physiological explanation of a mental fact
(trait, event, process) A is proposed when it is hypothesized that A is
identical with a trait of, or an event or a process in, some neural system B
(possibly modulated by some endocrine system C). For example, volun-
tary movement is controlled by certain neuron assemblies in the frontal
lobes; and visual imagery may be conjectured to be identical with an
activity in the visual cortex initiated in the same place or in some other
brain subsystem.
When submitting a physiological explanation of a behavioral fact we do
not just assert that a response is associated with a stimulus, much less
that it is caused by some immaterial mental entity-although it may be
caused by one of those brain processes we call 'mental.' Likewise, when
suggesting a physiological explanation of a mental fact we do not just
assert that it has a neural correlate, let alone that it is an effect of a mental
event occurring in an immaterial mind. In either case we propose a more
or less precise physiological mechanism working according to definite
though, alas, still poorly known laws.
The word 'mechanism' is used here in its broad sense of process in a
concrete system. For example, the conduction of a nerve pulse ("infor-
mation flow' ') along the axon of a neuron is a mechanism, and so is the
binding of a neurotransmitter to a receptor-as well as its blocking by
some other molecule. On the other hand, neither Piaget's proposed
"mechanisms" (e.g., that of equilibration), nor the once famous TOTE
(test-operate-test-exit) sequence is a mechanism proper, for neither ex-
plicitly involves the nervous system. Strictly speaking there are no behav-
ioral or mental mechanisms in themselves, that is, aside from physiologi-
cal (in particular neuromuscular and neuroendocrine) mechanisms-just
as there are no chemical mechanisms aside from reactants, or no social
mechanisms outside of people. Consequently, brainless psychology, be it
behaviorist or mentalist, can at best describe, never explain. Only biopsy-
chology can explain behavior and mind. (For the alternative standard
view see e.g., Borger & Cioffi, 1970; Marx & Hillix, 1973.)
Molar or pure psychology supplies some of the facts to be explained,
whereas biopsychology investigates the possible mechanisms explaining
those facts. This investigation involves, inter alia, the disclosure of the
neural "centers" and pathways that might "subserve" (i.e., control or
perform) the behavioral or mental acts of interest. Anatomy is thus an
important resource science for psychology.
Suppose, for example, that anatomical investigation reveals that a sen-
13.3. Explanation 279

(a) (b) (c)

FIGURE 13.2. S = Sensory system, M = Motor system, E = Emotive system, C =


Cognitive system. (a) The projection of S to M allows for the sensory regulation of
motor behavior. (b) E is in a position to receive inputs from S as well as to
influence M. (c) The projection of C to M allows for the cognitive steering of
behavior.

sory system S has afferences to a motor center M. Then we may conjec-


ture that a stimulus acting on S will cause a response of M: See Figure
13.2, part (a). If, on the other hand, both Sand M are connected to some
component E of the limbic system, we may guess that the sensory stimu-
lus may alter the emotional state, which may in tum regulate the motor
response: See Figure 13.2, part (b). Finally, if S is connected with some
cognitive center C, which is also connected with the emotive system E,
we may conjecture that the animal in question is capable of learning to
control its motor responses to the stimuli impinging upon S: See Figure
13.2, part (c). Moreover, the mere neighborhood of certain groups of
neurons may suggest looking for previously unknown connections among
them; it may predict behavioral or mental facts of a previously unknown
kind. Something similar occurs with neuroendocrine and neuroimmune
connections.
The best that can happen to any hypothesis concerning a mechanism is
to become a component of a theory (i.e., a hypothetico-deductive sys-
tem). One reason for this is that, by becoming interconnected, the various
hypotheses complement and support one another. A second reason is
that, whereas every experimental finding can be explained by a number of
alternative hypotheses, a theory, by making a large number of predictions
of several kinds, can face a whole array of experiments. It would be
miraculous for a totally false theory to account reasonably well for a
wealth of experiments: See Figure 13.3. (For details see Bunge, 1983b.)
In sum, psychology suffers from a glut of metaphors and a shortage of
theories, particularly well-confirmed theories of the mechanismic kind
(i.e., capable of explaining behavioral or mental phenomena). Therefore,
theory construction should be assigned priority in contemporary psy-
chology.
280 13. Concluding Remarks

TUD

h1 h2 h3

'\ t 1
\
\
I
I /
I
I
I \
I \
\ I
I \
I \
\
P3

I
I
I \
\
\
\ I I
ie o
I \
o
I \ I \
6 6 d b

FIGURE 13.3. (a) Experimental finding e may be taken to support not just one but
several rival hypotheses h" h2' h3, and so forth. (b) Experimental findings el to e6,
by supporting the predictions p" P2, and P3 derived from theory T and data D,
support T. Full lines: deduction. Dotted lines: confirmation.

13.4 Prospects
What is the future of psychology, assuming optimistically that humankind
will not be destroyed by a nuclear war, and that it will continue to engage
in scientific research? We cannot foretell the future of psychology, or of
any other discipline, because we do not know of any laws of the evolution
of knowledge. But we do know that the only serious limits to the growth
of knowledge are of a social kind, whence they can be overcome (Bunge,
1978). And we can do better than to prophesy and wait: We can shape the
future of psychology by planning for it.
Now, every plan must start by taking stock of the present. In our view
current psychology is characterized by the following traits:
(1) Rapid growth of experimental research, particularly in physiological
psychology, neuropsychology, psychophysics, developmental psy-
chology, clinical psychology, and biological psychiatry.
(2) Methodological sophistication in experimental basic research, partic-
ularly in biopsychology and psychophysics-far less so in applied
psychology.
(3) Theoretical stagnation. There are too few theories about behavioral
and mental processes, and it would seem that most of the existing ones
are wrong (Tulving, 1985b); worse, much theoretical work has been
misguided by the computer metaphor, whereas there is a dearth of
theories and models linking molar psychological variables to neuro-
physiological ones.
(4) Fragmentation. An exaggerated division of labor has resulted in weak
links among the various branches of psychology, to the point that
some of them (e.g., cognitive psychology) are becoming isolated from
the rest.
13.4. Prospects 281

(5) Ecological invalidity of much psychological research (i.e., insufficient


interest in such important everyday problems as "Why do we need
sleep?" "Why does the common cold impair our thought processes?"
"Why can we remember so many trivia while forgetting essentials?"
"What is the neuroendocrine mechanism of the process of falling in
love?" and "What is the neural mechanism of originality?").
(6) Divorce of practice from research. Much of applied psychology (clini-
cal, educational, consulting, etc.) is empirical or worse, namely domi-
nated by pseudoscience (e.g., psychoanalysis), and it continues to use
such primitive tools as projective and multiple choice tests.
(7) Philosophical obsolescence. Vitalism, teleology, and dualism, as well
as crude versions of rationalism, empiricism, and conventionalism,
continue to linger in psychology, aided and abetted by philosophers.

Everyone of the first six traits in the preceding list has been noted by
some psychologists, but the seventh is usually overlooked by everyone.
Yet the relevance of philosophy to psychology is quite obvious, as em-
phasized in chapter 1. This is particularly so with regard to the third
feature listed, namely theoretical stagnation. Indeed, it may be argued
that the main causes for the theoretical underdevelopment of psychology
are philosophical, particularly the following: (a) the positivist ban on
theorizing, so enthusiastically observed by radical behaviorism; (b) the
confusion of theory with metaphor; (c) mind-body dualism, which deta-
ches psychology from biology and encourages wild speculation about
immaterial (hence empirically inaccessible) entities and events; and (d)
the very existence of philosophical (or armchair) psychology, which gives
both theory and philosophy a bad name among experimental psycholo-
gists. Because philosophers have been largely responsible for this lamen-
table state of affairs, it behooves them to make amends. But of course it is
up to psychologists to purge their own brains of obsolete philosophies and
to engage more vigorously and rigorously in theorizing.
Having taken stock of the situation we can do something about it. For
example, theorizing would be encouraged by refusing to publish papers
restricted to presenting raw data, and by having psychology students
study more mathematics-in particular, by teaching them probability be-
fore mathematical statistics. Fragmentation can be decreased by multiply-
ing the numbers of cross-disciplinary research teams, workshops, and
seminars, as well as by requiring psychology students to learn more
neuroscience and social science. Ecological validity can be enhanced by
refusing to publish much correct but insignificant work more appropriate
for technical reports. Practice can be forced to marry research by requir-
ing all professionals to take a science degree presupposing intensive expo-
sure to psychological experiment and theory. And the philosophical per-
spective can be updated by having psychology students take some (good)
courses in the philosophy of science.
282 13. Concluding Remarks

By doing all that at the same time, psychologists would build a bright
future for their science. Psychology might well acquire in the twenty-first
century the glamour that distinguished physics in the first half of our own
century, and which biology possesses in the second half. However, imple-
menting these measures involves both a philosophical reorientation of the
psychological community, and persuading the powers that be that it
would be worthwhile to invest in learning about behavior and mind at
least a small fraction of what is being wasted on improving the overkill
capacity.

13.5 Philosophical Harvest


We proceed to summarizing the philosophical outputs of biopsychology
and social psychology noted in the previous chapters. We shall group
them into two sets: ontological (or regarding the nature of things) and
epistemological (or concerning our knowledge of things).

Ontological Crop
(1) Psychoneural identity: Mental processes are brain processes. Put
negatively: Mind is not separate from body, anymore than digestion is
detachable from the digestive tract.
(2) Emergentism: The subsystems of the nervous system that control
behavior or perform mental functions have properties that their compo-
nents lack. They have emerged in the course of evolutionary or develop-
mental processes, and some of them submerge as a result of sickness or
aging.
(3) Mind is causally efficient: Mental processes influence other brain
processes, and occasionally they have motor outlets. As well, they affect
(and are affected by) the other two regulatory systems of the body: the
endocrine and the immune.
(4) Localization cum integration: Except for memory and learning,
which are capabilities of all plastic neural systems, every mental "fac-
ulty" is the specific function of a special brain subsystem. However,
because the various subsystems are anatomically linked to one another,
no behavioral or mental "faculty" is separate from all the others. In
particular, cognition is fueled by motivation and it can steer movement.
Put negatively: Neither behavior nor mind is modular.
(5) Interaction with society: Behavior and mind-particularly learning,
perception, thought, and social behavior-are strongly influenced by so-
cial circumstances and, in turn, they contribute to shaping the latter
through both behavior and language.
13.5. Philosophical Harvest 283

Epistemological Crop
(6) Critical realism: There are things in themselves, i.e., existing inde-
pendently of the knowing subject, who can come to know some of them
partially and gradually. Put negatively: We do not make the world, though
we can change it-alas, not always for the better.
(7) Ratioempiricism: Scientific research, in psychology and elsewhere,
combines reason with experience. Put negatively: Radical rationalism,
though adequate in pure mathematics, is just as inadequate as radical
empiricism is in the fields of factual science.
(8) Reductionism: Psychology is a part of biology. That is, the explana-
tion of behavioral and mental processes, unlike their mere description,
calls for conjecturing and uncovering the corresponding neural (or neuro-
muscular, or neuroendocrine, or neuroimmune, or neuroendocrinoim-
mune) mechanisms. Put negatively: Nonbiological psychology is shallow
and incapable of explaining anything.
(9) Dependence upon social science: The explanation of some behav-
ioral and mental processes requires certain categories belonging to social
science. Put negatively: A psychology oblivious of the social matrix is just
as inadequate as a geography that ignores the atmosphere.
(10) Specificity and dependence: Although psychology is a very special
science, it is not independent but lies in the intersection of biology and
social science. Put negatively: There is no wall between the Naturwis-
senschaften and the Geisteswissenschaften-except in the brains of some
philosophers.

Let us comment briefly on the leading members of each of the two


groups of philosophical outputs, starting with (1). In light of the previous
chapters, the alternative to psychoneural monism, namely dualism, must
be seen as a serious metaphysical pathology. Like semantic aphasia, the
apraxias, and other neurological disorders, psychoneural dualism may be
seen as a disconnection syndrome: in this case as a result of the discon-
nection of psychology from biology, as well as from philosophy and sci-
ence. This double disconnection results in psychical blindness for physio-
logical, developmental, and evolutionary psychology, and also in the
estrangement of clinical psychology and psychiatry, on the one hand,
from neurology, endocrinology, and immunology on the other. Because
the syndrome originates in religious fundamentalism and is supported by
obsolete philosophy, it is advisable for psychologists to keep them both
outside their laboratory, and to intensify their commerce with both biol-
ogy and science-oriented philosophy.
As for (6), critical realism, all psychologists adopt it tacitly when exper-
284 13. Concluding Remarks

imenting, although they occasionally forget about it when theorizing. In


fact, all experimental psychologists and applied psychologists take it for
granted that their subjects of study, as well as their tools and their envi-
ronments, exist by themselves. (The fact that psychoanalysts analyze
mythical characters, such as Hamlet and Othello, only shows how far
from science they are. However, they always send their bills to real
persons.) In particular, all theories of perception presuppose that we
normally perceive things out there-when this is not so, one speaks of
illusions or hallucinations. Of course, such theories also admit that per-
ception is not passive: that the subject contributes his or her own memo-
ries and expectations. Something similar holds for theories of memory
and learning: They are all in line with critical or constructive (vs. naive)
realism. (For the varieties of realism see Bunge, 1983b. For a criticism of
the antirealist philosophies of physics see Bunge, 1973a and 1985a.) True,
some ethologists, like von Uexkiill (1921), and child psychologists, such
as Piaget (1954), have written about the "construction of reality" by the
animal or the child. But they usually mean the construction of maps,
models, images, or conceptual representations of reality. They do not
believe in omnipotence, much less in miracles. When in their right minds,
psychologists behave as critical realists, not as subjectivists, let alone
solipsists. Moreover, they are suspicious of signs of loss of touch with
reality, as in the cases of fabulation, autism, and schizophrenia. If anti-
realist philosophers were taken seriously they would be institutionalized,
for realism of some sort-whether naive, critical, or scientific-is a
symptom of sanity.

