MORITZA
MORITZA
net/publication/326669076
CITATIONS READS
0 1,622
1 author:
Gerald Ulrich
Charité Universitätsmedizin Berlin
221 PUBLICATIONS 1,330 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Gerald Ulrich on 30 July 2018.
IAG
IUGG
ICSU
ÖAW
in gratitude
Preface
Schuster, bleib bei deinem Leisten!
German proverb
“The rabbi spoke three times. The first talk was brilliant: clear and
simple. I understood every word. The second was even better: deep
and subtle. I didn’t understand much, but the rabbi understood all of
it. The third was by far the finest: a great and unforgettable experience.
I understood nothing, and the rabbi didn’t understand much either.”
This was one of Niels Bohr’s favorite anecdotes (Folse, 1985, p. 258).
All books on philosophy belong to one of the three types of the
rabbi’s talks. In an introductory text like the present book, the reader
is entitled to expect that it belongs to “Rabbi Type 1”. This is at least
what we have tried to achieve.
Our higher polytechnical schools have frequently become “Technical
Universities” or “Universities of Technology”. This implies that they
intend not only to give a profound scientific or professional education,
but also to offer a touch of “universality”. Now, the common interdisci-
plinary background of all scientific, engineering and medical disciplines
is becoming increasingly “philosophical”.
On all aspects of philosophy of mathematics and natural science
there are excellent and even brilliant monographs. The intention of the
present book is much more modest: to provide an absolutely introduc-
tory yet rather systematic and comprehensive textbook taylored to the
interests of students of science, technology, and medicine.
This book is written not by a professional philosopher, but by a
practicing scientist. The lack of philosophical depth is thus perhaps
partly made up by a knowledge of the mentality and interests of stu-
dents of scientific disciplines and by active research experience in sci-
ence, in which interdisciplinary and philosophical questions are arising
naturally and to an ever increasing extent.
vii
viii
?
Kant
?
Hegel
@
@
@
Kurt Gödel Niels Bohr
To avoid misunderstandings: the names stand for certain directions
of thinking and have no direct relation to “greatness”. (Otherwise one
might ask: why Plato and not Aristotle, why Bohr and not Einstein or
Heisenberg?)
Only the first three are “professional” philosophers. Gödel is the
greatest of modern logicians, and Bohr, through his “principle of com-
plementarity”, is the physicist who has made dialectics respectable in
the philosophy of natural science. When he was awarded a title of nobil-
ity for his scientific merits, he chose for his coat of arms the inscription
“Contraria sunt complementa ”.
Dialectics has frequently fallen into discredit by its dogmatic use
and misuse, and also Hegel has suffered from this circumstance. Nev-
ertheless dialectic thinking does have its use in natural philosophy, and
ix
Ruth Hödl did all the word–processing in her usual efficient, dedicated,
and painstaking way. Last but not least, my wife Gerlinde read various
versions of the manuscript and was my adviser in questions of biology
and theology, besides confirming that the book can be read also without
mathematics. All this help is gratefully acknowledged.
Helmut Moritz
xii
Contents
Preface vii
B Natural Science 71
3 Physics 73
3.1 Classical mechanics and determinism . . . . . . . . . . . 73
3.2 Deterministic chaos . . . . . . . . . . . . . . . . . . . . . 80
3.3 Probability . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.4 The theory of relativity . . . . . . . . . . . . . . . . . . . 90
3.5 Quantum theory . . . . . . . . . . . . . . . . . . . . . . 99
3.6 Elementary particles . . . . . . . . . . . . . . . . . . . . 111
3.7 Space and time; cosmology . . . . . . . . . . . . . . . . . 120
xiii
xiv CONTENTS
C Philosophy 195
5 Philosophy for scientists 197
5.1 Realism, idealism and dualism . . . . . . . . . . . . . . . 197
5.2 The three–world model . . . . . . . . . . . . . . . . . . . 207
5.3 Subject and object . . . . . . . . . . . . . . . . . . . . . 214
5.4 Historical landmarks . . . . . . . . . . . . . . . . . . . . 225
Index 285
Part A
Human Perception and
Thinking
Chapter 1
The human brain
3
4 CHAPTER 1. THE HUMAN BRAIN
Neurons
Each brain contains about 1011 nerve cells or neurons. Fig. 1.2
shows, very schematically, a typical nerve cell. The synaptic buttons
(some 1014 !) attach to dendrites or somas of neighboring cells. There is
a tiny “synaptic cleft” between a button and the next dendrite, which is
bridged by chemicals (neurotransmitters ) sent from button to dendrite.
The normal speed is only around 5 meters per second; the speed
can be increased by a myelin sheath of the axon up to 100 m/s. Again,
the reader is referred to the literature for details.
Important is the following fact: the “firing” of an impulse is all or
none. The shape and size of the impulse is irrelevant; important is
only whether the neuron fires or not. Synapses (consisting of synaptic
buttons and synaptic clefts) may be excitatory or inhibitory. Their
combined effect on the neuron determines whether it fires or not. The
system: all (1) or nothing (0) is essentially digital, using a dual number
system (consisting of zero’s and one’s) which is not unlike the dual
system used by a digital computer. Very simple examples may be found
in (Penrose 1989, pp. 393–394).
Each historical period compares the human body to its favorite ma-
chines: formerly it was a mechanical or heat engine, with a pump for
the heart. Now it is fashionable to compare the brain to a huge parallel–
processing digital computer. (There seems to be some “analogue com-
putation” too, but on a minor scale, by “non–spiking neurons”.) An
important difference to present computers is a high redundancy: if part
of the “brain computer” fails, other parts may take over.
Assumptions about the brain computation have lead to the con-
struction of artificial neural networks, which, together with more con-
ventional computing devices, play a considerable role in artificial intel-
ligence (AI).
H. Petsche (Vienna) pointed out to me that the “digital firing” (all
or none) mentioned above does not imply that the brain works exactly
like a digital computer. From (Cohen and Stewart 1994, p. 454) we
quote:
Nerve cells don’t “compute” digitally, but they do use discrete pulses to
communicate over large distances.