13.6 Summing Up
We have argued for the reductionist thesis that mental phenomena are
biological processes, as well as for the emergentist thesis that mentation is
a qualitative novelty emerging at certain points in the evolution of biopo-
pulations and in the development of individuals of some animal species.
Moreover, we have argued that the emergence of mental abilities can be
explained, at least in principle, by identifying it with the organization or
reorganization of neuronal systems (i.e., the change in connectivity), ei-
ther spontaneously (without any external causes) or in response to
changes occurring in other parts of the body or in the environment. We
have thus combined ontological emergentism with moderate epistemolog-
ical reductionism.
However, in the case of behavior and mind reduction is insufficient; it
must be supplemented with a study of processes occurring in adjoining
domains, sometimes higher-level ones. In particular, an adequate under-
standing of behavior and mentation in the case of gregarious animals calls
for the cooperation of social science. The two movements, reduction and
13.6. Summing Up 285

integration, supplement one another and they should be favored as a


means to decrease the current fragmentation of psychology-a fragmen-
tation that honors neither the unity of the whole animal nor its place in a
social context.
Philosophers can learn much from science, in particular from psychol-
ogy and neuroscience. Regrettably, most philosophers of psychology
have dealt with only folk psychology, early behaviorism, psychoanalysis,
or computer psychology, and have been effective only at bashing them. It
is high time philosophers became acquainted with contemporary scientific
psychology as well as with its neuroscientific basis and its sociological
cap. Only thus will they be able to enrich the philosophy of psychology
with novel and correct ideas while, at the same time, helping psycholo-
gists realize some of the philosophical presuppositions and implications of
their own work. Psychology has not advanced by getting rid of philosophy
but by replacing false or barren philosophical ideas with true or fertile
ones.
References
Adamson, W. c., & K. K. Adamson (Eds.). (1979). A handbook of specific
learning disabilities. New York: Gardner.
Ader, R., & N. Cohen. (1985). CNS-immune system interactions: Conditioning
phenomena. The Behavioral and Brain Sciences 8:379-395.
Agassi, J. (1975). Science influx. Dordrecht-Boston: Reidel.
Aggleton, J. P., & M. Mishkin. (1983). The amygdala: "Sensory gateway to the
emotions." In R. Plutchik & H. Kellerman (Eds.), Emotion: Theory, research
and experience: Vol. 3. Biological foundations of emotion. New York: Aca-
demic Press.
Aguayo, A. J. (1985). Axonal regeneration from injured neurons in the adult
mammalian central nervous system. In C. W. Cotman (Ed.), Synaptic plastic-
ity. New York: Guilford Press.
Alcock, J. E. (1981). Parapsychology: Science or magic? Oxford and Elmsford,
NY: Pergamon Press.
Alcock, J. E. (1984). Parapsychology's last eight years: A lack-of-progress report.
The Skeptical Inquirer 8:312-321.
American Psychiatric Association. (1980). Diagnostic and statistical manual of
mental disorders (3rd ed.). Washington, DC: American Psychiatric Associa-
tion.
Anderson, J. A., J. W. Silverstein, S. A. Ritz, & R. S. Jones. (1977). Distinctive
features, categorical perception, and probability learning. Some applications of
a neural model. Psychological Review 84:413-451.
Anderson, J. R. (1983). The architecture of cognition. Cambridge, MA: Harvard
University Press.
Andersson, M. (1982). Female choice selects for extreme tail length in a widow-
bird. Nature 299:818-820.
Ardila, R. (1979a). Los origenes del comportamiento humano. Barcelona: Fon-
tanella.
Ardila, R. (1979b). Walden tres. Barcelona: CEAC.
Ardila, R. (1980a). Terapia del comportamiento. Bilbao, Spain: Desch!e de
Brouwer.
Ardila, R. (1980b). Die Zukunft der Familie. Verhalten 2(2):71-82.
Ardila, R. (1985). La evaluaci6n comportamental como alternativa al diagn6stico
psiquiatrico tradicional. Revista Mexicana de Psicologia 2:62-68.
Ardila, R. (1986). Impacto psicol6gico de la guerra nuclear. Bogota: Catalogo
Cientifico.
Ardila, R. (1987). La sintesis experimental del comportamiento. Madrid: Alham-
bra.
Asano, T., T. Kojima, T. Matsuzawa, K. Kubota, & K. Murofushi. (1982). Object
and color naming in chimpanzees (Pan troglodytes). Proceedings of the Japan
Academy 58B:118-122.
Bachelard, G. (1938). La formation de l' esprit scientifique. Paris: Vrin.
288 References

Bachevalier, J., & M. Mishkin. (1984). An early and a late developing system
for learning and retention in infant monkeys. Behavioral Neuroscience 98:770-
778.
Bachevalier, J., J. K. Parkinson, & M. Mishkin. (1985). Visual recognition in
monkeys: Effects of separate vs. combined transection of fornix and amygdalo-
fugal pathways. Experimental Brain Research 57:554-561.
Bandura, A. (1978). The self system in reciprocal determinism. American Psy-
chologist 33:344-358.
Baranyi, A., & O. Feher. (1981). Synaptic facilitation requires paired activation of
convergent pathways in the neocortex. Nature 290:413-415.
Barlow, H. B. (1972). Single units and sensation: A neuron doctrine for percep-
tual psychology. Perception 1:371-380.
Bartlett, F. C. (1932). Remembering. Cambridge, U.K.: Cambridge University
Press.
Beaulieu, A. (1983). Le secret (?) de l'intelligence de Richelieu. Nouvelles de la
Repub/ique des Lettres 11:97-104.
Bekesy, G. von. (1967). Sensory inhibition. Princeton, NJ: Princeton University
Press.
Bekesy, G. von. (1968a). Problems relating psychological and electrophysiologi-
cal observations in sensory perception. Perspectives in Biology and Medicine
11: 179-194.
Bekesy, G. von. (1968b, March-April). Feedback phenomena between the
stringed instrument and the musician. The Rockefeller University Review.
Beninger, R. J., S. B. Kendall, & c. H. Vanderwolf. (1974). The ability of rats to
discriminate their own behaviours. Canadian Journal of Psychology 28:79-91.
Berlyne, D. E. (1975). Behaviourism? Cognitive theory? Humanistic psychol-
ogy?-To Hull with them all! Canadian Psychological Review 16:69-80.
Berthoz, A., & G. Melvill Jones (Eds.). (1985). Adaptive mechanisms in gaze
control: Facts and theories. Amsterdam-New York-Oxford: Elsevier.
Bindra, D. (1976). A theory of intelligent behavior. New York: Wiley.
Bindra, D. (1984). Cognition: Its origin and future in psychology. Annals ofTheo-
retical Psychology 1:1-29.
Bitterman, M. E. (1975). The comparative analysis of learning. Science 188:699-
709.
Bitterman, M. E. (1984). Learning in man and other animals. In V. Sarris & A.
Parducci (Eds.), Perspectives in psychological experimentation: Toward the
year 2000 (pp. 59-70). Hillsdale, NJ: Erlbaum.
Bliss, T. V. P. (1979). Synaptic plasticity in the hippocampus. Trends in Neuro-
Sciences 2:42-45.
Bliss, T. V. P., & T. Lflmo. (1973). Long-lasting potentiation of synaptic trans-
mission in the dentate area of the anaesthesized rabbit following stimulation of
the perforant path. Journal of Physiology 232:331-356.
Bloom, F. E. (1975). The gains in brain are mainly in the stain. In F. G. Worden,
J. P. Swazey, & G. Adelman (Eds.), The neurosciences: Paths of discovery (pp.
211-227). Cambridge, MA: MIT Press.
Boakes, R. (1984). From Darwin to behaviorism. London: Cambridge University
Press.
Boden, M. (1983). Artificial intelligence and animal psychology. New Ideas in
Psychology 1: 11-33.
References 289

Borger, R., & F. Cioffi (Eds.). (1970). Explanation in the behavioural sciences.
London and New York: Cambridge University Press.
Boring, E. G. (1942). Sensation and perception in the history of experimental
psychology. New York: Appleton-Century-Crofts.
Boring, E. G. (1950). A history of experimental psychology (2nd ed.). New York:
Appleton-Century-Crofts.
Bouchard, T. J., Jr., & M. McGue. (1981). Familial studies of intelligence: A
review. Science 212:1055-1059.
Bower, T. G. R. (1974). Development in Infancy. San Francisco: W. H. Freeman.
Boyd, R., & P. J. Richerson. (1985). Culture and the evolutionary process. Chi-
cago: University of Chicago Press.
Bredenkamp, J., & H. Feger (Eds.). (1983). Hypothesenprilfung, Enzyklopiidie
der Psychologie, (Series I, Vol. 5). G6ttingen: Hogrefe.
Brentano, F. (1955). Psychologie vom empirischen Standpunkt. Hamburg: Felix
Meiner. (Original work published 1874) English transl. of the chapter "The
distinction between mental and physical phenomena" in R. M. Chisholm (Ed.),
(1960). Realism and the background of phenomenology. Glencoe, IL: Free
Press.
Bridgman, P. W. (1927). The logic of modern physics. New York: Macmillan.
Bridgman, P. W. (1959). "The Logic of Modern Physics" after thirty years.
Daedalus 88:518-525.
Broad, C. D. (1949). The relevance of psychical research to philosophy. Philoso-
phy 24:291-309.
Brown, S. C. (Ed.). (1974). Philosophy of psychology. London: Macmillan.
Bruce, C. J., & M. E. Goldberg. (1984). Physiology of the frontal eye fields.
Trends in NeuroSciences 7:436-441.
Bruner, J. S., J. J. Goodnow, & G. Austin. (1956). A study of thinking. New York:
Wiley.
Brunswik, E. (1955). The conceptual framework of psychology. In O. Neurath, R.
Carnap, & C. Morris (Eds.), International encyclopedia of unified science (Vol.
1, No.6.). Chicago: University of Chicago Press.
Buhler, K. (1927). Die Krise der Psychologie. Jena: Gustav Fischer.
Bunge, M. (1956). Do computers think? British Journal for the Philosophy of
Science 7:139-148, 212-219.
Bunge, M. (1962). Intuition and science. Englewood Cliffs, NJ: Prentice-Hall.
(Reprinted 1975, Greenwood Press, Westport, CT).
Bunge, M. (l963a). The myth of simplicity. Englewood Cliffs, NJ: Prentice-Hall.
Bunge, M. (1963b). A general black box theory. Philosophy of Science 30:346-
358.
Bunge, M. (1964). Phenomenological theories. In M. Bunge (Ed.), The critical
approach: Essays in honor of Karl Popper (pp. 234-254). Glencoe, IL: Free
Press.
Bunge, M. (1967a). Scientific research: Part I. The search for system. New York:
Springer-Verlag.
Bunge, M. (l967b). Scientific research: Part II. The search for truth. New York:
Springer-Verlag.
Bunge, M. (1971). Is scientific metaphysics possible? Journal of Philosophy
68:507-520.
Bunge, M. (1973a). On confusing 'measure' with 'measurement' in the methodol-
290 References

ogy of behavioral science. In M. Bunge (Ed.), The methodological unity of


science (pp. 105-122). Dordrecht: Reidel.
Bunge, M. (1973b). Philosophy of physics. Dordrecht: Reidel.
Bunge, M. (1973c). Method, model and matter. Dordrecht: Reidel.
Bunge, M. (1974a). Sense and reference. Dordrecht: Reidel.
Bunge, M. (1974b). Interpretation and truth. Dordrecht: Reidel.
Bunge, M. (1977a). The furniture of the world. Dordrecht and Boston: Reidel.
Bunge, M. (1977b). Levels and reduction. American Journal of Physiology
233:R75-82.
Bunge, M. (1977c). General systems and holism. General Systems 22:87-90.
Bunge (1977d). Emergence and the mind. Neuroscience 2:501-509.
Bunge, M. (1978). The limits of science. Epistemologia 1:11-32. (Reprinted 1980
in The Physiologist 23:7-13).
Bunge, M. (1979a). A world of systems. Dordrecht and Boston: Reidel.
Bunge, M. (1979b). Causality in modern science. New York: Dover.
Bunge, M. (1979c). A systems concept of society: Beyond individualism and
holism. Theory and Decision 10: 13-30.
Bunge, M. (1980). The mind-body problem. Oxford and Elmsford, NY: Perga-
mon Press.
Bunge, M. (1981). Scientific materialism. Dordrecht and Boston: Reidel.
Bunge, M. (1982). Ciencia y desarrollo. Buenos Aires: Siglo XX.
Bunge, M. (1983a). Exploring the world. Dordrecht and Boston: Reidel.
Bunge, M. (1983b). Understanding the world. Dordrecht and Boston: Reidel.
Bunge, M. (1983c). Speculation: Wild and sound. New Ideas in Psychology 1:3-6.
Bunge, M. (1983d). Comment on a paper by Fedanzo. Journal of Social and
Biological Structures 6: 159-160.
Bunge, M. (1984). Philosophical problems in linguistics. Erkenntnis 21:107-173.
Bunge, M. (1985a). Philosophy of science and technology. Part I: Formal and
physical sciences. Dordrecht and Boston: Reidel.
Bunge, M. (1985b). Philosophy of science and technology: Part II. Life science,
social science and technology. Dordrecht and Boston: Reidel.
Bunge, M. (1985c). Seudociencia e ideologia. Madrid: Alianza Dniversidad.
Bunge, M. (1985d). Racionalidad y realismo. Madrid: Alianza Universidad.
Bunge, M. (1985e). From mindless neuroscience and brainless psychology to
neuropsychology. Annals of Theoretical Psychology 3:115-133.
Bunge, M. (19850. On research strategies in psychology. Reply to commentators.
Annals of Theoretical Psychology 3: 151-156.
Bunge, M. (1985g). Types of psychological explanation. In J. McGaugh (Ed.),
Contemporary psychology: Biological processes and theoretical issues (pp.
389-501). Amsterdam and New York: North Holland-Elsevier.
Bunge, M. (1986). A philosopher looks at the current debate on language acquisi-
tion. In I. Gopnik & M. Gopnik (Eds.), From models to modules (pp. 229-239).
Norwood, NJ: Ablex PubIs. Co.
Bumstine, T. E., W. T. Greenough, & R. C. Tees. (1984). Intermodal compensa-
tion following damage or deprivation: A review of behavioral and neural evi-
dence. In C. R. Almli & S. Finger (Eds.), Early brain damage (Vol. 1, pp. 3-
34). New York: Academic Press.
Burtt, E. A. (1932). The metaphysical foundations of modern physical science
(2nd ed.). London: Routledge & Kegan Paul.
References 291

Calford, M. B., M. L. Graydon, M. F. Huerta, J. H. Kaas, & J. D. Pettigrew.


(1985). A variant of the mammalian somatotopic map in a bat. Nature 313:477-
479.
Cartwright, D. S. (1979). Theories and models of personality. Dubuque, IA:
Brown.
Chang, F. F., & W. T. Greenough. (1984). Transient and enduring morphological
correlates of synaptic activity and efficacy change in rat hippocampal slice.
Brain Research 309:35-46.
Chomsky, N. (1959). Review of Skinner's Verbal Behavior. Language 35:26-58.
Chomsky, N. (1975). Reflections on language. New York: Pantheon Books.
Clement, J. (1982). Student's preconceptions in introductory mechanics. Ameri-
can Journal of Physics 50:66-71.
Cohen, P. R., & E. A. Feigenbaum (Eds.). (1981-1982). The handbook of artifi-
cial intelligence (3 vols.). Los Altos, CA: William Kaufmann.
Conel, J. L. (1939-1967). The postnatal development of the human cerebral cor-
tex, 7 volumes. Cambridge, MA: Harvard University Press.
Cowey, A., & L. Weiskrantz. (1975). Demonstration of cross-modal matching in
rhesus monkeys, Macaca mulatta. Neuropsychologia 13:117-120.
Craik, K. J. W. (1943). The nature of explanation. Cambridge, U.K.: Cambridge
University Press.
Crews, D., & M. C. Moore. (1986). Evolution of mechanisms controlling mating
behavior. Science 231:121-125.
Crick, F. (I 984a). Function of the thalamic reticular complex: The searchlight
hypothesis. Proceedings of the National Academy of Sciences of the USA
81 :4586-4590.
Crick, F. (1984b). Memory and molecular turnover. Nature 312: 101.
Critchley, M. (1960). Evolution after Darwin. Chicago: University of Chicago
Press.
Damasio, A. R., H. Damasio, & G. W. Van Hoesen. (1982). Prosopagnosia:
Anatomic basis and behavioral mechanisms. Neurology 32:331-341.
Damasio, A. R., & G. W. Van Hoesen. (1983). Emotional disturbances associated
with the focal lesions of the limbic frontal lobe. In K. M. Heilman & P. Satz
(Eds.), Neuropsychology of human emotion (pp. 85-110). New York: Guilford
Press.
Darwin, C. R. (1872). The expression of the emotions in man and animals. Lon-
don: Appleton.
Davidson, D. (1970). Mental events. In L. Foster & J. W. Swanson (Eds.), Expe-
rience and theory (pp. 79-101). Amherst: University of Massachusetts
Press.
Davidson, D. (1974). Psychology as philosophy. In S. C. Brown (Ed.), Philosophy
of psychology (pp. 41-52). London: Macmillan.
Davidson, J. M. (1980). The psychobiology of sexual experience. In J. M. David-
son & R. J. Davidson (Eds.), The psychobiology of consciousness (pp. 271-
332). New York: Plenum.
Davidson, J. M. and R. J. Davidson (Eds.). (1980). The psychobiology of con-
sciousness. New York: Plenum.
Dawson, M. E., & J. J. Furedy. (1976). The role of awareness in human differen-
tial autonomic classical conditioning: The necessary-gate hypothesis. Psy-
chophysiology 13:50-53.
292 References

Delius, J. D., & B. Nowak. (1982). Visual symmetry recognition by pigeons.


Psychological Research 44:199-212.
Dell, G. (1985). Positive feedback in hierarchical connectionist models: Applica-
tion to language production. Cognitive Science 9(1):3-24.
Dennett, D. C. (1978). Brainstorms: Philosophical essays on mind and psychol-
ogy. Montgomery, VT: Bradford Books.
Descartes, R. (1649). Traite de l'ame. In C. Adam & P. Tannery (Eds.) Oeuvres
de Descartes, 11 vols. Paris: Cerf, 1909.
Desimone, R., T. D. Albright, C. G. Gross, & C. Bruce. (1984). Stimulus-selec-
tive properties of inferior temporal neurons in the macaque. Journal of Neuro-
science 4:2051-2062.
Diaconis, P. (1978). Statistical problems in ESP research. Science 201:131-
136.
Dickinson, A., & N. J. Mackintosh. (1978). Classical conditioning in animals.
Annual Review of Psychology 29:587-612.
Dimond, S. J. (1976). Brain circuits for consciousness. Brain, Behavior and Evo-
lution 13:376-395.
Dimond, S. J. (1980). Neuropsychology. London: Butterworths.
Dixon, N. F. (1971). Subliminal perception. London: McGraw-Hill.
Doty, R. W., Sr. (1975). Consciousness from neurons. Acta Neurobiologica Ex-
perimentalis 35:791-804.
Dumont, J. P. C., & R. M. Robertson (1986). Neuronal circuits: Evolutionary
perspective. Science 233:849-853.
Dworkin, B. R., & N. E. Miller. (1986). Failure to replicate visceral learning in
acute curarized rat preparation. Behavioral Neuroscience 100:299-314.
Easter, S. S., Jr., D. Purves, P. Rakic, & N. C. Spitzer. (1985). The changing
view of neural specificity. Science 230:507-511.
Eccles, J. C. (1953). The neurophysiological basis of mind. Oxford: Clarendon
Press.
Eccles, J. C. (Ed.). (1966). Brain and conscious experience. Berlin-Heidelberg-
New York: Springer-Verlag.
Eccles, J. C. (1980). The human psyche. Berlin-Heidelberg-New York: Springer-
Verlag.
Eccles, J. C. (1982). How the self acts on the brain. Psychoneuroendocrinology
7:271-283.
Eccles, J. C, and D. N. Robinson. (1985). The wonder of being human. Boston
and London: New Science Library (Shambhala).
Edelman, G. M. (1978). Group selection and phasic reentrant signaling: A theory
of higher brain function. In G. M. Edelman & V. B. Mountcastle, The Mindful
Brain (pp. 51-100). Cambridge, MA: MIT Press.
Edelman, G. M., & V. B. Mountcastle (1978). The mindful brain. Cambridge,
MA: MIT Press.
Engel, B. T., & J. A. Joseph. (1981). Modulation of baroceptor sensitivity during
operant cardiac conditioning. Advances in Physiological Sciences 17:177-180.
Erikson, J. R., & M. Reiss Jones. (1978). Thinking. Annual Review of Psychology
29:61-90.
Estes, W. K. (1979). Experimental psychology: An overview. In E. Hearst (Ed.).
The first century of experimental psychology (pp. 623-667). Hillsdale, NJ:
Erlbaum.
References 293

Estes, W. K. (1984). Human learning and memory. In P. Marler, & H. S. Terrace


(Eds.), The biology of learning (pp. 616-628). Berlin-Heidelberg-New York:
Springer-Verlag.
Evarts, E. V. (1973). Motor cortex reflexes associated with learned movement.
Science 179:501-503.
Evarts, E. V., Y. Shinoda, & S. P. Wise. (1984). Neurophysiological approaches
to higher brain function. New York: Wiley (lnterscience).
Eysenck, H. J. (1971). The IQ argument: Race, intelligence and education. New
York: Library Press.
Eysenck, H., & G. Wilson. (1973). The experimental study of Freudian theories.
London: Methuen.
Falmagne, R. J. (Ed.). (1975). Reasoning: Representation and processes. Hills-
dale, NJ: Erlbaum.
Farah, M. J., M. S. Gazzaniga, J. Holtzmann, & S. M. Kosslyn. (1985). A
left hemisphere basis for visual mental imagery? Neuropsychologia 23: 115-
118.
Fernandez Guardiola, A. (Ed.). (1979). La conciencia. Mexico, D. F.: Trillas.
Feyerabend, P. K. (1975). Against method. Reprint: London: Verso. 1978.
Fishbein, E., M. Deri, M. Sainati Nello, & M. Sciolis Marino. (1985). The role of
implicit models in solving verbal problems in multiplication and division. Jour-
nal of Research in Mathematics Education 16:3-17.
Fisher, S., & R. P. Greenberg. (1977). The scientific credibility of Freud's theories
and therapy. New York: Basic Books.
Flohr, H., & W. Precht (Eds.). (1981). Lesion-induced neuronal plasticity in
sensorimotor systems. Berlin-Heidelberg-New York: Springer-Verlag.
Fodor, J. A. (1975). The language of thought. New York: Crowell.
Fodor, J. A. (1981). The mind-body problem. Scientific American 244(1): 114-123.
Fodor, J. A. (1983). The modularity of mind. Cambridge, MA: MIT Press.
Fox, B. H., & B. H. Newberry (Eds.). (1984). Impact of psychoendocrine sys-
tems in cancer and immunity. Lewiston, NY-Toronto: Hogrefe.
Freud, s. (1929). Introductory lectures on psychoanalysis (2nd ed.). London:
Allen & Unwin.
Freud, S. (1953-1965). Standard edition of the complete psychological works.
London: Hogarth.
Freud, S. (1962). The ego and the id. London: Hogarth Press & Institute of
Psycho-Analysis.
Fuster, J. M. (1984a). Behavioral electrophysiology of the prefrontal cortex.
Trends in NeuroSciences 7:408-414.
Fuster, J. M. (1984b). The cortical substrate of memory. In L. R. Squire and N.
Butters (Eds.), Neuropsychology of memory (pp. 279-286). New York:
Guilford Press.
Fuster, J. M., & G. E. Alexander. (1970). Delayed response deficit by cryogenic
depression of frontal cortex. Brain Research 20:85-90.
Gallup, G. C. (1977). Self-recognition in primates: A comparative approach to the
bi-directional properties of consciousness. American Psychologist 32:329-338.
Garcia, J. (1981). The logic and limits of mental aptitUde testing. American Psy-
chologist 36: 1172-1180.
Garcia, J., & R. A. Koelling. (1966). Relation of cue to consequence in avoidance
learning. Psychonomic Science 4: 123-124.
294 References

Gazzaniga, M. S. (1967). The split brain in man. Scientific American 217(2):24-


28.
Gentner, D., & J. Grudin. (1985). The evolution of mental metaphors in psychol-
ogy: A 90-year retrospective. American Psychologist 40: 181-182.
Gerstein, G., A. Aertsen, M. Bloom, E. Espinosa, S. Evanczuk, & M. Turner.
(1985). Multi-neuron experiments: Observation of state in neural nets. In H.
Haken (Ed.), Complex-system operational approaches (pp. 58-70). Berlin-Hei-
delberg-New York: Springer-Verlag.
Geschwind, N. (1965). Disconnexion syndromes in animals and man. Brain
88:237-294 and 585-644.
Geschwind, N. (1970). The organization of language and the brain. Science
70:940-944.
Geschwind, N. (1974). Selected papers on language and the brain. Dordrecht-
Boston: Reidel.
Geschwind, N., & A. M. Galaburda. (1985). Cerebrallateralization. Archives of
Neurology 42:428-459, 521-552, 634-654.
Gibson, J. J. (1950). The perception of the visual world. Boston: Houghton Mif-
flin.
Goddard, G. V. (1980). Component properties of the memory machine: Hebb
revisited. In D. W. Juscyk & R. M. Klein (Eds.), The nature of thought (pp.
231-247). Hillsdale, NJ: Erlbaum.
Goldberg, G. (1985). Supplementary motor area structure and function: Review
and hypotheses. The Behavioral and Brain Sciences 8:567-616.
Goldman, P. S. (1971). Functional development of the prefrontal cortex in early
life and the problem of neuronal plasticity. Experimental Neurology 32:366-
387.
Goldman-Rakic, P. S. (1982). Organization offrontal association cortex of normal
and experimentally brain-injured primates. In M. Arbib, D. Caplan, & J. C.
Marshall (Eds.), Neural models of language (pp. 469-483). New York: Aca-
demic Press.
Goody, J. (1977). The domestication of the savage man. London: Cambridge
University Press.
Gould, J. L. (1986). The locale map of honey bees: Do insects have cognitive
maps? Science 232:861-863.
Gould, J. L., & P. Marler. (1984). Ethology and the natural history of learning. In
P. Marler & H. S. Terrace (Eds.), The biology of learning (pp. 47-74). Berlin-
Heidelberg-New York: Springer-Verlag.
Gould, S. J. (1981). The mismeasure of man. New York: Norton.
Graham-Brown, T. (1914). The intrinsic factors in the act of progression in the
mammal. Proceedings of the Royal Society 84:308-319.
Greenough, W. T. (1984). Structural correlates of information storage in the
mammalian brain: A review and hypothesis. Trends in NeuroSciences 7:229-
233.
Gregory, R. L. (1973). The confounded eye. In R. L. Gregory & E. H. Gombrich
(Eds.), Illusion in Nature and Art (pp. 49-95). New York: Scribner's Sons.
Griffin, D. R. (1984). Animal thinking. Cambridge, MA: Harvard University
Press.
Griinbaum, A. (1984). Thefoundations of psychoanalysis. Berkeley, CA: Univer-
sity of California Press.
References 295

Halgren, E. (1982). Mental phenomena induced by stimulation in the limbic sys-


tem. Human Neurobiology 1:251-260.
Hansel, E. E. M. (1980). ESP and parapsychology: A critical re-evaluation. Buf-
falo, NY: Prometheus.
Harlow, H. F. (1958). The nature oflove. The American Psychologist 13:673-685.
Hart, J., Jr., R. S. Berndt, & A. Caramazza. (1985). Category-specific naming
deficit following cerebral infarction. Nature 316:439-440.
Haugeland, J. (Ed.). (1981). Mind design. Cambridge, MA: MIT Press.
Hearst, E. (Ed.). (1979). The first century of experimental psychology. Hillsdale,
NJ: Erlbaum.
Hebb, D. O. (1949). The organization of behavior. New York: Wiley.
Hebb, D. O. (1953). On motivation and thought. (Reprinted in Hebb [1982] pp. 17-
21.)
Hebb, D. O. (1961). The role of experience. (Reprinted in Hebb [1982], pp. 89-
95.)
Hebb, D. O. (1963). The semiautonomous process: Its nature and nurture. Ameri-
can Psychologist 18: 16-26.
Hebb, D. O. (1972). Textbook of psychology (3rd ed.). Philadelphia: Saunders.
Hebb, D. O. (1980). Essay on mind. Hillsdale, NJ: Erlbaum.
Hebb, D. O. (1982). The conceptual nervous system. H. A. Buchtel, Ed. Oxford
and New York: Pergamon.
Hebb, D.O., W. E. Lambert, & G. R. Tucker. (1971). Language, thought and
experience. The Modern Language Journal 55:212-222.
Heilman, K. M., & P. Satz (Eds.). (1983). Neuropsychology of human emotion.
New York: Guilford Press.
Herrnstein, R. J. (1984). Objects, categories, and discriminative stimuli. In H. C.
Roitblat, T. G. Bever, & H. S. Terrace (Eds.), Animal cognition. Hillsdale, NJ:
Erlbaum.
Hess, W. R. (1957). The functional organization of the diencephalon. J. R.
Hughes, Ed. New York: Grune and Stratton.
Hilgard, E. R. (1977). Divided consciousness: Multiple controls in human thought
and action. New York: Wiley.
Hilgard, E. R. (1980). Consciousness in contemporary psychology. Annual Re-
view of Psychology 31:1-26.
Hirsh, R. (1974). The hippocampus and contextual retrieval of information from
memory: A theory. Behavioral Biology 12:421-444.
Hobson, J. A., R. Lydic, & H. A. Baghdoyan. (1986). Evolving concepts of sleep
cycle generation: From brain centers to neuronal populations. The Behavioral
and Brain Sciences 9:371-400.
Hoffmann, B. (1962). The tyranny of testing. New York: Crowell.
Hofstadter, D. R., & D. C. Dennett (Eds.). (1981). The mind's I. Fantasies and
reflections on self & soul. New York: Basic Books.
Hollard, V. D., & J. D. Delius. (1982). Rotational invariance in visual pattern
recognition by pigeons and humans. Science 218:804-806.
Homans, G. C. (1974). Social behavior: Its elementary forms (Rev. ed.). Orlando,
FL: Harcourt Brace Jovanovich.
Hoyle, G. (1976). Approaches to understanding the neurophysiological basis of
behavior. In J. C. Fentress (Ed.), Simpler networks and behavior (pp. 21-38).
Sunderland, MA: Sinauer Associates.
296 References

Huarte de San Juan, J. (1976). Examen de ingenios para las ciencias. Madrid:
Editora Nacional. (Original work published 1575)
Hubel, D. H. (1982). Exploration of the primary visual cortex, 1955-1978. Nature
299:515-524.
Hubel, D. H., & T. N. Wiesel. (1962). Receptive fields, binocular interaction, and
functional architecture in the cat's visual cortex. Journal of Physiology
160:106-154.
Hubel, D. H., & T. N. Wiesel. (1968). Receptive fields and functional architecture
of monkey striate cortex. Journal of Physiology 195:215-243.
Hull, C. L. (1952). A behavior system. New Haven, CT: Yale University
Press.
Humphrey, N. (1983). Consciousness regained. Oxford and New York: Oxford
University Press.
Huxley, A. (1932). Brave new world. New York: Harper.
Ingvar, D. H. (1979). "Hyperfrontal" distribution of regional cerebral blood flow
in resting wakefulness: On the functional anatomy of the conscious state. Acta
Neurologica Scandinavica 60: 12-25.
James, W. (1950). Principles of Psychology (2 vols). New York: Dover. (Original
work published 1890)
Jarochewski, M. (1975). Psychologie im 20. Jahrhundert. Berlin: Yolk und Wis-
sen.
Jasper, H. H., & G. Bertrand. (1966). Thalamic units involved in somatic sensa-
tion and voluntary and involuntary movements in man. In D. P. Purpura & M.
D. Yahr (Eds.), The thalamus (pp. 365-390). New York: Columbia University
Press.
Jerison, H. J. (1973). Evolution of the brain and intelligence. New York: Aca-
demic Press.
John, E. R., Y. Tang, A. B. Brill, R. Young, & K. Ono (1986). Doubled-labeled
metabolic maps of memory. Science 233: 1167-1175.
Johnson, R. D., & C. H. Jones. (1984). Attitudes toward the existence and scien-
tific investigation of extrasensory perception. The Journal of Psychology
117:19-22.
Johnson-Laird, P. N. (1983). Mental models. Cambridge, MA: Harvard Univer-
sity Press.
Johnson-Laird, P. N., & P. C. Wason (Eds.). (1977). Thinking. Readings in cogni-
tive science. Cambridge, U.K.: Cambridge University Press.
Kamin, L. J. (1974). The Science and politics of IQ. New York: Wiley.
Kandel, E. R. (1976). Cellular basis of behavior. San Francisco: Freeman.
Kandel, E. R., & J. H. Schwartz (Eds.). (1981). Principles of neural science. New
York-Amsterdam: Elsevier.
Keeser, W., & M. Bullinger. (1984). Process-oriented evaluation of a cognitive-
behavioural treatment for clinical pain: A time-series approach. In B. Bromm
(Ed.), Pain measurement in man. Neurophysiological correlates of pain (pp.
417-428). Amsterdam and New York: Elsevier.
Kirton, M. J. (1976). Adaptors and innovators: A description and measure. Jour-
nal of Applied Psychology 61:622-629.
Kmetz, J. M. (1978). Plant primary perception: The other side of the leaf. The
Skeptical Inquirer 2(2):57-61.
Knapp, A. G., & J. A. Anderson. (1974). Theory of categorization based on
distributed memory storage. Journal of Experimental Psychology 10:616-637.
References 297