That “brain computation” is not really “digital”. For a detailed
treatment of “neural computing” cf. (Kohonen 1988, Chapter 9: a
very readable non–mathematical chapter in a otherwise rather techni-
cal book). Also the discussion in (Penrose 1989, pp. 392–399) is highly
relevant.
Let us also not forget that “computation” or logical thinking is
only a tiny fraction of our intellectual activity which also includes joys,
desires, fears, poetry, music, etc., cf. end of sec. 2.1.
Suggested additional reading. The interesting reader may start with
the didactic masterpiece (Ornstein and Thompson 1984) and continue
with more technical books such as (Eccles 1973) or (Thompson 1985).
8 CHAPTER 1. THE HUMAN BRAIN
The relation between brain and mind is one of the most controversial
problems in philosophy. It is closely related to the problem of matter
and mind.
At a first glance, there seems to be no problem at all. Our thoughts,
feelings, and emotions are clearly mental , belonging to the mind. A
tree, a house, or a stone are clearly material , consisting of matter: we
can kick a stone, live in a house, or fall from a tree, breaking our leg.
What could be more material?
Modern science gives a rather different picture of matter. It consists
of extremely small molecules, separated by large distances. Molecules
consist of atoms, and the atoms have a tiny nucleus with electrons or-
biting around them like planets orbiting around the sun, and separated
by similarly huge (in a relative sense) distances. So matter essentially
consists of empty space. If my foot kicks a stone, an empty space kicks
another empty space (to be honest, what matters are the forces that
act in these “empty spaces”).
If modern science makes matter less “material”, it also makes mind
less “mental”, as the philosopher Bertrand Russell said. The study of
our brain has shown us that, e.g., emotions are related to the activity
of the (material!) limbic system, and even the most sublime manifes-
tations of human thinking seem to correspond to the firing of certain
neurons in the cortex.
A fair and didactically useful comparison between brain and mind
seems to be possible by modern computer terminology: the brain is
the hardware, the mind is the software of human thinking. The same
software can be run on different computers, and the ideas contained in
the present book (supposing that it contains any) can be understood
by different readers, that is, by different brains.
1.2. BRAIN AND MIND 9
(2) Mind is a new aspect of brain activity that emerged when biolog-
ical evolution led to the genus “man”.
10 CHAPTER 1. THE HUMAN BRAIN
REPLY
Dear Sir:
Your astonishment’s odd:
I am always about in the Quad.
And that’s why the tree
Will continue to be,
since observed by
Yours faithfully,
God.
Plato, Kant, Fichte, and Hegel are also considered idealists, al-
though in a somewhat different sense: they regard mind as a
primary concept, and matter a concept derived from mind.
This view is not so far from opinions of some modern physicists
who argue as follows. It is very difficult to define matter directly.
For a physicist it is natural to say that matter is what satisfies the
laws of physics. Now these laws are expressed by mathematical
formulas, which are certainly mental rather than material struc-
tures. Similarly, the entities satisfying mathematical formulas are
mathematical functions, which are equally mental. Hence matter
is a mental construction.
Such an argument can be refuted, as almost all philosophical ar-
guments can. Nevertheless it seems to incorporate at least a spark
of truth. Finally, we shall discuss
12 CHAPTER 1. THE HUMAN BRAIN
Auditory perception
Although our main emphasis will be on visual perception, the sense
of hearing also presents great interest. We listen to the voice of a friend,
or to a piece of music.
The ear is an organ of extreme sensitivity. If it would be only a little
more sensitive, we would constantly hear an unbearable background
noise produced by the thermal motion of the air molecules. The inner
ear contains a beautifully designed resonance mechanism which, so to
speak, performs a very precise and detailed harmonic analysis of the
auditory signal, a sound wave. The individual frequencies are heard
distinctly and separately. If we strike a chord of four simultaneous notes
1.3. HUMAN PERCEPTION 15
on the piano, we hear all four frequencies distinctly in their true pitch
(if we have an absolute ear) or in their correct relation of frequencies
(which is enough for the average musician and music–lover). Thus if we
strike on the piano simultaneously the notes c and e, we shall hear both
of them distinctly and not “averaged” to an intermediate note d, say.
This is important because in visual perception precisely this “averaging
effect” occurs.
On the other hand, auditory perception in man does not give very
precise information on the direction from which the sound comes, and
on the location of the source, except when these faculties are especially
trained in blind people. (Animals such as bats can “hear” directions
very precisely!)
It is also worth noting that each half of the brain receives informa-
tion from both ears.
Visual perception
This is the classical form of perception, on which philosophers tra-
ditionally have laid the greatest emphasis. It is well known that each
eye acts like a small photographic camera. A lens generates a picture
of the outer world on the retina, again usually a very precise picture,
which is then transmitted by its optical nerve to the cortex. The main
relay stations of both “optical tracks” are the two lateral geniculate
bodies (LGB), one in the right and the other in the left part of the
thalamus, cf. Fig. 1.1 on p. 4. The left hemisphere of the brain (or the
left half–brain) receives its visual information from both eyes, but from
the right half of the visual field of each eye! It is similar for the left
visual field of each eye, from which information goes exclusively to the
right LGB and the right visual cortex.
Optical processing occurs mainly in the retina, the LGB’s and the
visual cortex of both hemispheres.
Slight differences in the direction of the optical axes lead to small
differences in the respective images on both retinas, called parallaxes,
and they provide the third dimension, depth. This is called stereoscopic
vision.
In view of the high state of present electronic image processing, it is
tempting (but not cogent) to compare the processing of visual data of
the eye with our image processing technologies, and stereoscopic vision
has been employed already for several decades by photogrammetry for
the same purposes, namely to construct a three–dimensional model
from two–dimensional images.