Koffka, K. (1935). Principles of Gestalt psychology. Orlando, FL: Harcourt


Brace Jovanovich.
Kohler, W. (1929). Gestalt psychology. New York: Liveright.
Krech, D. (1950). Dynamic systems, psychological fields, and hypothetical con-
structs. Psychological Review 57:283-290.
Krechevsky, I. (1932). "Hypotheses" versus "chance" in the pre-solution period
in sensory discrimination learning. (University of California Publications in
Psychology, Vol. 6, No.3). Los Angeles, CA: University of California Press.
Kripke, S. (1971). Identity and necessity. In M. K. Munitz (Ed.), Identity and
individuation (pp. 135-164). New York: New York University Press.
Kuhn, T. S. (1962). The structure of scientific revolutions. Chicago: University of
Chicago Press.
Lagerspetz, K. (1981). Combining aggression studies in infra-humans and man. In
P. F. Brain & D. Benton (Eds.), Multidisciplinary approaches to aggression
research (pp. 389-400). Amsterdam and New York: Elsevier.
Lagerspetz, K. M. J., & P. Niemi (Eds.). (1984). Psychology in the 1990s. Am-
sterdam and New York: Elsevier/North-Holland.
Lamendella, J. T. (1977). General principles of neurofunctional organization and
their manifestations in primary and nonprimary language acquisition. Lan-
guage Learning 27:155-196.
Laming, D. (1985). Some principles of sensory analysis. Psychological Review
92:462-485.
Land, E. H. (1983). Recent advances in retinex theory and some implications for
cortical computations: Color vision and the natural image. Proceedings of the
National Academy of Sciences of the USA 80:5163-5169.
Larson, J., & G. Lynch. (1986). Induction of synaptic potentiation in hippocam-
pus by patterned stimulation involves two events. Science 232:985-988.
Lashley, K. S. (1929). Brain mechanisms and intelligence. Chicago: University of
Chicago Press.
Lashley, K. S. (1941). Coalescence of neurology and psychology. Proceedings of
the American Philosophical Society 84:461-470.
Layzer, D. (1974). Heritability analysis of IQ scores: Science or numerology?
Science 183:1259-1266.
LeDoux, J. E., D. H. Wilson, & M. S. Gazzaniga. (1979). Beyond commisuro-
tomy: Clues to consciousness. In M. S. Gazzaniga (Ed.), Handbook of behav-
ioral neurobiology (Vol. 2, pp. 543-554). New York and London: Plenum.
Lehrman, D. S. (1953). A crititique of Konrad Lorenz's theory of instinctive
behavior. Quarterly Review of Biology 28:337-363.
Lenin, V. I. (1947). Materialism and empirio-criticism. Moscow: Foreign Lan-
guages Publishing House. (Original work published 1908)
Levine, M. (1974). A cognitive theory of learning. Hillsdale, NJ: Erlbaum.
Libet, B. (1965). Cortical activation in conscious and unconscious experience.
Perspectives in Biology and Medicine 9:77-86.
Libet, B. (1978). Neuronal vs. subjective timing, for a conscious sensory experi-
ence. In P. A. Buser & A. Rougeul-Buser (Eds.), Cerebral correlates of con-
scious experience (pp. 69-82). Amsterdam: North-Holland.
Libet, B. (1985). Unconscious cerebral initiative and the role of conscious will in
voluntary action. The Behavioral and Brain Sciences 8:529-539.
Lieberman, P. (1984). The biology and evolution of language. Cambridge, MA:
Harvard University Press.
298 References

Lieberman, P. (1985). On the evolution of human syntactic ability. Its pre-adap-


tive bases-motor control and speech. Journal of Human Evolution 14:657-
668.
Lloyd Morgan, C. (1894). An introduction to comparative psychology. London:
Walter Scott.
Locke, S. E., & M. Hornig-Rohan (1983). Mind and immunity: Behavioral immu-
nology, an annotated bibliography. Washington, DC: Institute for the Ad-
vancement of Health.
Lombardi, C. M., C. F. Fachinelli, & J. Delius. (1984). Oddity of visual patterns
conceptualized by pigeons. Animal Learning & Behavior 12:2-6.
Luce, R. D., R. R. Bush, & E. Galanter (Eds.). (1963-1965) Handbook of mathe-
matical psychology (3 vols.). New York: Wiley.
Luria, A. R. (1973). The working brain. Hardmonsworth: Penguin Books.
Luria, A. R. (1975). Neuropsychology: Its sources, principles, and prospects. In
F. G. Worden, J. P. Swazey, & G. Adelman (Eds.), The neurosciences: Paths
of discovery. Cambridge, MA: MIT Press.
Luria, A. R. (1976). Cognitive development. Its cultural and social foundations.
Cambridge, MA: Harvard University Press.
Luria, A. R. (1979). The making of mind. A personal account of soviet psychol-
ogy. M. Cole and S. Cole, Eds. Cambridge, MA: Harvard University Press.
MacCorquodale, K., & P. E. Meehl. (1948). On a distinction between hypotheti-
cal constructs and intervening variables. Psychological Review 55:95-107.
MacKay, D. M. (l969).1nformation. mechanism. and meaning. Cambridge, MA:
MIT Press.
MacKay, D. M. (1978). Selves and brains. Neuroscience 3:599-606.
Macphail, E. M. (1982). Brain and intelligence in vertebrates. Oxford: Clarendon
Press.
Mahoney, M. J. (1976). Scientist as subject. The psychological imperative. Cam-
bridge, MA: Ballinger.
Maine de Biran. (1823-1824). Nouveaux essais d'anthropologie. In P. Tisserand
(Ed.), Oeuvres complt?tes, Vol. 14. Paris: Alcan & PUF.
Majerus, M. E. N., P. O'Donald, & J. Weir. (1982). Female mating preference is
genetic. Nature 300:521-523.
Malamut, B. L., R. C. Saunders, & M. Mishkin. (1984). Monkeys with combined
amygdalo-hippocampus lesions succeed in object discrimination learning de-
spite 24-hour intertrial periods. Behavioral Neuroscience 98:759-769.
Mandler, G. (1984). The construction and limitation of consciousness. In V.
Sarris & A. Parducci (Eds.). Perspectives in psychological experimentation:
Toward the year 2000 (pp. 109-126). Hillsdale, NJ: Eribaum.
Mandler, G., & W. Kessen. (1974). The appearance of free will. In S. C. Brown
(Ed.), Philosophy of psychology (pp. 305-324, 340-342). London: Macmillan.
Margolis, J. (1984). Philosophy of psychology. Englewood Cliffs, NJ: Prentice-
Hall.
Marks, D. F. (1986). Investigating the paranormal. Nature 320:119-124.
Marier, P., & H. S. Terrace (Eds.). (1984). The biology of learning. New York:
Springer-Veriag.
Marr, D. (1982). Vision: A computational investigation in the human representa-
tion of visual information. San Francisco: Freeman.
Marshall, J. C. (1977). Minds, machines and metaphors. Social Studies of Science
7:475-488.
References 299

Martin, U., H. Martin, & M. Lindauer. (1978). Transplantation of a time-signal in


honeybees. Journal of Comparative Physiology A 124: 193-201.
Marx, M. H., & W. A. Hillix. (1973). Systems and theories in psychology (2nd
ed.). New York: McGraw-Hill.
Masterton, R. B., C. B. G. Campbell, M. E. Bitterman, & N. Hotton (Eds.).
(1976a). Evolution of brain and behavior in vertebrates. Hillsdale, NJ:
Erlbaum.
Masterton, R. B., W. Hodos, & H. J. Jerison (Eds.). (l976b). Evolution, brain,
and behavior. Hillsdale, NJ: Erlbaum.
Matarazzo, J. D., N. E. Miller, S. M. Weiss, & J. A. Herd. (1984). Behavioral
health. New York: Wiley.
Matsuzawa, T. (1985). Use of numbers by a chimpanzee. Nature 315:57-59.
Matthews, L. J., & J. H. Patton. (1975). Failure to shift following disconfirmation
in concept identification. Journal of Experimental Psychology of Human
Learning and Memory 1:91-94.
Maudsley, H. (1876). The physiology of mind. London: Macmillan.
Mayer, R. E. (1977). Thinking and problem-solving. Glenview IL: Scott, Fores-
man.
Mayes, A. (Ed.). (1983). Memory in animals and humans. Wokingham, U.K.:
Van Nostrand Reinhold.
McClelland, D. C., & R. A. Clark. (1953). Discrepancy hypothesis. In D. C.
McClelland, J. W. Atkinson, R. A. Clark, & E. L. Lowell (Eds.), The achieve-
ment motive (pp. 42-66). New York: Appleton-Century-Crofts.
McCloskey, M., A. Caramazza, & B. Green. (1980). Curvilinear motion in the
absence offorces: Naive beliefs about the motion of objects. Science 210: 1139-
1141.
McCulloch, W. S. (1965). Embodiments of mind. Cambridge, MA: MIT Press.
McGuigan, F. J. (1978). Cognlllve psychopnysiology: Principles of covert behav-
ior. Englewood Cliffs, NJ: Prentice-Hall.
McKenna, F. P. (1985). Another look at the "new psychophysics." British Jour-
nal of Psychology 76:97-109.
McNaughton, B. L., R. M. Douglas, & G. V. Goddard. (1978). Synaptic enhance-
ment in fascia dentata: Cooperativity among coactive afferents. Brain Research
157:277-293.
Meehl, P. E. (1978). On soft and hard psychology. Journal of Consulting and
Clinical Psychology 46:806-834.
Melvill Jones, G. (1977). Plasticity in the adult vestibulo-ocular reflex arc. Philo-
sophical Transactions of the Royal Society of London B 278:319-334.
Melvill Jones, G., A. Berthoz, & B. Segal. (1984). Adaptive modification of the
vestibulo-ocular reflex by mental effort in darkness. Experimental Brain Re-
search 56:149-153.
Merzenich, M. M., R. J. Nelson, M. P. Stryker, M. S. Cynader, A. Schoppmann,
& J. M. Zook. (1984). Somatosensory cortical map changes following digit
amputation in adult monkeys. Journal of Comparative Neurology 224: 591-
605.
Metzinger, T. (1985). Neuere Beitriige zur Diskussion des Leib-Seele-Problems.
Frankfurt: Peter Lang.
Miller, G. A. (1956). The magical number seven, plus or minus two: Some limits
on our capacity for processing information. Psychological Review 63:81-97.
Miller, G. A. (1964). Mathematics and psychology. New York: Wiley.
300 References

Miller, G. A. (1980). Computation, consciousness and cognition. Behavioral and


Brain Sciences 3: 146.
Milner, B. (1959). The memory deficit in bilateral hippocampal lesions. Psychiat-
ric Research Reports 11 :43-58.
Milner, B. (1982). Some cognitive effects of frontal-lobe lesions in man. Philo-
sophical Transactions of the Royal Society of London. B 298:211-226.
Milner, B., & M. Petrides. (1984). Behavioural effects of frontal-lobe lesions in
man. Trends in NeuroSciences 7:403-407.
Milner, P. M. (1957). The cell assembly: Mark II. Psychological Review 64:242-
252.
Milner, P. M. (1970). Physiological psychology. New York: Holt, Rinehart and
Winston.
Mimura, K. (1986). Development of visual pattern discrimination in the fly de-
pends on light experience. Science 232:83-85.
Mishkin, M. (1982). A memory system in the monkey. Philosophical Transactions
of the Royal Society of London B 298:85-95.
Mishkin, M., B. Malamut, & J. Bachevalier. (1984). Memories and habits:
Two neural systems. In G. Lynch, J. L. McGaugh, & N. M. Weinberger
(Eds.), Neurobiology of learning and memory (pp. 65-77). New York: Guilford
Press.
Mishkin, M., & H. L. Petri. (1984). Memories and habits: Some implications for
the analysis of learning and retention. In L. R. Squire & N. Butters (Eds.),
Neuropsychology of memory (pp. 287-294). New York: Guilford Press.
Mishkin, M., B. J. Spiegler, R. C. Saunders, & B. L. Malamut. (1982). An animal
model of global amnesia. In S. Corkin, K. L. Davis, J. H. Growden, E. Usdin,
& R. J. Wurtman (Eds.), Alzheimer's disease: A report of progress. New York:
Raven Press.
Moore, J., & A. Newell. (1974). How can Merlin understand? In L. W. Gregg
(Ed.), Cognition and knowledge. Hillside, NJ: Erlbaum.
Morris, R. G. M., E. Anderson, G. S. Lynch, & M. Baudry. (1986). Selective
impairment oflearning and blockade oflong-term potentiation by an N-methyl-
D-aspartate receptor antagonist, AP5. Nature 319:774-776.
Motley, M. T. (1985). Slips of the tongue. Scientific American 253(3): 116-127.
Mountcastle, V. B. (1978). An organizing principle for cerebral function. In G. M.
Edelman & V. B. Mountcastle, The mindful brain (pp. 1-49). Cambridge, MA:
MIT Press.
Mountcastle, V. B., R. A. Andersen, & B. C. Motter. (1981). The influence of
attentive fixation upon the excitability of the light-sensitive neurons of the
posterior parietal cortex. Journal of Neuroscience 1:1218-1235.
Mountcastle, V. B., J. C. Lynch, A. Georgopoulos, H. Sakata, & C. Acuna.
(1975). Posterior parietal association cortex of the monkey. Journal of Neuro-
physiology 38:871-908.
Munn, C. A. (1986). Birds that "cry wolf." Nature 319:143-145.
Murray, E. A., & M. Mishkin. (1985). Amygdalectomy impairs cross modal asso-
ciation in monkeys. Science 228:604-606.
Nathans, J., T. P. Piantanida, R. L. Eddy, T. B. Shows, & D. S. Hodges. (1986).
Molecular genetics of inherited variation in human color vision. Science
232:203-210. •
Neisser, U. (1976). Cognition and reality. San Francisco: Freeman.
References 301

Neisser, U. (1982a). Memory: What are the important questions? In U. Neisser


(Ed.), Memory observed: Remembering in natural contexts (pp. 3-19). San
Francisco: Freeman.
Neisser, U. (Ed.). (1982b). Memory observed. Remembering in natural contexts.
San Francisco: Freeman.
Newell, A. (1982). The knowledge level. Artificial Intelligence 18:87-127.
Newell, A., & H. A. Simon. (1963). Computers in psychology. In R. D. Luce, R.
R. Bush, & E. Galanter (Eds.), Handbook of mathematical psychology (Vol. I,
pp. 361-428). New York: Wiley.
Newell, A., & H. A. Simon. (1972). Human problem solving. Englewood Cliffs,
NJ: Prentice-Hall.
Newell, A., & H. A. Simon. (1981). Computer science as empirical inquiry. In J.
Haugeland (Ed.), Mind design (pp. 35-66). Cambridge, MA: MIT Press.
Nottebohm, F. (1981). A brain for all seasons: Cyclical anatomical changes in
song control nuclei of the canary brain. Science 214:1368-1370.
Oakley, D. A. (1983). The varieties of memory: A phylogenetic approach. In
Mayes (Ed.), pp. 20-82.
Oakley, D. A. (Ed.). (1985). Brain and mind. London: Methuen.
Oatley, K. (1980). Representing ourselves: Mental schemata, computational met-
aphors, and the nature of consciousness. In G. Underwood & R. Stevens
(Eds.), Aspects of consciousness (Vol. 2). London: Academic.
O'Keefe, J., & L. Nadel. (1978). The hippocampus as a cognitive map. Oxford:
Clarendon Press.
Olds, J. (1975). Mapping the mind onto the brain. In F. G. Worden, J. P. Swazey,
& G. Adelman (Eds.). The neurosciences: Paths of discovery (pp. 375-400).
Cambridge, MA: MIT Press.
Ornstein, R. E. (Ed.). (1973). The nature of human consciousness. San Francisco:
Freeman.
Orwell, G. (1949). 1984. New York: Harcourt Brace Jovanovich.
Osgood, C. E. (1953). Method and theory in experimental psychology. New York:
Oxford University Press.
Paillard, J., F. Michel, & C. E. Stelmach. (1983). Localization without content: A
tactile analogue of "blindsight." Archives of Neurology 40:548-551.
Paivio, A. (1975). Neomentalism. Canadian Journal of Psychology 29:263-291.
Palm, G. (1981). Towards a theory of cell assemblies. Biological Cybernetics
39:181-194.
Paradis, M. (1985). On the representation of two languages in one brain. Lan-
guage Sciences 7:1-39.
Paradis, M. (Ed.). (1983). Readings on aphasia in bilinguals and polyglots. Mon-
treal: Didier.
Parducci, A., & V. Sarris. (1984). The experimental approach: Dead end or via
regia? In V. Sarris & A. Parducci (Eds.), Perspectives in psychological experi-
mentation: Toward the year 2000 (pp. 1-14). Hillsdale, NJ: Erlbaum.
Patry, J.-L. (Ed.). (1982). Feldforschung. Methoden und Probleme sozialwissens-
chaftlicher Forschung unter natiirliche Bedingungen. Bern, Stuttgart, Wien:
Hans Huber.
Paunonen, S. V., & J. N. Jackson. (1985). Idiographic measurement strategies for
personality and prediction: Some unredeemed promissory notes. Psychological
Review 92:486-511.
302 References

Pavlov, I. P. (1927). Conditioned reflexes. Oxford: Oxford University Press.