16 CHAPTER 1. THE HUMAN BRAIN
Evolutionary epistemology
Epistemology is nothing else than “theory of knowledge ”. The fa-
mous biologist Konrad Lorenz (1903–1989) was the last professor on
the philosophical chair of Immanuel Kant in Königsberg. Studying
Kant’s philosophy, he recognized that Kant’s a priori conditions for
human knowledge could be identified with man’s perceptual structures
(eye, ear, etc.) developed in the course of human evolution. Man’s
world view is largely, though by no means exclusively, conditioned by
his physical and mental constitution. For instance if, instead of per-
ceiving, in the electromagnetic spectrum, only the wave lengths from
400 to 700 nm, we would be able to visualize also other frequencies,
then our visual representation would be quite different. (We do per-
ceive infrared, but as heat.) Or think of the other extreme, a blind
person.
20 CHAPTER 1. THE HUMAN BRAIN
Set theory
The notion of set should nowadays been known from school. (In the
present treatment, we shall pursue an intuitive approach, like in school,
sacrificing full logical rigor to simplicity.) A set consists of elements. It
can be discrete or continuous (Fig. 2.1).
Figure 2.1: A discrete set (A) and a continuous set (B) in the plane
23
24 CHAPTER 2. LOGIC AND MATHEMATICS
s∈M . (2.1)
A∪B (2.3)
and intersection
A∩B (2.4)
(Fig. 2.3).
From Fig. 2.2 we see that
s∈A∩ B ∩ C , s ∈ M , s ∈ M ∩ A∩ B ∩C ,
A∪B ⊂ M , A ∩B ⊂ M , A∪ M = M ,
A∩M =A ,
etc.
We also need the concept of complementary set (with respect to some
comprehensive set M ), see Fig. 2.4. In an obvious notation we may
write
A=M −A . (2.5)
In the latter case “aRb ” means “a son is younger than (his) father”.
There are many excellent introductions to set theory, e.g. (Halmos
1960).
Numbers
Let us try to define the integers 1, 2, 3, . . . Consider the number 2.
Obviously, “2” may be considered the property which is common to all
pairs:
2 = set of all pairs .
−2 = 1 − 3 = (1, 3) = (5, 7) = · · ·
0 = 1 − 1 = (1, 1) = (2, 2) = · · ·
Logic of propositions
A proposition, denoted by p, q, r, . . ., is a sentence or a statement.
E.g.
p ... It rains.
q ... Today is Friday.
r ... The street is wet.
Propositions can be true or false. In the first case, we assign the truth
value T or simply 1, in the second case, F or simply 0.
The following symbols are used:
The reader might wish to check this by means of the concrete ex-
ample just given.
The third truth table can be explained in a similar way; note that it
is formally identical to the multiplication table for the numbers 1 and
0. This was one of the starting points for Boole’s logical algebra!
The fourth truth table is almost self–evident: p ⇔ q is true if p = 1,
q = 1 or p = 0, q = 0 and false otherwise. (p = 1 is an abbreviation for
“the sentence p has the truth value 1”.) Other truth tables can easily
be established.
We have the basic laws (which should be checked by means of our
special example):
p∨ ∼ p There holds always p or not–p;
p is either true or false (nothing
else): law of the excluded middle,
“tertium non datur ”
∼ (p ∧ ∼ p) p and not–p cannot hold
simultaneously:
law of contradiction
∼ (∼ p) ⇔ p double negation = affirmation (2.11)
p⇒p∨q “If it rains, then it rains or
today is Friday”
p ∧ (p ⇒ q) ⇒ q “modus ponens ”
(we can put — ponere — q as true)
(p ⇒ q)∧ ∼ q ⇒∼ p “modus tollens ”
(we can remove — tollere —
p as false)
By means of these and similar laws, all logical conclusions (deduc-
tions, proofs) can be performed in a purely formal manner (e.g., also
by an automatic computer). Therefore we speak of formal logic, in our
case of propositional calculus.
We thus have a “mathematization” of logic, in a similar way as
set theory has provided a “logization” of mathematics. Formal logic
is understood to consist primarily of set theory and the calculus of
propositions.
There is a close connection between these two branches of formal
(or symbolic, or mathematical ) logic, expressed in a correspondence
between ∧, ∨ in propositional logic and ∩, ∪ in set theory:
30 CHAPTER 2. LOGIC AND MATHEMATICS
x∈A∩B ⇔ x∈A∧x∈B ,
(2.12)
x∈A∪B ⇔ x∈A∨x∈B .
Here x ∈ A (x belongs to the set A) is regarded as a proposition, say
p, and x ∈ B is another proposition, say q.
A final note on terminology: logical truths such as (2.11) are called
analytical truths or tautologies. They hold always, independently of the
special meaning of p or q.
1945; Penrose 1989, p. 423), and what about people who are equally
fluent in several languages? In which language do they think? My
answer, based on introspection, is: in none. I also think generally in
images and indistinct structures (unless I prepare explicit formulations
in a certain language for lectures or for writing).
It seems, however, that linguistic formulations play, in philosophy
or history, a much larger role than in mathematical sciences. This may
be one of the reasons why English is used as a lingua franca almost
universally in natural sciences but much less so, e.g., in history. Phi-
losophy is often difficult to translate, and poetry is frequently almost
untranslatable.
We do no wish to downplay the importance of language for human
thinking. It has been claimed that a child who has not learned to speak
one (any) human language by the age of about 12 years, will forever be
unable to think as a normal adult (Cohen and Stewart 1994, p. 355).
Language is indeed an indispensable prerequisite for human thinking.
Thus there is no doubt that linguistic analysis is important in logic
and philosophy, but it seems to be a tool rather than a goal in itself.
Our expression need not be verbal. Getting a red face may express
our embarrassment better than any words; and furiously walking out
of the room could be our final argument. The behaviorist psychologists
may not be so wrong after all!
Earlier we have talked about nonverbal thinking in mathematics.
What about the performance of a Beethoven sonata by a pianist? In
sec. 1.1 we have seen that the cerebellum is heavily involved. Does this
mean that the performing pianist is not active intellectually? If you
think so, don’t tell it to her; otherwise you may get a painful nonverbal
answer!
These examples from elementary psychology show that “verbal
thinking” is indeed only a small fraction of our total intellectual activ-
ity. They are not meant to discredit logic but to place logical thinking
into its proper, lofty but highly abstract, place.
Logical antinomies
Already around 1900, Russell, analyzing work by Gottlob Frege
(1848–1925), met with the first difficulties in reducing mathematics to
logic (more precisely, to set theory). He found antinomies or paradoxes,
that is, logical contradictions.
Russell’s antinomy: Let M be the set of all sets which do not contain
themselves as elements. Does M contain itself? Answer: it contains
itself as element, if and only if it does not contain itself! Thus the
concept of the set M is obviously contradictory.
This requires some abstract thinking. A more concrete form has
been given to it also by Russell: A barber in a village shaves exactly
those of its male inhabitants who do not shave themselves. Does the
barber shave himself? Yes, exactly then if he does not shave himself!
Russell’s antinomies and similar paradoxes show that the concept of
“set” in its “naive” form may be contradictory. Various methods have
been tried in order to avoid such antinomies:
(1) Limitation to simple logical systems in which it is impossible to
even formulate such antinomies.
(2) Axiomatic set theory: the axioms should exclude “dangerous”
sets. However, such attempts so far have been largely “ad hoc”. Even
after having excluded known antinomies, one cannot be sure that later
on one cannot find new antinomies in some remote corner of the system.
So far, the consistency of axiomatic set theory has not been proved!
Hilbert’s program of “formalism ” (around 1920) was intended to
give formal calculi for logic and mathematics whose consistency was
to be proved according to mechanical rules (as realized, e.g., by an
automatic computer).
Gödel’s theorem
Hilbert’s program received a death blow in 1931 when the Austrian
logician Kurt Gödel proved that such a program is impossible.
2.3. LOGICAL PARADOXES AND GÖDEL’S THEOREM 35
Self–reference
The common feature of all these paradoxes is self–reference: Rus-
sell’s set contains itself, the liar asserts something about what he himself
is just saying, etc.
It seems that self–reference has an importance far beyond some
esoteric paradoxes: its applications reach from human thinking (“the
thinking thinks the thinking”) and artificial intelligence to cosmology.
Hofstadter (1979, 1985) has written two remarkable heavy volumes
about it. Self–reference also seems to lie at the basis of dialectic think-
ing. Thus we shall meet it frequently in our book.
Let us here mention the delightful little book (Smullyan 1980),
which is a veritable treasure of paradoxes.
The logical antinomies and Gödel’s theorem seem to indicate that
even formal logic and mathematics cannot be made arbitrarily precise.
We may speak of a Gödelian uncertainty of logic and mathematics as
2.3. LOGICAL PARADOXES AND GÖDEL’S THEOREM 37
Foundations of mathematics
We have already met two ways of founding arithmetic or the theory
of natural numbers (positive integers), which may then serve as a basis
of other branches of mathematics:
(1) Logicism: arithmetic can be reduced to logic (set theory), cf. (2.7)
in sec. 2.1.
The main reason for all the problems and paradoxes in logic and
mathematics is infinity. Russell’s set is “highly infinite”; self–reference
implied “zero distance”: the barber shaves himself (he has “zero dis-
tance from himself”), the liar says something about which he is just
saying (both statements are coincident or have “zero distance”), and
“zero”, or “infinitely small”, is also an aspect of infinity, at least in the
present context (remember that one frequently bluntly puts 1/0 = ∞,
although mathematicians may frown on it).
38 CHAPTER 2. LOGIC AND MATHEMATICS
Inexact concepts
Exact
√ are certainly the concepts of mathematics: numbers such as
1, 2, 3, 2, π; geometrical points, straight lines, circles, spheres, etc.
Rather exact are empirically defined physical concepts such as “pre-
cisely” drawn lines, circles, “physical points” realized, e.g., by the in-
tersection of two empirical lines (+) or as the center of small circles (◦)
found on any illustration in a book on geometry. Rather exact are also
the concepts of “logical atomism” (see the end of sec. 2.1): Mr. Smith,
his house, a lamp or a book on his table, Mr. Smith’s dog, an apple,
and a huge amount of similar “well–defined” concrete objects.
Clearly, even the “points” A and B are not defined physically with ab-
solute precision: on paper, they are defined not much better than to an
accuracy of 0.1 mm; in nature, perhaps by a cross (+) on a stone to an
accuracy of 1 mm; most precisely, by marks engraved on metal or glass
to an accuracy of about 0.001 mm. At any rate, the distance d cannot
be measured to an accuracy higher than the accuracy to which physi-
cal points are defined. Let the result of our measurement be 23.281 m.
If the points are defined to 1 mm only, the result may equally well
be 23.280 or 23.282 m. In geodesy, we may write 23.281 m ± 1 mm,
there the “standard error” is denoted by ± 1 mm. This is not precisely
equivalent to the statement that the measurement lies in a “confidence
interval”, say between 23.280 and 23.282 m, but we may disregard this
distinction for the present purpose; cf. also sec. 4.4.
Here we already see that classical principles of logic such as the law
of contradiction or the law of the excluded middle (sec. 2.1, formulas
(2.11)) no longer hold with empirical concepts: p ∨ ∼ p, “d is either
23.281 or not”, is violated, there may as well be d = 23.281 m or d =
23.2812 m (6= 23.281 m). The same holds with the law of contradiction:
∼ (p ∧ ∼ p): d = 23.281 m (p) and d = 23.2812 m (not–p) may hold
simultaneously.
Inexact are many concepts of ordinary life, starting from our body:
nose (where in our face is the boundary line between our nose and an
adjacent cheek?), finger (where on our hand does it begin?), but also
a tall building (how high?), a hiking trail in the mountains (which to
his embarrassment, the hiker frequently finds very difficult to follow),
smoke, a cloud (see also the end of sec. 2.1), or the Earth’s atmosphere
in general (at its “edge” there are less and less molecules in a cubic
meter; if there are only a few stray molecules or ions, is this still atmo-
sphere or already empty space?).