Pavlov, I. P. (1955). Selected works. Moscow: Foreign Languages Publishing
House.
Pears, D. (1975). Questions in the philosophy of mind. London: Duckworth.
Penfield, W. (1975). The mystery of the mind. Princeton, NJ: Princeton University
Press.
Penfield, W., & P. Perot. (1963). The brain's record of auditory and visual experi-
ence: A final summary and discussion. Brain 86:595-696.
Perrez, M. (1979). 1st die Psychoanalyse eine Wissenschaft? (2nd ed.). Bern,
Stuttgart, Wien: Hans Huber.
Petermann, B. (1932). The Gestalt theory and the problem of configuration. Lon-
don: Routledge & Kegan Paul.
Petrides, M. (1982). Motor conditional associative-learning after selective pre-
frontal lesions in the monkey. Behavioral and Brain Research 5:407-413.
Petrides, M. & S. D. Iversen. (1979). Restricted posterior parietal lesions in the
rhesus monkey and performance on visuospatial tasks. Brain Research 161 :63-
77.
Petrides, M., & B. Milner. (1982). Deficits on subject-ordered tasks after frontal-
and temporal-lobe lesions in man. Neuropsychologia 20:249-262.
Piaget, J. (1954). The construction of reality in the child. New York: Basic
Books.
Piaget, J. (1964). Six etudes de psychologie. Geneve: Gonthier.
Piaget, J. (1965). Etudes sociologiques. Geneve: Librairie Droz.
Piaget, J. (1971). Insights and illusions of philosophy. London: Routledge &
Kegan Paul.
Piaget, J. (1976). Le comportement, moteur de {'evolution. Paris: Gallimard.
Plum, F., & J. B. Posner (1980). The diagnosis of stupor and coma (3rd ed).
Philadelphia: Davis.
Poppel, E. (1977). Introduction: Relating perceptual phenomena to neuronal
mechanisms. Neurosciences Research Program Bulletin 15:323-326.
Poppel, E. (1985). Grenzen des Bewusstseins. Stuttgart: Deutsche Verlags-An-
stalt.
Poppel, E., R. Held, & D. Frost. (1973). Residual visual function after brain
wounds involving the central visual pathways in man. Nature 243:295-
296.
Popper, K. R. (1959). The logic of scientific discovery. London: Hutchinson.
Popper, K. R. (1972). Objective knowledge. Oxford: Clarendon Press.
Popper, K. R., & J. C. Eccles. (1977). The self and its brain. New York: Springer
International.
Pratt, C. C. (1939). The logic of modern psychology. New York: Macmillan.
Pribram, K. H. (1971). Languages of the brain: Experimental paradoxes and
principles in neuropsychology. Englewood Cliffs, NJ: Prentice-Hall.
Prioleau, L., M. Murdock, & N. Brody. (1983). An analysis of psychotherapy vs.
placebo studies. Behavioral and Brain Sciences 6:275-310.
Purves, D., & R. D. Hadley. (1985). Changes in the dendritic branching of adult
mammalian neurones revealed by repeated imaging in situ. Nature 315:404-
406.
Putnam, H. (1960). Minds and machines. In S. Hook (Ed.), Dimensions of mind
(pp. 148-179). New York: New York University Press.
References 303

Putnam, H. (1975). Mind, language, and reality. Cambridge, U.K.: Cambridge


University Press.
Pylyshyn, Z. W. (1978). Computational models and empirical constraints. Behav-
ioral and Brain Sciences 1:93-99.
Pylyshyn, Z. W. (1980). Computation and cognition: Issues in the foundation of
cognitive science. Behavioral and Brain Sciences 3: 111-132.
Pylyshyn, Z. W. (1984). Computation and cognition. Cambridge, MA: MIT Press.
Quinn, W. G. (1984). Work in invertebrates on the mechanisms underlying learn-
ing. In P. Marler & H. S. Terrace (Eds.), The biology of learning (pp. 197-246).
Berlin-Heidelberg-New York: Springer-Verlag.
Rachman, S. (Ed.). (1963). Critical essays on psychoanalysis. New York: Mac-
millan.
Rager, G. (1981). The significance of neuronal cell death during the development
of the nervous system. In H. Flohr & W. Precht (Eds.), Lesion-induced neuro-
nal plasticity in sensorimotor systems (pp. 3-12). Berlin-Heidelberg-New
York: Springer-Verlag.
Rakic, P. (1985). Limits of neurogenesis in primates. Science 227:1054-1056.
Ramachandran, V. S., & S. M. Anstis. (1983). Perceptual organization in moving
patterns. Nature 304:529-531.
Ramachandran, V. S., & S. M. Anstis. (1986). The perception of apparent mo-
tion. Scientific American 254(2):102-109.
Rasmussen, T., & B. Milner. (1977). The role of early left-brain injury in deter-
mining lateralization of cerebral speech functions. Annals of the New York
Academy of Sciences 229:355-369.
Reed, T. E. (1983). Nerve conduction in mice: A new method with results and
analysis of variation. Behavioral Genetics 13:257-265.
Reed, T. E. (1984). Residual latency (delay at the neuromuscular junction): Nor-
mative values and heritability in mice. Behavioral Genetics 14:209-219.
Restle, F. (1976). The selection of strategies in cue learning. Psychological Re-
view 69:329-343.
Robinson, D. N. (1985). Philosophy of Psychology. New York: Columbia Univer-
sity Press.
Roe, A., & G. G. Simpson (Eds.). (1958). Behavior and evolution. New Haven,
CT: Yale University Press.
Roth, J., D. LeRoith, J. Shiloach, J. L. Rosenzweig, M. A. Lesniak, & J.
Havrankova. (1982). The evolutionary origins of hormones, neurotransmitters,
and other extracellular chemical messengers. New England Journal of Medi-
cine 306:523-527.
Rusiniak, K. W., C. P. Palmeri no , A. G. Rice, D. L. Forthman, & J. Garcia.
(1982). Flavor-illness aversions: Potentiation of odor by taste with toxin but
not shock in rats. Journal of Comparative Physiological Psychology 94:527-
539.
Russell, E: S. (1976). Report of the ad hoc committee. Genetics 83:s99-slO1.
Ryle, G. (1949). The concept of mind. London: Hutchinson.
Ryle, G. (1960). Dilemmas. Cambridge: University Press.
Sagi, D., & B. Julesz. (1986). Enhanced detection in the aperture of focal atten-
tion during simple discrimination tasks. Nature 321:693-695.
Sarris, V. (1986). Lehrbuch der experimentalen Psychologie: Methodologische
Grundlagen (2 vols.). Munchen: Reinhardt (UTB).
304 References

Sarris, V., & A. Parducci (Eds.). (1984). Perspectives in psychological experi-


mentation: Toward the year 2000. Hillsdale, NJ: Erlbaum.
Saunders, R. C., E. A. Murray, & M. Mishkin. (1984). Further evidence that
amygdala and hippocampus contribute equally to recognition memory. Neurop-
sychologia 22 :785-796.
Schacter, D. L. (1983). Amnesia observed: Remembering and forgetting in a
natural environment. Journal of Abnormal Psychology 92:236-242.
Schacter, D. L. (1985). MUltiple forms of memory in humans and animals. In N.
M. Weinberger, J. L. McGaugh, & G. Lynch (Eds.), Memory systems of the
brain: Animal and human cognitive processes. New York: Guilford Press.
Schacter, D. L., J. L. Harbluk, & D. R. McLachlan. (1984). Retrieval without
recollection: An experimental analysis of source amnesia. Journal of Verbal
Learning and Verbal Behavior 23:593-61l.
Schmidt, H. (1975). Toward a mathematical theory of psi. Journal of the Ameri-
can Society for Psychical Research 69:301-319.
Schmitt, F. 0., F. G. Worden, G. Adelman, & S. G. Dennis (Eds.). (1981). The
organization of the cerebral cortex. Cambridge, MA: MIT Press.
Schwartz, N. B. (1984). Endocrinology as a paradigm, endocrinology as author-
ity. Endocrinology 114:308-313.
Scoville, W. B., & B. Milner. (1957). Loss of recent memory after bilateral
hippocampal lesions. Journal of Neurology, Neurosurgery and Psychiatry
20: 11-2l.
Scribner, S. (1975). Recall of classical syllogisms: A cross-cultural investigation
of error on logical problems. In R. J. Falmagne (Ed.), Reasoning: Representa-
tion and processes (pp. 153-173). Hillsdale, NJ: Erlbaum.
Scribner, S., & M. Cole. (1981). The psychology of literacy. Cambridge, MA:
Harvard University Press.
Searle, J. R. (1983). Intentionality: An essay in the philosophy of mind. Cam-
bridge, U.K.: Cambridge University Press.
Segall, M., D. T. Campbell, & M. J. Herskovits. (1966). The influence of culture
on visual perception. Indianapolis: Bobbs-Merrill.
Shallice, T. (1972). Dual functions of consciousness. Psychological Review
79:383-393.
Shallice, T. (1979). Case study approach in neuropsychological research. Journal
of Clinical Neuropsychology 1:183-211.
Shallice, T. (1982). Specific impairments of planning. Philosophical Transactions
of the Royal Society of London B 298: 199-209.
Shallice, T., E. K. Warrington, & R. McCarthy. (1983). Reading without seman-
tics. Quarterly Journal of Experimental Psychology 35A: 111-138.
Sharpless, S. (1964). Reorganization of function in the nervous system-use and
disuse. Annual Review of Physiology 26:357-388.
Simon, H. A. (1979). Information-processing models of cognition. Annual Review
of Psychology 30:363-396.
Simon, H. A. (1980). The behavioral and social sciences. Science 209:72-
78.
Singer, W. (1982). Central core control of developmental plasticity in the kitten
visual cortex: I. Diencephalic lesions. Experimental Brain Research 47:209-
222.
Skinner, B. F. (1938). The behavior of organisms. New York: Appleton-Century-
Crofts.
References 305

Skinner, B. F. (1945). The operational analysis of psychological terms. Psycho-


logical Review 52:270-277.
Skinner, B. F. (1948). Walden two. New York: Macmillan.
Skinner, B. F. (1950). Are theories of learning necessary? Psychological Review
57: 193-216.
Skinner, B. F. (1953). Science and human behavior. New York: Macmillan.
Skinner, B. F. (1969). Contingencies of reinforcement. New York: Appleton-
Century-Crofts.
Skinner, B. F. (1971). Beyondfreedom and dignity. New York: Knopf.
Skinner, B. F. (1978). Reflexions on behaviorism and society. Englewood Cliffs,
NJ: Prentice-Hall.
Sloman, A. (1978a). The computer revolution in philosophy: Philosophy, science,
and models of mind. Atlantic Highlands, NJ: Humanities Press.
Sloman, A. (1978b). What about their internal languages? Behavioral and Brain
Sciences 1:602-603.
Smith Churchland, P. (1980). A perspective on mind-brain research. Journal of
Philosophy 77: 185-207.
Smith Churchland, P. (1986). Neurophilosophy. Cambridge, MA: MIT Press.
Spence, K. W. (1948). The postulates and methods of "behaviorism." Psycholog-
ical Review 55:67-78.
Squire, L. R. (1986). Mechanisms of memory. Science 232: 1612-1619.
Squire, L., & N. J. Cohen. (1985). Human memory and amnesia. In J. L.
McGaugh, G. Lynch, & N. M. Weinberger (Eds.), The neurobiology oflearn-
ing and memory (pp. 3-64). New York: Guilford Press.
Squire, L. R., N. J. Cohen, & J. A. Zouzounis. (1984). Preserved memory in
retrograde amnesia: Sparing of a recently acquired skill. Neuropsychologia
22:145-152.
Sternberg, R. J. (1985). Beyond IQ. New York: Cambridge University Press.
Stevens, S. S. (1935). The operational definition of psychological concepts. Psy-
chological Review 42:517-527.
Stoerig, P., M. Hiibner, & E. Poppel. (1985). Signal detection analysis of residual
vision. Neuropsychologia 23:589-599.
Stone, G., N. Adler, & F. Cohen. (1979). Health psychology: A handbook. San
Francisco: Jossey-Bass.
Strawson, P. F. (1959). Individuals. London: Methuen.
Suppes, P. (1975). From behaviorism to neobehaviorism. Theory & Decision
6:269-285.
Suppes, P. (Ed.). (1978). Impact of research on education: Some case studies.
Washington, DC: National Academy of Education.
Swash, M., & c. Kennard (Eds.). (1985). Scientific basis of clinical neurology.
Edinburgh: Churchill Livingstone.
Taylor, C. (1964). The explanation of behaviour. London: Routledge & Kegan
Paul.
Teuber, H.-L. (1959). Some alterations in behavior after cerebral lesions in man.
In Evolution of nervous controlfrom primitive organisms to man (pp. 157-194).
Washington, DC: American Association for the Advancement of Science.
Teuber, H.-L. (1978). The brain and human behavior. Proceedings of the XXIst
International Congress of Psychology (pp. 119-163). Paris: PUF.
Thompson, R. F. (1975). Introduction to physiological psychology. New York:
Harper & Row.
306 References

Tolman, E. C. (1932). Purposive behavior in animals and men. New York: Cen-
tury.
Tolman, E. C., & I. Krechevsky. (1933). Means-end-readiness and hypothesis.
Psychological Review 40:60-70.
Tranel, D., & A. R. Damasio. (1985). Knowledge without awareness: An auto-
nomic index of facial recognition by prosopagnosis. Science 228: 1453-1454.
Treisman, A. (1982). Perceptual grouping and attention to visual search for fea-
tures and for objects. Journal of Experimental Psychology 8:194-214.
Treisman, A. M., & G. Gelade. (1980). A feature-integration theory of attention.
Cognitive Psychology 12:97-136.
Treisman, A., & R. Paterson. (1984). Emergent features, attention, and object
perception. Journal of Experimental Psychology 10: 12-31.
Triandis, H. C. (Ed.). (1980). Handbook of cross-cultural psychology (6 vols.).
Boston: Allyn & Bacon.
Triandis, H. C., V. Vassiliou, G. Vassiliou, Y. Tanaka, & A. V. Shanmugam.
(1972). The analysis of subjective culture. New York: Wiley.
Tulving, E. (1983). Elements of episodic memory. Oxford: Clarendon Press.
Tulving, E. (1984). Multiple learning and memory systems. In K. M. J. Lager-
spetz & P. Niemi (Eds.). Psychology in the 1990s (pp. 163-184). Amsterdam
and New York: ElsevierlNorth-Holland.
Tulving, E. (1985a). Memory and consciousness. Canadian Psychology 26: 1-12.
Tulving, E. (1985b). How many memory systems are there? American Psycholo-
gist 40:385-398.
Tunnell, G. B. (1977). Three dimensions of naturalness: An expanded definition of
field research. Psychological Bulletin 84:426-437.
Tuomela, R. (1973). Theoretical concepts. Wien-New York: Springer-Verlag.
Turing, A. (1950). Can a machine think? Mind NS 59:433-460.
Tversky, A., & D. Kahneman. (1971). Belief in the law of small numbers. Psycho-
logical Bulletin 76:105-110.
Uexkiill, J. von. (1921). Umwelt und Innenwelt der Tiere (2nd ed.). Berlin:
Springer-Verlag.
Ullmann, L. P., & Krasner, L. (1975). A psychological approach to abnormal
behavior (2nd. ed.). Englewood Cliffs, NJ: Prentice-Hall.
Underwood, G., & R. Stevens (Eds.). (1980). Aspects of consciousness (2 vols.).
London: Academic.
Ungerleider, L. G., & M. Mishkin. (1982). Two cortical visual systems. In D. J.
Ingle, M. A. Goodale, & R. J. W. Mansfield (Eds.), Analysis of visual behavior
(pp. 549-586). Cambridge, MA: MIT Press.
Uttal, W. R. (1978). The psychobiology of mind. Hillsdale, NJ: Erlbaum.
Valentine, E. R. (1982). Conceptual issues in psychology. London: Allen & Un-
win.
Van der Loos, H., & T. A. Woolsey. (1973). Somatosensory cortex: Structural
alterations following early injury to sense organs. Science 179:395-398.
Van Rillaer, J. (1980). Les illusions de la psychanalyse. Brussels: Pierre Mardaga.
Varela, J. (1971). Psychological solutions to social problems. New York: Aca-
demic Press.
Varela, J. (1977). Social technology. American Psychologist 32:914-923.
Vining, D. R. (1986). Social versus reproductive success: The central theoreti-
cal problem of human sociobiology. Behavioral and Brain Sciences 9:167-
216.
References 307