Fuzzy sets
Many sets in ordinary life are not very precisely defined, e.g., the
Earth’s atmosphere just mentioned, considered as a set of molecules.
Such sets with ill–defined boundaries are called “fuzzy” (Fig. 2.7).
Obviously, a cloud is a fuzzy set of water droplets. But also more
abstract sets in ordinary life are “fuzzy”. Think of the set of honest
people. Who belongs to it? Saints, yes; thieves, no; but what about
myself, whose behavior frequently is less than perfect? Or consider the
set of red apples: which shades of red are permitted? Is an apple still
red or already closer to orange? What about a red apple with some
little yellow spots?
40 CHAPTER 2. LOGIC AND MATHEMATICS
Fuzzy propositions
The same “fuzziness” also applies to propositions (statements) of
everyday life. A few examples: “It is hot today” (which temperature?),
“You are stupid”, “Mahler’s symphonies are great”.
Among the most popular unprecise statements are weather fore-
casts. “There will be light rain showers tomorrow.” In the U.S.A.,
newspapers are frequently more explicit: “There is a probability of
25 % for rain tomorrow”. Intuitively this is pretty clear, but what does
it mean precisely? Will rain fall in 25 % of the next day (consisting of
12 h, say), that is, will we have 25 % × 12 = (25/100) × 12 = 3 hours
of rain tomorrow? If instead it rains for 4 hours, will we cease to trust
our newspaper or the weather forecast in general?
Obviously not, we have to live with unprecise statements and try
to interpret them reasonably, guided by experience. Much more diffi-
cult is the task to “precisely” formulate these “unprecise” statements,
especially if we want to simulate human thinking by a computer. Fre-
quently it does not work to consider some facts as precisely true and
contradictory facts simply as false, as we ourselves unfortunately do
often enough, following the example of small children (“I am right!”
“No, I am right!”)
Fuzzy logic
There are various cases and possibilities of formalizing these “fuzzy”
concepts. Probably the oldest one (Gauss and Legendre, around 1800)
is
2.4. INEXACT CONCEPTS, “FUZZY LOGIC” 41
Informal reasoning
The arguments we hear in daily life, in quarrels and disputes, even
in university lectures, are hardly capable of being expressed in the sym-
bolism of formal logic as, for instance, mathematical proofs may be.
Curiously enough, the same holds also for the “informal reason-
ing” in philosophical arguments. Even philosophical concepts such as
causality or determinism, matter and mind, freedom and law, are by no
means sharply defined. They subtly change their meaning during a dis-
cussion. Sometimes this is intentional, sometimes it passes unnoticed.
Lucas (1970, p. 58) speaks of “chameleon words ”.
In philosophical and other discussions, statements are not in general
simply true or false. There are arguments pro and con, some carry great
weight, some arguments are rather weak. Discussions may be intended
to prove the opponent wrong, but this seldom happens. Mostly the
participants in discussions and the readers of philosophical books are
invited to follow the arguments, to appreciate their strength, validity,
and cogency, and finally to form their own opinion.
Nobody could imagine to replace, in a discussion, the two opposing
philosophers by opposing computers. So far, no philosophical book has
44 CHAPTER 2. LOGIC AND MATHEMATICS
Introduction
We may define dialectics as that aspect of human thinking that
transcends “algorithmic reasoning” which could as well (or better) be
performed by a computing machine. A beautiful example has been
pointed out by J.N. Findley in his article “The contemporary relevance
of Hegel” in (Findley 1963). It is Gödel’s theorem.
Formalized thinking based on axioms and, in principle, performable
by a machine, is called an object language, or symbolic language or for-
mal language. The informal language in which we speak about the op-
erations of formal logic, is called a metalanguage (considerations about
mathematics are called metamathematics!). In the present book, the
metalanguage is (Austrian) English. It is a useful convention to con-
sider the metalanguage to be of higher level than the object language.
Using this terminology, we may describe Gödel’s proof (sec. 2.3) as
follows.
In a certain symbolic object language, a proposition G is constructed
which asserts its own formal unprovability. By informal reasoning,
however, it can be shown that G nevertheless is true. Let us quote
Findley (1963):
But the unproveable sentence at the same time soars out of this logico–
mathematical tangle [of the formulation of G in the object language] since
the proof of its unproveability in one language [the object language] is itself
a proof of the same sentence in another language of higher level [the meta-
language], a situation than which it is not possible to imagine anything more
Hegelian.
2.5. DIALECTIC THINKING 45
or better
thesis −→ antithesis
& .
synthesis
The synthesis (S) is not simply a compromise between thesis (T ) and
antithesis (A), but S is a higher–level standpoint from which both T
and A become understandable and even compatible. Thus, in the triad,
the synthesis S, so to speak, lies on a higher level than both T and A.
(It would be absolutely wrong to regard A as the simple logical opposite
∼ T (non–T ) in the straightforward sense of formal logic, cf. sec. 2.1.
46 CHAPTER 2. LOGIC AND MATHEMATICS
Example 5.
A geodetic example: measurement of all three angles α, β, γ in a
triangle (sec. 2.4).
Thesis T : α, β, γ have been measured.
Antithesis A: α + β + γ 6= 180◦ (contradiction!).
Synthesis S: least–squares adjustment as described in sec. 2.4.
As we have seen above (Fig. 2.8) the important property is that the
synthesis lies on a higher level. This is also illustrated by an example
from everyday life.
Example 6.
A discussion in which two participants defend two apparently
contradictory positions (T and A). A skilful chairman manages to
convince both partners that their opinions, if seen from a proper
perspective, are really compatible (S). If all three participants
have good will, this usually succeeds, often surprisingly, giving all
a profound feeling of satisfaction, especially if the initial positions
T and A have appeared quite contradictory.
Let us return to empirical science. All measurements are affected
by measuring errors, and no physical theory can be expected to hold
absolutely. We shall now give an example for a measurement and one
for a theory (cf. secs. 4.4 and 3.4).
Example 7.
T : A measured distance d has the value 18.85 m.
A: This cannot be true because of measuring errors, rounding–
off, etc.
S: d = 18.85m ± 0.007m. This shows that it is an empirical value
and at the same time estimates its accuracy.
Example 8.
T : Newtonian mechanics holds exactly.
A: No, since relativity theory gives better results.
S: Newtonian mechanics holds to a certain accuracy which is
defined by relativity theory (v/c 1, cf. sec. 3.4).
The dialected process can sometimes be iterated in order to come closer
and closer to reality, as the following two examples show.
Example 9. √
A trivial example from mathematics. What is the value of 2?
√ √
T : 2 = 1.4 A: 2 = 1.5
48 CHAPTER 2. LOGIC AND MATHEMATICS
Example 10.
Iterative solution of an equation x − φ(x) = 0, assuming φ(x) to
change only slowly with x.
T : A certain approximate value x0 gives a solution.
A: x0 − φ(x0 ) 6= 0 (contradiction!)
S = T1 : x1 = φ(x0 ) is a better solution
A1 : x1 − φ(x1 ) 6= 0 (contradiction)
S1 = T2 : x2 = φ(x1 ) is again better
A2 : x2 − φ(x2 ) 6= 0 (2.17)
S2 : x3 = φ(x2 )
etc.
2.5. DIALECTIC THINKING 49
freedom — law
accident — necessity
content — form (e.g., in art)
analysis — synthesis
self–preservation — care for others
conservation — progress
subject — object
matter — mind
justice — love
continuous — discrete
theory — practice
simplicity — complexity
competition — cooperation
order — chaos
and many others.
own actions and his own thinking, has missed the great opportunity
to become a human person. Formal “thinking” can also be performed
by machines and higher animals; only self–conscious “thinking about
thinking” is probably restricted to man only. We again remind the
reader what we have said about the dialectic nature of Gödel’s proof
at the beginning of the present section.
“Thinking about thinking” may involve several or even many levels
in the way of Fig. 2.9. A striking example, due to Poul Martin Møller,
may be found in (Bohr 1963, p. 13). A Danish student says:
I get to think about my own thoughts of the situation in which I find myself.
I even think that I think of it, and divide myself into an infinite retrogressive
sequence of “I”s who consider each other. I do not know at which “I” to
stop as the actual, and in the moment I stop at one, there is indeed again
an “I” which stops at it. I become confused and feel dizziness as if I were
looking down into a bottomless abyss, and my ponderings result finally in a
terrible headache.
According to Rosenfeld, this is a “delightfully humorous illustration of
Hegelian dialectics” (Folse 1985, p. 54). It also strongly reminds of
Fichte (Versuch einer neuen Darstellung der Wissenschaftslehre 1797,
II(2); Werke Band I, p. 526, simplified and modernized):
When you say that you are conscious of yourself, you distinguish your think-
ing “I” from the “I” about which you are thinking: the “I” as a subject
(S1 ) from the “I” as an object (O1 ). But in this process you are necessarily
regarding the subject S1 as the object O2 of a new subject “I” (S2 ), and so
on: S1 = O2 , S2 = O3 , S3 = O4 , etc. ad infinitum.
(4) The geometry of dialectics. If we represent syntactic “thinking
about objects” by a movement in a plane, then the dialectic triad (T ,
A, S) certainly does not correspond to a triangle lying in this plane.
The synthesis S rises above it (Fig. 2.8). To use the same metaphor,
dialectics corresponds to a third dimension, rising above the basic plane
of formal logic. This procedure may even be iterated, cf. Fig. 2.9.
The triad, or triangular motion, is by no means a necessary char-
acteristic of dialectic thinking, as we have already mentioned. In-
stead of the triad (T − A − S) Fichte frequently uses a 5–term pentad
(T − A − T A − A T − S), where T A is a partial synthesis with the
accent on T , and A T similarly with the accent on A, and the mathe-
matician Speiser (1952) uses a 7–term process in his attempt at a mod-
ern reconstruction of Hegel’s “Logic ”. Weizsäcker (1992) uses the term
“Kreisgang” (circular movement), in which the observer, so to speak,
52 CHAPTER 2. LOGIC AND MATHEMATICS
Example 12.
T : steering an automobile
A: skidding (on ice, etc.) (negation)
S: countersteering to restore the course (negation of negation).
motion and led to Kepler’s laws and Newton’s mechanics. (The syn-
thesis lies on almost too high a level: according to Einstein’s general
theory of relativity (sec. 3.4), all reference systems and all origins are
theoretically (!) equivalent, cf. Example 3 on p. 46.)
By directing the attention of the observer from the object to his
own perceptual apparatus (cf. sec. 1.4), Kant claimed to have provided
a Copernican revolution in philosophy. In fact, it was followed by a
period of intense flourishing of philosophy (Fichte, Schelling, Hegel).
By looking at the antinomy of the liar from an unexpected angle,
Gödel was able to prove extremely profound and far–reaching theorems
in logic (sec. 2.3). An unprovable proposition is recognized to be true.
A mathematical example which is less trivial than it looks, providing
important
√ generalizations, is as follows.
2 is not a rational number; therefore it was not considered a num-
ber by the ancient Greeks. It can, however, be approximated with
arbitrary accuracy by the rational numbers
x1 = 1.4
x2 = 1.41
x3 = 1.414 (2.18)
x4 = 1.4142
..
.
The sequence x1 , x2 , x3 , . . . does not have a rational number√ as a limit,
it “leads to nowhere”. The solution is Cantor’s definition of 2 as pre-
cisely the entire sequence {x1 , x2 , x3 , . . .}! This principle is frequently
used in mathematics, cf. formula (2.8) on p. 27.
In adjustment theory (secs. 2.4, 2.6, and 4.4), the troublesome and
annoying measuring errors are re-interpreted as stochastic variables
with an interesting mathematical theory.