Vygotsky, L. S. (1978). Mind in society. The development of higher psychological


processes. M. Cole, V. John-Steiner, S. Scribner, and E. Souberman (Eds.).
Cambridge, MA: Harvard University Press.
Walker, S. (1983). Animal thought. London: Routledge & Kegan Paul.
Warrington, E. K. (1982). Neuropsychological studies of object recognition. Phil-
osophical Transactions of the Royal Society of London B 298: 15-33.
Warrington, E. K., & T. Shallice. (1984). Category specific semantic impair-
ments. Brain 107:829-854.
Warrington, E. K., & A. M. Taylor. (1978). Two categorical stages of object
recognition. Perception 7:695-705.
Warrington, E. K., & L. Weiskrantz. (1974). The effect of prior learning on
subsequent retention in amnesic patients. Neuropsychologia 12:419-428.
Warrington, E. K., & L. Weiskrantz. (1982). Amnesia: A disconnexion syn-
drome? Neuropsychologia 20:233-248.
Wason, P. C., & P. N. Johnson-Laird. (1972). Psychology of reasoning. Cam-
bridge, MA: Harvard University Press.
Watson, J. B. (1913). Psychology as the behaviorist views it. Psychological Re-
view 20:158-177.
Weiskrantz, L. (1977). Trying to bridge some neuropsychological gaps between
monkey and man. British Journal of Psychology 68:431-445.
Weiskrantz, L. (1980). Varieties of residual experience. Quarterly Journal of
Experimental Psychology 32:365-386.
Weiskrantz, L. (1982). Comparative aspects of studies of amnesia. Philosophical
Transactions of the Royal Society of London B298:97-109.
Weiskrantz, L. (1985). Introduction: Categorization, cleverness and conscious-
ness. Philosophical Transactions of the Royal Society of London B308:3-19.
Weizenbaum, J. (1976). Computer power and human reason. San Francisco:
Freeman.
Welch, I. D., G. A. Tate, & F. Richards (Eds.). (1978). Humanistic psychology: A
source book. Buffalo, NY: Prometheus Books.
Whetherick, N. (1979). The foundations of psychology. In N. Bolton (Ed.), Philo-
sophical problems in psychology (pp. 89-110). London: Methuen.
Wiener, N. (1948). Cybernetics: Control and communication in the animal and
the machine. Cambridge, MA: MIT Press.
Wiesel, T. (1982). Postnatal development of the visual cortex and the influence of
environment. Nature 299:583-591.
Williams, D. R., & H. Williams. (1969). Automaintenance in the pigeon: Sus-
tained pecking despite contingent nonreinforcement. Journal Experimental
Analysis of Behavior 12:511-520.
Wise, S. P., & K. H. Mauritz. (1985). Set-related neuronal activity in the premo-
tor cortex of rhesus monkeys: Effects of changes in motor set. Proceedings of
the Royal Society of London B 223:331-354.
Wittgenstein, L. (1967). Zettel. G. E. M. Anscombe & G. H. von Wright (Eds.).
Oxford: Basil Blackwell.
Wolpe, J. (1958). Psychotherapy by reciprocal inhibition. Stanford, CA: Stanford
University Press.
Wolpe, J. (1981). Behavior therapy versus psychoanalysis. American Psycholo-
gist 36:159-164.
Wood, C. C. (1982). Implications of simulated lesion experiments for the interpre-
tation of lesions in real nervous systems. In M. Arbib, D. Caplan, & J. C.
308 References

Marshall (Eds.), Neural models of language (pp. 485-509). New York: Aca-
demic Press.
Worden, F. G., J. P. Swazey, & G. Adelman (Eds.). (1975). The neurosciences:
Paths of discovery. Cambridge, MA: MIT Press.
Wright, A. A., H. C. Santiago, D. K. Kendrick, & R. G. Cook. (1985). Memory
processing of serial lists by pigeons, monkeys, and people. Science 229:287-
289.
Yates, J. (1985). The content of awareness is a model of the world. Psychological
Review 92:249-284.
Young, J. Z. (1973). Memory as a selective process. In Australian Academy of
Science, Report: Symposium on Biological Memory (pp. 25-45). Canberra.
Zeki, S. (1980). The representation of colours in the cerebral cortex. Nature
284:412-418.
Zoeke, B., & V. Sarris. (1983). A comparison of "frame of reference" paradigms
in human and animal psychophysics. In H.-G. Geissler, H. F. J. M. Buffart, E.
L. J. Leeuwenberg, & V. Sarris (Eds.). Modern issues in perception. Amster-
dam: North-Holland.
Zola-Morgan, S., L. R. Squire, & M. Mishkin. (1982). The neuroanatomy of
amnesia: Amygdala-hippocampus versus temporal stem. Science 218: 1337-
1339.
Zuriff, G. E. (1985). Behaviorism: A conceptual reconstruction. New York: Co-
lumbia University Press.
Name Index
Acuna, C., 215 Bartlett, P. C., 110, 179, 181, 214
Adamson, K. K., 256 Baudry, M., 189
Adamson, W. C., 256 Bayes, T., 85
Adelman, G., 142 Beaulieu, A., 166
Ader, R., 145 Bekesy, G. von, 76, 94, 151, 175
Adler, N., 252 Bellarmino, Cardinal, 116
Aertsen, A., 164 Benedict, R., 222
Agassi, J., 19 Beninger, R. J., 248
Aggleton, J. P., 208, 244 Bergson, H., 267
Aguayo, A. J., 143 Berkeley, G., 8
Albright, T. D., 162 Berlyne, D. E., 130
Alcmaeon, 5, 166 Bernard, c., 144
Alcock, J. E., 114 Berthoz, A., 151, 242
Alzheimer, A., 162,240 Bertrand, G., 177
Andersen, R. A., 197 Bindra, D., 8, 10,99, 111, 163,
Anderson, E., 189 185
Anderson, J. A., 80, 190, 191, Binet, A., 256, 262
203,204 Bitterman, M. E., 49, 156, 160
Anderson, J. R., 108 Blakemore, c., 164
Andersson, M., 84 Bliss, T. V. P., 150
Anstis, S. M., 102, 108 Bloom, P. E., 75
Aquinas, St. Thomas, 9 Bloom, M., 164
Ardila, R., 229, 230, 254, 259, Boas, P., 222
261, 273 Borger, R., 278
Aristotle, 3, 9, 11, 35, 54, 94, Boring, E. G., 7, 30, 75, 94, 123
119, 125, 205, 222 Bouchard, T. J., Jr., 122
Armstrong, D., 8 Bower, T. G. R., 153
Asano, T., 205 Braille, L., 174
Augustine, St., 8 Bredenkamp, J., 63
Austin, G., 124 Brentano, P., 238
Averroes, 9 Bridgman, P. W., 73, 125, 126
Ayer, A. J., 8, 9 Brill, A. B., 180
Broad, C. D., 8, 50
Bachelard, G., 61 Broca, P., 83, 94, 154, 161, 166
Bachevalier, J., 182 Brody, N., 113
Baghdoyan, H. A., 163 Bruce, C., 162
Bandura, A., 124 Bruner, J., 124
Baranyi, A., 150 Brunswik, E., 30
Barlow, H. B., 164 Buchner, L., 8
310 Name Index

Buffon, G. L., 206 Davidson, D., 49, 135


Bullinger, M., 69 Davidson, J. M., 171,233,241
Bunge, M. A., 8, 10, 12, 15, 16, Davidson, R. J., 233
19,21, 34, 35, 37, 44, 45, Dawson, M. E., 179
47, 50, 56, 59, 63, 65, 72, Delius, J. D., 202, 203, 205
74, 77, 78, 92, 93, 95, 98, Dell, G., 249
101, 102, 111, 112, 119, Dennett, D. C., 50, 108
120, 126, 127, 128, 133, Dennis, S. G., 162
174, 176, 192, 194,201, Deri, M., 211
202,208,211,212,216, Descartes, R, 10,26,71, 75, 94,
245, 249, 254, 260, 266, 113
270,271,274,275,279, Desimone, R., 3, 4, 8, 162
280,284 Dewey, J., 26
Burnstine, T. E., 151, 174 Diaconis, P., 114
Burt, C., 155,256,257 Dickinson, A., 179
Burtt, E. A., 19 Diderot, D., 8
Bush, R. R, 118 Dietzgen, J., 9
Dilthey, W., 223
Calford, M. B., 174 Dimond, S. J., 10, 164,233
Campbell, C. B. G., 156 Dingler, H., 125
Campbell, D. T., 226 Dixon, N. F., 239
Caramazza, A., 211 Doty, R W., Sr., 233
Carnap, R., 8 Douglas, R. M., 150
Cartwright, D. S., 123 Down, J. L., 155
Casanova, 172 Duhem, P., 116
Chang, F. F., 188 Dumont, J. P. C., 156
Chomsky, N., 5, 8, 95, 96, 109, Dworkin, B., 144
135, 154, 193
Cioffi, F., 278 Easter, S. S., Jr., 152
Claparede, E., 214, 251 Ebbinghaus, H., 94
Cohen, F., 252 Eccles, J. c., 4, 8, 9, 10,22,50,
Cohen, N. J., 146, 162 84, 113,215,233,247,267
Cohen, P. R, 108 Eddy, R L., 121
Cole, M., 154,213 Edelman, G. M., 10,233
Comte, A., 35, 116 Einstein, A., 75, 253, 254
Cowey, A., 76 Engel, B. T., 169
Craik, K. J. W., 206 Engels, F., 267
Crews, D., 160 Epicurus, 8
Crick, F., 178, 183 Erikson, J. R., 109
Cynader, M. S., 174 Espinosa, E., 164
Estes, W. K., 80, 108, 111
Da Vinci, L., 8, 253 Evanczuk, S., 164
Damasio, A. R, 233, 240 Evarts, E. V., 169, 178, 197,
Damasio, H., 240 198, 215
Darwin, C. R, 75, 156, 167,267 Eysenck, H. J., 113, 155
Name Index 311

Fachinelli, C. F., 205 Goethe, J. W., 37


Farah, M. J., 200 Goldberg, G., 215
Fechner, G. T., 92, 94, 175 Goldberg, M. E., 162
Feger, H., 63 Goldman, P. S., 153
Feher, 0., 8, 150 Goldman-Rakic, P. S., 153
Feigenbaum, E. A., 108 Golgi, N., 76
Feigl, H., 8 Goodnow, J. J., 124
Fernandez-Guardiola, A., 233 Gould, J. L., 149, 168
Feyerabend, P. K., 8, 61 Gould, S. J., 122, 125, 145
Fichte, J. G., 8 Graham-Brown, T., 205
Fishbein, E., 8, 211 Graydon, M. L., 174
Fisher, S., 113 Green, B., 211
Flohr, H., 150 Greenberg, R. P., 113
Flourens, P., 94 Greenough, W. T., 151, 174, 188
Flynn, J., 161 Gregory, R. L., 159, 199
Fodor, J. A., 95, 96, 108, 111, Griffin, D. R., 233
164,200 Gross, C. G., 162
Forthman, D. L., 187 Grudin, J., 276
Fox, B. H., 146 Griinbaum, A., 113
Frankl, P., 17 Guilford, J. P., 256
Freud, S., 23, 26, 58, 68, 112,
113,115,152,182,200, Hadley, R. D., 143
238, 249, 254, 255, 268, Haeckel, E., 158
275,277 Halgren, E., 201
Frost, D., 196,240 Hamlet, 284
Furedy, J. J., 4, 8, 179 Hansel, E. E. M., 114
Fuster, J. M., 83, 180 Harbluk, J. L., 109
Harlow, H. F., 38
Galaburda, A. M., 152 Haugeland, J., 50, 108
Galanter, E., 118 Havrankova, J., 156
Galen, 166, 218 Hearst, E., 7
Galilei, G., 116, 176 Hebb, D.O., 8, 10,76,93, 123,
Gall, F. J., 160, 164 133, 134, 148, 163, 166,
Gallup, G. C., 248 178, 179, 181, 185, 190,
Garcia, J., 187,214 191, 195, 204, 206, 216,
Gauss, C. F., 199 255
Gazzaniga, M. S., 200, 233, 242, Hegel, G. W. F., 8
247 Held, R., 196,240
Gelade, G., 103, 197,240 Helmholtz, H. L. F. von, 8, 11,
Gentner, D., 276 94, 159
Geogopoulos, A., 215 Henry M. (H. M.), 182, 214, 240
Gerstein, G., 164 Herbart, J. F., 72
Geschwind, N., 97, 152, 162,244 Herd, J. A., 252
Gibson, J. J., 120 Herder, J. G., 222
Goddard, G. V., 150, 180 Hermes, 143
312 Name Index

Herodotus, 222 Jones, C. H., 114


Herrnstein, R J., 202 Jones, R. S., 190, 203
Herskovits, M. J., 226 Joseph, J. A., 169
Hess, W. R, 161, 170 Jourdain, M., 66
Hilgard, E. R., 233, 242 Julesz, B., 120, 196
Hillix, W. A., 30, 51, 52, 278 Jung, C. G., 152
Hippocrates, 166, 200, 218
Hirsh, R., 5, 124, 182, 188 Kaas, J. H., 174
Hobbes, T., 8 Kahneman, D., 76
Hobson, J. A., 8, 163 Kamin, L., 122, 125, 155
Hodges, D. S., 121 Kandel, E. R., 141, 149, 168
Hodos, W., 156 Kant, I., 9, 17,49, 116,207
Hoffman, B., 214 Keeser, W., 69
Hofstadter, D. R., 50 Kendall, S. B., 248
d'Holbach, P. H., 8 Kennard, C., 68
Hollard, V. D., 202, 203 Kessen, W., 216
Holzman, J., 200 Kirton, M. J., 206
Homans, G., 274 Kluckhohn, C., 222
Hornig-Rohan, M., 145 Kmetz, J. M., 27
Hotton, N., 156 Knapp, A. G., 203
Hoyle, G., 149 Koelling, R. A., 187
Huarte de San Juan, J., 166 Koffka, K., 100
Hubel, D. H., 103, 153, 164, Kohler, W., 92, 100
203 Kojima, T., 205
Hiibner, M., 240 Korsakoff, S. S., 181,240
Huerta, M. P., 174 Kosslyn, S. M., 200
Hughlings Jackson, J., 8, 113 Krech, D., 134
Hull, C. L., 130, 133, 135, 170 Krechevsky, I., 124
Hume, D., 116,238 Kripke, S., 50
Humphrey, N., 233, 246 Kubota, M., 205
Huxley, A., 258 Kuhn, T. S., 61, 224
Huxley, T. H., 8
La Mettrie, J. O. de, 8
Ibn Khaldun, 222 Lacan, J., 4, 17,58, 112, 115
Ingvar, D. H., 143, 233, 243 Laming, D., 176
Land, E. H., 176
Jackson, J. N., 68 Laplace, P. S. de, 199
James, Wm., 8, 26, 216, 244 Larson, J., 148
Jarochewski, M., 9 Lashley, K. S., 8, 15, 80, 84,
Jasper, H. H., 177 123, 124, 134, 135
Jerison, H. J., 8, 156 Layzer, D., 122
John, E. R., 180 Le Bon, G., 222
Johnson, RD., 114 Le Chatelier, H., 119
Johnson-Laird, P. N., 108, 164, LeDoux, J. E., 233, 247
211,213,233 Lehrman, D. S., 121, 154, 193
Name Index 313