Mathematically “ill–posed problems” have become central in the
recently fashionable “chaos theory” (sec. 3.2) and “inverse problems”
(sec. 3.8)
Illness, pain, and grief if regarded from the proper point of view,
may be recognized as positive factors in the development of human
personality, “making a virtue of necessity”.
In art, this is the principle of tragedy.
In philosophy, finally, the contrast between materialism (matter is
primary, mind is derived from matter) and idealism (mind is primary,
56 CHAPTER 2. LOGIC AND MATHEMATICS
law of contradiction ∼ (p ∧ ∼ p)
(2.19)
double negation ∼ (∼ p) = p
hold.
If dialectics is to apply to the real world, it must be wider than
formal logic, somewhat like the “fuzzy logic” of sec. 2.4. Then these
laws may not hold, as the examples of sec. 2.4 and Examples 7 to 11
of the present section have shown. Thus we may say:
Dialectic logic is a logic of rational approximation.
This is not the whole story, however. Take, for instance, the process
of acquiring knowledge. According to Fichte, the “I” (the subject, T )
confronts the “non–I”, the surrounding world (the object, A). The “I”
continuously takes some information from the surrounding world and
thus gradually increases its knowledge through a continuous sequence
of syntheses (S).
In this process, the “I” of one year ago (or one minute ago), is
identical and yet not identical to the “I” now since, in the meantime,
I have undergone a development. The law of contradiction no longer
holds.
This important point was very well elaborated by Havemann (1964,
pp. 48-49). Thinking begins with the dialectic (not formal–logic!) con-
tradiction between identity and difference. Throughout the years we
change: I am certainly different from what I was 50 years ago, never-
theless I feel that somehow I have retained my personal identity, and
also legally I have definitely remained the same person.
Thus we may also say:
Dialectic logic is a logic of temporal evolution.
In several places above we have seen the importance of the concept
of complementarity due to Niels Bohr. Thus we say:
58 CHAPTER 2. LOGIC AND MATHEMATICS
A necessary and sufficient condition is that the infinite sum (2.26) “con-
verges”, as mathematicians say, that is, that s results as a finite number.
This is Hilbert space; it is the generalization of n–dimensional Euclidean
space for n → ∞.
Hilbert space is the mathematical tool for quantum mechanics, as
we shall see in sec. 3.5.
Spaces of higher dimension and Hilbert space appear awesome con-
cepts for the non–initiated. The main (or even only) problem is psycho-
logical. The psychological barrier is best overcome by practical com-
putations. For instance, solving a linear system of equations with 5
unknowns means that we work in 5–dimensional Euclidean space. And
if your pocket calculator solves a system with 10 unknowns, you work
in 10–dimensional space! What could be easier than that?
“But I cannot visualize these spaces!”, you say. Let me tell you
a secret. Nobody can visualize a higher–dimensional space, but all
mathematicians speak fluently about points, straight lines, planes, or
subspaces in n–dimensional space. The picture they have in mind when
they speak in this way, is our usual three–dimensional space (or even
the two–dimensional plane). Computing in n–space and at the same
time visualizing 3–space may sound a little schizophrenic but it is al-
ways done in this way, and it always works! So don’t worry. This
schizophrenic way of visualization even works in infinite–dimensional
Hilbert space, as we shall see in sec. 3.5.
Differential calculus. It was invented simultaneously and indepen-
dently by the great philosopher Leibniz (1646–1716) and by the perhaps
even greater physicist Newton (1642–1727). This led to a terrible quar-
rel about priority: even great personalities are only human beings.
What we need here is extremely simple: the Pythagorean theo-
rem (2.21) for infinitely small (infinitesimal ) differences dx, dy, and ds
(Fig. 2.17); ds is called line element. We obviously have
and in n dimensions
a11 x + a12 y = l1 ,
(2.30)
a21 x + a22 y = l2 .
The coefficients aij are known, the left–hand sides l1 and l2 have
been measured, and x and y are unknowns to be determined. For
clarification, let me give a numerical example:
2x − 3y = 10 ,
(2.31)
x + y = 30 .
The solution is x = 20, y = 10, as we immediately find on substitution.
It frequently happens that more observations li are made, in order
to check the determination and to improve its numerical value. For
64 CHAPTER 2. LOGIC AND MATHEMATICS
2x − 3y = 10 ,
x + y = 30 , (2.32)
x−y = 8 .
The solution of the second and third equations gives x = 19, y = 11,
which obviously is different from the solution of the first and second
equations, namely x = 20 and y = 10. We say that the three equations
(2.32) are inconsistent.
Let
a11 x + a12 y = l1 ,
a21 x + a22 y = l2 , (2.33)
a31 x + a32 y = l3 .
l1 + v1 = a11 x + a12 y ,
l2 + v2 = a21 x + a22 y , (2.34)
l3 + v3 = a31 x + a32 y .
l+v = Ax , (2.36)
where
l1 v1
x
l = l2 , v = v2 , x= (2.37)
y
l3 v3
are vectors and
a11 a12
vT v =⇒ minimum , (2.39)
where
vT = [v1 v2 v3 ] (2.40)
is the transpose of the vector v in (2.37).
The solution can be written in the matrix form
−1
x = AT A AT l , (2.41)
A n
What is interesting here is that the solution (2.41) of the problem de-
fined by (2.36) and (2.39) is a projection of a vector l in n–dimensional
space onto an m–dimensional subspace.
Figure 2.19: The plane Ω as the “union” of two orthogonal straight lines
(subspaces) X and Y .
2.6. GEOMETRY: DIMENSIONS TWO TO INFINITY 67
Ω=X∪Y .
a∈Ω ,
a∈X , a 6∈ Y
b∈Y , b 6∈ X .
c 6∈ X , c 6∈ Y .
Let us compare this situation with the set theory of ordinary logic
(sec. 2.1). Here Ω is a set (represented by a square) which is decomposed
68 CHAPTER 2. LOGIC AND MATHEMATICS
Figure 2.20: The set Ω as the union of the two sets X and Y .