Leibniz, G., 8, 11, 113 Maslow, A. H., 17


Lenin, V. 1., 9, 258 Masterton, R. B., 156
LeRoith, D., 156 Matarazzo, J. D., 252
Lesniak, M. A., 156 Matsuzawa, T., 205
Levine, M., 124 Matthews, L. J., 212
Lewin, K., 215 Maudsley, H., 186
Libet, B., 215, 233 Mauritz, K. H., 178
Lieberman, P., 158 Maxwell, J. C., 75, 119
Lindauer, M., 149 Mayer, R. E., 124
Lloyd Morgan, C., 167,267 McCloskey, M., 211
Locke, J., 94 McCulloch, W. S., 111
Locke, S. E., 145 McDougall, W., 8
Lombardi, C. M., 205 McGue, M., 122
Lfllmo, T., 150 McGuigan, F. J., 146
Lorenz, K., 121, 193 McKenna, F. P., 176
Lotze, R. H., 8 McLachlan, D. R., 109
Luce, R. D., 118 McNaughton, B. L., 150
Lucretius, 8 Mead, M., 222
Lugaro, 185 Medawar, P., 145
Luria, A. R., 8, 101, 154, 179, Meehl, P. E., 85, 132, 134
210,213,217,243,246, Melvill Jones, G., 151,242
247 Merzenich, M. M., 174
Lydic, R., 163 Metzinger, T., 50
Lynch, G. S., 148, 189 Michel, F., 240
Lynch, J. C., 215 Mill, J. S., 104, 176
MacCorquodale, K., 132, 134 Miller, G. A., 72, 106, 178, 239
MacDougall, Wm., 238 Miller, N. E., 144, 252
Mach, E., 8, 116, 151 Milner, B., 182, 214, 217, 240
MacKay, D. M., 106, 108 Milner, P., 10, 161, 178, 197
Mackintosh, N. J., 179 Mimura, K., 149
Macphail, E. M., 75 Mishkin, M., 182, 183, 188, 196,
Mahoney, M. J., 258 199, 208, 209, 233, 244
Maine de Biran, 268 Moleschott, J., 8
Majerus, M., 160 Moliere, 66
Malamut, B. L., 182,244 Montesquieu, 222
Mandler, G., 178, 216, 233 Moore, J., 129
Margolis, J., 22 Moore, M. C., 160
Marks, D. F., 114 Morris, R. G. M., 189
Marler, P., 169 Motley, M. T., 249
Marr, D., 108, 110, 199 Motter, B. C., 197
Marshall, J. C., 276 Mountcastle, V. B., 8, 10, 197,
Martin, H., 149 215, 233
Martin, U., 149 Murdock, M., 113
Marx, K., 9, 75, 261 Murofushi, K., 205
Marx, M. H., 30, 51, 52, 278 Murray, E. A., 182, 196,244
314 Name Index

N. N., 240 Petrides, M., 214


Nadel, L., 189 Pettigrew, J. D., 174
Nathans, J., 121 Piaget, J., 3, 26, 35, 68, 92, 101,
Neisser, U., 38, 179 102, 153, 154, 173,214,
Nelson, R. J., 174 277, 278, 284
Newberry, B. H., 146 Piantanida, T. P., 121
Newell, A., 80, 108, 129 Pinel, P., 254
Newton, I., 59, 75, 85, 119, 167 Plato, 8, 9, 111,208,276
Nottebohm, P., 188 Plum, P., 241
Nowak, B., 205 Poppel, E., 195, 196,233,240
Popper, K. R., 4, 8, 9, 10,50,
O'Donald, P., 160 65, 113,208, 211, 224, 247
O'Keefe, J., 189 Posner, M., 241
Oakley, D. A., 277 Pratt, C. C., 125
Oatley, K., 233 Precht, W., 150
Oedipus, 112 Pribram, K., 237
Ohm, G., 131,212 Prioleau, L., 113
OIds, J., 8, 161, 164 Ptolemy, 116, 119
Ono, K., 180 Purves, D., 143, 152
Ornstein, R. E., 241 Putnam, H., 108
Orwell, G., 258 Pylyshyn, Z., 108, 111
Osgood, C. E., 71
Othello, 284 Quine, W. v. 0.,8,9

Paillard, J., 240 Rachman, S., 113


Paivio, A., 111 Rager, G., 152
Palermino, C. P., 187 Rakic, P., 144, 152
Palm, G., 191 Ramachandran, V. S., 102, 108
Papez, J. W., 170 Ramon y Cajal, S., 8, 76, 143
Parducci, A., 80 Reed, T. E., 122
Parkinson, J. K., 182 Reiss Jones, M., 109
Patry, J. L., 70 Rensch, B., 8
Paterson, R., 103 Restle, P., 124
Patton, J. H., 212 Rice, A. G., 187
Paunonen, S. V., 68 Richards, P., 49
Pavlov, I., 8, 93, 94, 118, 129, Richelieu, Cardinal 166
166 Rignano, E., 214
Pears, D., 50 Ritz, S. A., 190, 203
Peirce, C. S., 82 Robertson, R. M., 156
Penfield, W., 8, 76, 170 Robinson, D. N., 9, 10,22
Pericles, 229 Roe, A., 156
Perot, P., 170 Rogers, C., 17,255
Perrez, M., 113 Romanes, G .. 75
Petermann, B., 105 Rorschach, H., 64, 253
Petri, H. L., 182, 188 Rosenzweig, J. L., 156
Name Index 315

Roth, J., 156 Smith Churchland, P., 93


Rusiniak, K. W., 187 Socrates, 207
Russell, B., 8 Spearman, c., 95
Russell, E. S., 155 Spence, K. W., 124, 125
Ryle, G., 11 Sperry, R., 8
Sagi, D., 120 Spiegler, B. J., 244
Sainati Nelo, M., 211 Spinoza, B., 8
Sakata, H., 215 Spitzer, N. C., 152
Santa Claus, 214 Squire, L. R., 162, 163,209
Sapir, E., 222 Stelmach, C. E., 240
Sarris, V., 63, 80, 176 Sternberg, R. J., 214
Saunders, R. C., 182, 244 Stevens, R., 233
Schacter, D. L., 109, 162, 182, Stevens, S. S., 125, 176
233, 240 Stoerig, P., 240
Schlick, M., 8 Stone, G., 252
Schmidt, H., 114 Strawson, P. F., 4
Schmitt, F. 0., 142 Stryker, M. P., 174
Schneirla, T. C., 8 Suppes, P., 125, 268
Schoppman, A., 174 Swash, M., 68
Schwartz, J. H., 141
Schwartz, N. B., 156 Tanaka, Y., 225
Sciolis Marino, M., 211 Tang, Y., 180
Scoville, W. B., 182 Tanzi, A., 185, 190, 191, 195, 204
Scribner, S., 154,213 Tate, G. A., 49
Searle, J. R., 50 Taylor, c., 275
Sechenov,1. M., 94 Taylor, F. V., 258
Segal, B., 242 Tees, R. C., 151, 174
Segall, M., 226 Teilhard de Chardin, P., 8, 156
Sellars, R. W., 267 Teuber, H.-L., 15, 83
Shallice, T., 69, 217, 233, 239 Theophrastus, 9
Shanmugam, A. V., 225 Thompson, R. F., 10
Shannon, c., 105 Thoreau, H. D., 259
Shiloach, J., 156 Thorndike, E. C., 94
Shinoda, Y., 178, 197, 198, 215 Thucidides, 222
Shows, T. B., 121 Thurstone, L. L., 120, 191
Silverstein, J. W., 190,203 Tinbergen, N., 121
Simon, T., 256 Titchener, E. B., 251
Simon, H. A., 80, 108,212 Tolman, E. C., 124, 130, 133,
Simpson, G. G., 156 135, 170
Singer, W., 179 Tranel, D., 233, 240
Skinner, B. F., 8, 37, 118,125, Treisman, A., 103, 196, 197,240
130, 132, 135,251,252, Triandis, H. C., 225
255, 258, 259, 274 Trotsky, L., 258
Sloman, A., 108, 110 Tulving, E., 182,233,240,243,
Smart, J. J. c., 8, 9 250,280
316 Name Index

Tunnell, G. B., 70 Weizenbaum, J., 111


Tuomela, R., 132 Welch, 1. D., 49
Turing, A., 8,108,117,142 Wernicke, C., 83, 94, 154, 161,
Turner, M., 164 166,272
Tversky, A., 76 Wertheimer, M., 100
Whetherick, N., 7
Uexktill, J. von, 177, 238, 284 Whitehead, A. N., 8, 267
Underwood, G., 233 Whorf, B. L., 222
Ungerleider, L. G., 199 Wiesel, T. N., 103, 153, 164
Uttal, W. R., 10, 276 Williams, D. R., 192
Williams, H., 192
Valentine, E. R., 74 Wilson, D. H., 233, 247
Van der Loos, H., 153 Wilson, G., 113
Van Gogh, V., 253 Windelband, W., 223
Van Hoesen, G. W., 240 Wise, S. P., 178, 197, 198,215
Van Rillaer, J., 113 Wittgenstein, L., 8, 11,24
Vanderwolf, C. H., 248 Wolpe, J., 113, 135
Varela, J., 258, 260 Wood, C. C., 80
Vassiliou, G., 225 Woolsey, T. A., 153
Vassiliou, V., 225 Worden, F. G., 142
Vico, G., 222 Wundt, W., 8, 53, 92, 94, 222,
Vogt, C., 8 223, 251
Vygotsky, L. S., 101, 130, 154,
214, 215, 246, 247
Yates, J., 247
Young, J. Z., 149
Warrington, E. K., 184,203,244
Young, R., 180
Wason, P. c., 164, 213
Watson, J. B., 8, 99, 116, 130,
146, 156,222,251 Zeki, S., 176
Weir, J., 160 Zoeke, B., 176
Weiskrantz, L., 76, 182, 184, Zola-Morgan, S., 209
196,202,233,240,241, Zook, J. M., 174
244, 248, 249 Zouzounis, J. A., 162
Weiss, S. M., 252 Zuriff, G. E., 117, 125, 131
Subject Index
Affect, 98, 169-173 Artificial intelligence, 29, 110
AI. see Artificial intelligence Assembly, 102-103
Aims Association, 99,121,187, 189-191
of psychology, 34-39, 59 cross-modal, 195-196
of psychotechnology, 260-261 free, 64
of research, 45, 53-55 Associationism, 104, 189
Alertness, see Attention and Atlas of mental maps, 174, 209
Awareness Atomism, 45-46, 163; see also
Ambiguous figure, 123-124 Individualism, Faculty
Amnesia, 181-182,239-241,244 psychology, and Neuron-
Analogy, 276; see also Metaphor ism
Analysis, 19, 103-104, 196-197 Attention, 103, 177-179, 194,
Anatomy, 189-190, 278-279 196-197, 238, 242
Animism, 8, 14 Automatization of behavior, 169
Anomaly, ontological, 10 Autonomism
Anthropology, 222, 224 methodological, 268
Antiexperimentalism, 79-80 psychophysical, 8, 14, 15
Antitheoretical bias, 37 Awareness, 234, 235-236
Aphasia, 161 altered states of, 241
Approach, 43-61 animal, 239
applied, 44 development, 239
atomistic, 45-46 divided, 242, 245
behaviorist, 51 evolution, 245-246
biological, 51
doctrinaire, 44, 49 Background, 19, 44, 52, 53, 55,
empirical, 44, 48-49 113
holistic, 46 formal, 57, 59
humanistic, 44, 49 philosophical, 57, 58
interdisciplinary, 48, 270-274 specific, 57, 59
mathematical, 44 Behavior, 26-27, 222-223
mentalist, 51 abnormal, 252-255
nonscientific, 48-50 adaptive, 173
scientific, 44 automatic, 168-169
systemic, 46-48 determinants of, 119-120
technological, 44 maladaptive, 173
VUlgar, 44, 48 Behaviorism, 5, 6, 18, 36, 51,
Argument, 34 53-56, 70, 81, 100, 116-
Ars inveniendi, 212 135, 139-140, 154, 186,
Artifact, 29, 74, 78-79 192, 273-274
318 Subject Index

methodological, 5, 116 Composition, 47-48


ontological, 5 Computation, 95, 199
philosophical, 116 Computationalism, 108-111, 199,
radical, 3, 26 212, 214-215, 218
Biology, 5, 6, 7, 15, 113, 115, Computer, 29
139, 268-269; see also -brain analogy, 80, 107, 160
Neurobiology mind,210
Black-boxism, 117-120 model,80
metaphysical, 117 Concept, 125-126, 201-207
methodological, 117 abstract, 202, 204-205
Blindsight, 195-196, 239-240 empirical, 201-202
Bond, 101 Conditioning, 118, 248
Bottom-up research strategy, Confirmation, 81, 103
140, 269 Connectivity, neuronal, 143,
Brain, 141-146 146-147
Brain function, specific, 146, change in, 150, 180, 204; see
164, 170, 185 also Plasticity
basic, 166-184 hard, 147
higher, 185-218 soft, 147
Bridge formula, 266, 271 Consciousness, 63, 82, 159, 179,
209, 234, 236
Case study, 69 altered states of, 241
Categorization, 35, 201-215 animal, 238
mathematical model of, 203- causal efficacy, 242, 245
204 collective, 221
Causality, 77 content, 236
Cell assembly, 188, 197,203, degrees, 244
206,211 development, 239, 245
Chance, 114 evolution, 245-246
Channel capacity, 106 stream, 238
Class, social, 227-228 Consensus, 25, 248
Classification in psychology, 32, Construct, see Concept, Hypoth-
97-98 esis, Problem, and Theory
CNS (Central Nervous System), Control, experimental, 114, 249
141-165 Controversy
Cognition, 33, 95, 98, 170, 172, hermeneutic, 9, 23
207-215 scientific, 9, 30
Cognitivism, 170,207,212,214- Correlation, statistical, 85
217 Cortex, cerebral, 141-142,243-
Coincidence, 210 244
Common sense, 111 Cortico-limbic system, 181-182,
Communication system, 107, 246 191, 244
Community, scientific, 58 Cortico-striate system, 182, 191
Comparative study, 155-156 Cortico-thalamic system, 244
Competition, interneuronal, 152 Counterrevolution in science,
Compensation, intermodal, 174 115, 134
Subject Index 319

Creativity, 205-206, 214 Ecological validity, 38, 281


Critical period, 149, 153-154 Education, 255-257
Criticism, 22, 23 EEG (Electroencephalography),
Culture, 224-229 74
"carpentered", 226 Efficacy, 38
design, 258-259 Efficiency, 38
subjective, 225 Embodiment, 111
Curiosity, 178 Embryological process, 152
Emergence, 7, 12, 13,46, 100,
Data, 81, 85, 99, 127-128 101-102, 151, 188, 191,
Definition, 96, 126-127, 265-266 205, 245-246, 266-267
axiomatic, 126 developmental, 102, 151-153
explicit, 126 dynamic aspect of, 267
implicit, 126 evolutionary, 102
operational, 73, 126, 247-248 Emotion, 161-162,208; see also
reductive, 266 Affect
Delayed matching to sample Empiricism, 5, 9, 201-202, 205-
task, 83 206
Dendritic branching, 143 Encephalization, 157-158
Deprivation, sensory, 123,201 Endocrine system, 157
Depth, conceptual, 23, 31 Energy, conservation of, 10
Description, 34, 36, 54, 139-140, Engram, 181-184, 208
182,274 Environment, 47-48
Determinism, 216 Environmentalism, 120-125,
Development, 121, 151-155,239, 154
245-246 Epiphenomenalism, 8, 14-15,
cognitive, 153, 158 113-114
Diagram, commutative, 131 Epistemology, 20-21, 53, 62
Dialectics, 9 Ethology, 120-121
Disconnection syndrome, 97, 99, Event, 12, 174
162, 244, 283 Evolution, 94, 98, 102, 107, 155-
Discrimination, 175, 196 160, 168, 185, 193-194,
Disease, mental, 253-255 199, 202, 210, 246, 267,
Disregard for exceptions, 210-211 277
Distinction vs. separation, 33, Exactness, 22, 23
97,272 Excitation, neuronal, 147-148
Domain of a science, 57-58 Executive function, 216-217
Double dissociation, principle of, Expectation, 178
83 Experiment, 10,68,76-81,83,
Dream, 112-113,200-201,241 93, 114, 160, 247-249
Drive, 170-171 philosophical presuppositions
Dualism, psychophysical, 7-17, of,77-79
26, 50, 52, 58, 92-94, 113, Experimental analysis of behav-
156, 159, 160, 166, 178, ior, 252
180, 184, 224, 267, 283 Experimental animal, choice of,
Dynamics, 167 79
320 Subject Index

Explanation, 35-36, 51, 54, 78- Generalizer, 204


79, 109, 115, 119, 131, Genetics, 154-155,254
132-134, 139-141, 150, Gestalt, 100, 123, 200
173, 182, 183,211,222, Gestalt school, 46, 53, 95, 100-
274-280 105, 197, 210
developmental, 277 Glue formula, see Bridge for-
environmental, 277 mula
evolutionary, 277 Group-study method, 68-69
genetic, 276
mentalist, 275-276 Habit, 182
metaphorical, 276 Habituation, 148-149, 186, 188
pseudo-, 275--276 Haeckel's "law", 158
psychological, 121, 167-169, Hallucination, 201, 241
274-279 Hedonism, 192
physiological, 277-278 Historico-cultural school, 5, 6
tautological, 275 Holism, 29, 46, 142, 146, 160-
teleological, 275 161, 163
Extrapolation of animal experi- Homogenization, cultural, 231-
mentation to humans, 81- 232
82 Humanities, 5
Humanization, 228
Factor analysis, 95 Hypnosis, 243
Faculty, mental, 95-97 Hypothesis, 36, 81-82,112,114,
Faculty psychology, 23, 95-97, 265, 274-275, 279
272 negative, 84
Fallacy, 76 null, 84, 114
Family, 230 programmatic, 15, 85
Fatigue, neuronal, 124 statistical, 85
Feature, 102-103 Hypothesizing, 124, 199
detector, 103, 164, 195, 197 Hypothetico-deductive system,
-integration theory of atten- 191; see also Theory
tion, 196-197
Fechner's law, 175-176 Idea, 16
Feedback, 118, 119 Idealism, philosophical, 5, 6, 7,
Field study, 68-70, 84 8, 15,51,215
Fragmentation of psychology, Identity hypotheses, 12-17, 31,
30-31, 272-273, 281 50, 55, 58, 92, 104, 165,
Framework, 58 166, 178, 184,218,265-
Future of psychology, 280-282 270
Fuzziness, conceptual, 9, 10, 12, strong, 12, 139
112 weak, 12
Geisteswissenschaft, 5 Ideology, 60, 75, 115
General outlook, 19; see also Idiographic, 223; see also Case
Worldview study
Gene, 193-194, 276 Illusion, visual, 226
Subject Index 321

Imagery, 200, 206 Invertebrate, 28, 157, 179


Imaging techniques, 143 IQ, see Intelligence
Immaterialism, 16, 17,89
Immune system, 145-146 Kinematics, 167
Imponderable, 94 Know-how, 182,209; see also
Independence of psychology, Habit and Knowledge,
273; see also Autono- procedural
mism, methodological Knowledge, 149, 194,208-209
Indicator, 73, 86, 92, 127, 162, background, 44, 57
238, 247-248 declarative, 240
hypothesis, 73, 127-130,213 engineering, 110; see also Arti-
psychological, 129 ficial intelligence
Individual, 25, 49 fund of, 57, 59
Individualism, 274; see also innate, 208
Atomism kinds of, 209
Induction, 65, 82 limits to, 194, 280
Inference, 81-85 objective, 208-209
Information, 33, 106 ordinary, 11, 44
processing, 105-111, 239 procedural, 240
theory, 105-106 scientific, 11
Inheritance, 122-123 subjective, 208
Inidealism, 16, 17 tacit, 2
Inhibition, neuronal, 147-148, objective, 208-209
151 Know-that, see Knowledge,
lateral, 151, 175 declarative
Innateness hypothesis, 96, 193
Insight, 212 Laboratory study 84; see also
Instinct, 121, 192-193 Observation, Experiment
Insulation, experimental, 78-79 and Measurement
Instrument, 75 Language, 96, 206-207, 222
Integration, conceptual, 270-274; evolution, 158, 246-247
see also Merger learning, 154, 193, 210
Intelligence, 75, 122-123, 125, Law, 109
155, 213, 228 probabilistic, 78
testing, 213-214, 256-257 psychological, 49-50, 119, 131
Intention, 169,215-217,238,275 Lawfulness, 77
Intentionality, 238 Learning, 27, 28, 118, 121, 148-
Interactionist, psychophysical, 8, 149, 171-173, 180-181,
10, 14 185-194, 226-230, 239,
Interpolation method, 85 248
Introspection, 63-64, 68, 70-71, associative, 189-191
98-99 behaviorist definition, 186
Introspectivism, 99 equation, 190-191
Intuitionism, 46, 100, 206 impairment, 189
Invention, 66, 75-76, 82 laws, 191-192,229
322 Subject Index

mechanism, 188-189 Mathematics, 59


multiple trial, 187-188 Matter, 7
neurobiological definition, 187 Measure, 72
single-trial, 181, 188, 192 Measurement, 71-76, 125-126
verbal, 64 technique, 73-76
visceral, 144-145 Mechanism, 54, 114, 118, 119,
Level, 140, 269-270 140, 278-279
Limbic system, 170, 171,207- "Meditation", 241
208,244 Memory, 109-110, 149, 179-184,
Linguistics, 95 188,244
Localization of mental functions, episodic, 162, 181-182, 188
96, 160-164 mental, 180
Localizationism, 160-166, 191, of skills, 162, 181-182; see
218 also Habit
mosaic, see Localizationism somatic, 180
strong, 162 system, 162
weak, 163 Mentalism, 5, 6, 51-53, 46, 89-
Location, 197-198 115, 134, 139-140
Logic, 13 nonscientific, 52-53
Logical empiricism, see Positiv- scientific, 52-53
ism Merger of fields and theories,
Love, 38 271
Message, 107
Metaphor, 181
Magnitude, 71-73 Metaphysics, 116, 117
additive, 72 Method,63
extensive, 72 experimental, 56
intensive, 72 general,63
Malfunction, brain, 162; see also inductive, 65
Disease, mental scientific, 56, 65-66
Map special, 62-63
body, 174 Methodics, 45, 53-55
cognitive, 149, 209 Methodology, 62-86
navigational, 189 Microreductionism, 266-267,
of the world, 177 269; see also Neuronism
sensory, 173-174 Mind, 7-17, 26, 89, 266
somatotopic, see Map, body action on matter, 8, 10, 145
Marxism, 9 mapping onto brain, 166
Mass action "law", 80 origin, 157-158
Materialism, philosophical, 5, 6, power, 245
10, 12, 16, 17,22 privacy, 75
eliminative, 8 Mind-body dualism, see Dual-
emergentist, 7, 8, 10,91-92 ism, psychophysical
reductive, 8 Mind-body problem, 7, 8, 15, 52
vulgar, 7, 9, 10 Mistake, 249
Subject Index 323

Model, 54, 114, 127-128, 191- Normality, behavioral, 252-254


192, 199-200 criteria, 253-254
box -and-arrow, 107-108 Normative claim, 19
input-output, 142 Novelty
mental,211 detector, 177
Modularity, see Faculty, mental qualitative, 13; see also Emer-
Modus ponens, 212 gence
Molar study, 140
Monism, neutral, 8, 14-15 Object, 92
Monism, psychophysical, 7-17; constancy, 202
see also Identity hypothe- Objectifier, see Indicator
ses Objectivity, 34, 74
Mood, 145 Observation, 66, 68-71
Morality of scientific research, clinical, 68-70
21,53 field, 69-70
Motivation, 171-172 scientific, 66, 68-69
Motor act, 119 Ontology, 11, 20; see also Me-
Movement taphysics
automatic, 167-169 Onus probandi, 84
voluntary, 31-32, 169,215-216 Operationalization, 127-128
Mystery, 43 Operationism, 125-130, 247-248
Organicism, 46
Nativism, 120, 154 Orgasm, 241
Nature-nurture controversy, 101, Other minds, 209-210
121-123, 154-155
Naturwissenschaft, 5 Panpsychism, 8, 156
Neobehaviorism, 130-132; see Parallelism, psychophysical, 8,
also Variable, intervening 14, 15, 104, 113
Neurobiology, 139-165 Parapsychology, 39, 50, 111-112,
developmental, 151-155 114-115
evolutionary, 155-160 Passivity, 144
Neuroendocrine system, 156, 171 Pattern recognition, 195,210-211
Neurolinguistics, 161 Percept, 195, 200, 202
Neurology, 143 Perception, 27, 102-104, 120,
Neuromuscular system, 119, 169 194-201, 226, 235-237
Neuron, 141-142; see also Cell nonconscious, 239-240
assembly Person, 3
Neuronism, 163-164 Personality
Neuroscience, 16, 31-32, 59, 96, dual,242
135, 267; see also Neuro- theories, 112, 123, 159
biology Phenomenon, 116
Neurotransmitter, 107, 143, 156 Phenomenalism, 8, 116, 176
Nomothetic, 223 Philosophy, 2; see also Episte-
Nonsense syllables, 64 mology and Ontology
Norm, 229-230 inherent in science, 17-21
324 Subject Index

interaction with psychology, 2, Psychobiology, 5, 6, 13,26,51,


3, 281, 285 55-56, 139-141, 166-218;
of mind, 2, 3, 7-17, 58, 110- see also Psychology,
111 physiological
of psychology, 21-24 Psycholinguistics, 5
Physicalism, 7, 91 Psychology
Physics, 25 aims of, 34-39
Planning, 214, 215, 217 applied, 60-61, 251-262
Plasticity biological, 51
behavioral, 150 classical, 94-99
functional, 150 clinical, 48-49, 112
mathematical model, 148 cognitive, 23, 33, 52, 89, 105-
neural, 96, 146-154, 169, 174, 111,132,207-215,272-
180, 189 273
Potentiation, neuronal, 148 comparative, 167
Poverty, theoretical, 37 computational, 108-111
Positivism, 5, 6, 116, 125 cross-cultural, 226-227
Preconception, 211 definitions, 26-27
Predicate, 13 dialectical materialist, 130
Predictability, 216 ecological, 120
Prediction, 34, 36 educational, 125, 255-257
Presupposition, 16, 86, 109 evolutionary, 94, 155-160,
of scientific research, 17-21 277
Prewiring, 152 folk, 3, 17,24,48, 111
Problem, 38, 43, 65, 67 humanistic, 5-6, 17, 39, 49
conceptual, 233-234 industrial, 257-258
empirical, 233-234 information-processing, 33, 95,
methodological, 63 105-111, 181
solving, 211-213 molar, 278
Problematics, 44, 52, 54-55, 59 philosophical, 4, 17, 18,50,
Program 99, 111
computer, 108, 110, 212 place of, 223-224, 271-274
mental,216-217 physiological, 6, 10, 31, 94
Property, 101, 126 pop, 111-115
emergent, 13, 101-102,266 pseudoscientific, 60; see also
global, 101 Parapsychology and Psy-
primary, 176-177 choanalysis
resultant, 266 relevance to philosophy, 285
secondary, 177 scientific, 7, 16, 18, 34-39, 56-
Prosopagnosia, 240 61
Pseudoscience, 112 social,221-232
Psi, 78-79, 114 Psychophysics, 176
Psychiatry, 252-255 Psychosomatic disorder, 146
Psychoanalysis, 5, 38, 39,111- Psychotechnology, 251-262
115, 123,253-254 Purpose, 238; see also Intention
Subject Index 325

Qualia, 176; see also Property, Response, 131-132, 168,204


secondary Responsiveness, 235
Quantitation, 71-72, 247 Revolution, scientific, 61, 75,
111, 115, 134
Randomness, 78 Reward, 172
Rationalism, 201-202, 206, 207 Rewiring, 150-151
Rationality, 212-213 Role, 227
Reactivity, 234 Rorschach test, 64
Readiness, 178; see also Atten- Rule, 63, 108-109,212
tion
Realism, epistemological, 8, 9,
19, 35, 53, 78, 91-92 Sapir-Whorf hypothesis, 222
critical, 283-284 Scholasticism, 11, 22
naive, 176 Schools in psychology, 30
scientific, 176-177, 202 Science, 110
Reality, 9, 77, 90-91 applied, 35
"construction", 285 basic, 35
Reallocation, functional, 151 behavioral, 27, 222-224
Recognition, 197-198 cognitive, 208, 273
Recording of events, 179-180 natural, 5, 223-224
Reduction, 265-274 social, 5, 6, 28, 112, 223-224,
Reductionism, 12, 13, 164, 223, 268-269, 273-274
265-270 spiritual, 5
biosociological, 269 Search, 103
emergentist, 266 Sectorial thinking, 23, 144
epistemic, 268, 270 Self, see Consciousness
moderate, 266 Self-assembly, 205
ontological, 265-267, 270 Self-awareness, 234, 236, 248
radical, 266 Self-consciousness, 237, 245-247
Reference, 238; see also Refer- antero-, 237
ent full, 237, 242
Referent, 271 pro-, 237
central, 28-29 Semantics, 125
of psychology, 27-29 Sensation, 173-177, 194-195
of social science, 28 Sensitivity, 235
peripheral, 28-29 Sensor, 173-174,235
Reflex, 151, 169,215 Sensory system, 195-196
Reflexology, 104 Set, 123, 178
Refutation, 84 motor, 178
Reinforcer, 129, 192 Sex, 170-171,249
Repair, neural, 143-144, 150 role, 230
Representation, internal, 111, Signal, 106
140, 206, 238 Simulation, computer, 80-81
Research, scientific, 19, 23, 24 Sleep, 163
program, 192 Slip of the tongue, 249
326 Subject Index

Social matrix of behavior, 221- inhibitory, 147


232 plastic, 147
Socialization, 228-230 Synaptogenesis, 150
Society, 28, 81, 221 Synthesis, 103-104; see also
S-O-R psychology, 117, 130-131, Integration and Merger
135 System, 46-47, 100, 101, 161
S-O-R-F psychology, 118 composition, 47
Spoonerism, 249 conceptual, 100
Soul,89 environment, 47
Split-brain, 242 functional, 100-101, 164
Spontaneity, neuronal, 142, 144, input-output, 117
151, 191, 205-206 material, 100
S-R psychology, 117-125, 130, neuronal, 142, 146; see also
135, 151, 168 Cell assembly
Specificity of psychology, 273 structure, 47, 100
Speculation, 11, 50, 115 Systemicity, 23
Speech, 83, 140, 161, 217, 247 Systemism, 46, 163
Spirit, 5, 6
Spiritualism, see Idealism Tanzi-Hebb hypothesis, 190-191,
State 195,204
brain, 10 Technician, 65
internal, 118, 172 Technique, 62-65, 68
mental, 10 imaging, 74-75
Statistics, mathematical, 78, 81, invasive, 73-74
84-85, 114, 122 noninvasive, 73-74
Status, social, 227-228 scientific, 64
Species, 35 semiscientific, 64
Steering, cognitive, 215 Technology, 110
Stevens's law, 176 Telekinesis, 10, 114
Stimulation, 170, 196,201 Teleology, 121, 159,201,275
Stimulus, 120, 123, 131-132, 142, Telepathy, 113-114
147, 168,204,234 Test
Storage metaphor, 181 mental, 213-214, 256-257
Stress, psychological, 146 of theories, 127-128
Structure, 47-48, 100 Testability, 23,112-113
Style in philosophy, 2, 22-24 Thalamus, 177
Subjective experience, 76, 89, Theology, 6, 10
90-94 Theory, 12, 19,36-37,55, 73,
Subjectivism, 34-35, 91 85, 108, 114, 126-128,
Subliminal perception, 235, 244 132, 133, 191-192,212,
Submergence, 151, 244 250,275,279,281
Symbol, 205 axiomatized, 126, 133
Synaptic junction, 145-149 black box, 37
elastic, 147 depth,37
excitatory, 147 exactness, 36-37
Subject Index 327

generality, 36-37 ontological, 38


merger, 271 practical, 39
strength, 36-37 Valuation, 92, 172-173
truth, 37 Value system, 172, 225, 228,
Therapy, psychological, 254-255 231-232
Thing, concrete, 16 Variable, 112
Thinking, 109, 110, 124, 206- intervening, 53, 130-132
207, 236-237 Vertebrate, 157
divergent, 214; see also Crea- Vestibulo-ocular reflex, 150-151
tivity Verstehen, 5
Thurstone's learning function, Vision, 176-177, 195-199
120, 191 residual, 239-240; see also
Time series analysis, 69 Blindsight
Tomography, 75
Top-down research strategy, 46, Whole, 100, 101-103; see also
140,269 Gestalt and System
Tower of London test, 217 Will, 32, 93, 215-217
Tradition, 7, 43 free, 215-216; see also Inten-
Turing machine, 142 tion
Turing test, 117 Wiring, see Connectivity
World
external, 90-91
Unconscious, 237, 249
inner, 90-91
Unification of psychology, 31-33
possible, 50
Unobservable, 73, 129
real, 177
Utopia, 258
Worlds, Popper's three, 9, 10
Worldview, 19, 45, 104, 246
Vagueness, 82
Validity Yogis, 144-145
alethic, 39
ecological, 38, 39 Zeroth law, 78, 119
methodological, 38 Zoology, 29, 59

Anda mungkin juga menyukai