Ω=X ∪Y
X ∩Y =∅
X⊂Ω
Y ⊂Ω
a ∈ X , a 6∈ Y
b ∈ Y , b 6∈ X
c∈X , c 6∈ Y .
Denoting
Ω − X = Y = X̄ (not X)
Ω − Y = X = Ȳ (not Y )
c ∈ X or c ∈ X̄
p or ∼ p
p∨ ∼ p ,
c 6∈ X , c 6∈ Y = X̄ ,
either c ∈ X or c ∈ X̄ (p ∨ ∼ p)
73
74 CHAPTER 3. PHYSICS
Two point masses (Fig. 3.1) attract each other with a force F of mag-
nitude F , proportional to the masses m1 and m2 , and inversely propor-
tional to the square of their distance l; this is the famous inverse square
law . Here G denotes a universal constant, the gravitational constant.
We also have the equality of action and reaction: the two forces F1 and
F2 in Fig. 3.1 are equal in magnitude and opposite in direction. The
magnitude of both F1 and F2 is given by (3.11).
The analogy between (3.12) and (3.13), on the one hand, and (3.14)
and (3.15), on the other hand, is obvious.
Thus it is not surprising that both principles are due to Gauss, who
also recognized the deep analogy between them.
It may be shown that Gauss’ principle applied to a free particle on
a surface, does give geodesic motion. For the sphere, motion along a
great circle is obvious, cf. Fig. 3.2.
mẍ1 = F1 (x1 , x2 , x3 ) ,
mẍ2 = F2 (x1 , x2 , x3 ) , (3.16)
mẍ3 = F3 (x1 , x2 , x3 ) ,
3.1. CLASSICAL MECHANICS AND DETERMINISM 77
1
q̇i = pi ,
m (3.18)
ṗi = Fi (q1 , q2 , q3 ) .
∂H
q̇i = ,
∂pi
∂H (3.19)
ṗi = − ,
∂qi
Imagine the figure formed by these two curves and their infinitely many
intersections . . . ; these intersections form a kind of meshwork, tissue, or
infinitely dense network . . . One is struck by the complexity of this figure
which I do not even attempt to draw. Nothing is better suited to give us an
idea of the complexity of the three–body problem and in general of all the
problems of dynamics in which there is no uniform integral [of the motion]
...
The modern theory of general nonlinear dynamical systems is con-
sidered to start with Poincaré’s work. The subject then lay relatively
dormant, known only to a few specialists, until 1954 when the fa-
mous Russian mathematician Andrei Kolmogorov (1903–1987) and his
younger colleague Vladimir Arnold started with a general and system-
atic treatment of such strange trajectories. In 1963 there followed an
independent paper on an application to meteorology by the American
Edward Lorenz. Then the subject exploded. Currently it is probably
the most popular subject of mathematics, known to a broad general
public.
Let me try to explain what Lorenz did. He took the equations of
mathematical weather prediction, simplified them and studied the so-
lution numerically with the help of a computer. These solutions proved
to be extremely unstable: two solutions with almost identical initial
conditions started to diverge wildly (Fig. 3.3). Since the data of me-
teorology are unavoidably insufficient and inaccurate, the initial con-
ditions are not exactly known; small deviations result in completely
different behavior. This is the reason why it is hardly meaningful to
make detailed weather predictions more than a few days ahead. (In
astronomy, predictions are good for tens or even hundreds of years, in
spite of Poincaré . . . )
Let us repeat:
stability: small causes produce small effects;
instability: small causes produce large effects.
82 CHAPTER 3. PHYSICS
3.3 Probability
God does not throw dice.
Albert Einstein
5032
f1 = = 0.5032 ,
10000
(3.30)
4968
f2 = = 0.4968 ,
10000
3.3. PROBABILITY 87
lim f1 = p1 = 0.5 ,
n→∞
(3.31)
lim f2 = p2 = 0.5 .
n→∞
Special relativity
Einstein’s special theory of relativity deals with inertial systems.
An inertial system according to Newton’s theory is a system on which
no force acts, so that equations (3.1) and (3.7) (p. 73–74) hold:
ẍ = 0 ; (3.33)
x = at + b . (3.34)
Eq. (3.33) says that there is no acceleration, and (3.34) says that the
motion of an inertial system is uniform, that is with constant velocity
along a straight line. An example is a spaceship in intergalactic space
3.4. THE THEORY OF RELATIVITY 91
whose rockets have been shut off and which, of course, continues to
move with constant velocity to reach the nearest galaxy.
A system which moves uniformly with respect to an inertial system,
is also an inertial system. According to classical mechanics, the two
are related by a Galilei transformation:
x̄ = x − vt ,
ȳ = y ,
(3.35)
z̄ = z ,
t̄ = t .
These equations can be directly found by inspecting Fig. 3.6.
c0 = c . (3.37)
(4) Special relativity has been confirmed in all relevant cases with-
out any exception. It is perhaps the most accurately confirmed
theory of physics. This is particularly important for high veloci-
ties v almost = c, which occur with very fast–moving elementary
particles in high–energy accelerators (CERN in Geneva, etc.).
3.4. THE THEORY OF RELATIVITY 93
x4 = i c t (3.40)
√
where i = −1 is the “imaginary unit”. Then, and on putting x1 = x,
x2 = y, x3 = z as usual, we may write the line element in the form
is no more unreal than space. What is true is that the imaginary trans-
formation is simply a mathematical artifice to bring the line element
into the Pythagorean form (3.41).
Beautiful as the form (3.41) is, however, it must be admitted that it
may give a false impression that space and time are not different at all.
An imaginary transformation, though easy mathematically, is a serious
operation from the point of view of geometrical and physical reality. A
circle
x2 + y 2 = 1 (3.42)
becomes, on replacing y by iy,
x2 − y 2 = 1 , (3.43)
which is a hyperbola and thus a quite different geometrical figure. This
simple example is relevant for relativity, as we shall see in sec. 3.7; cf.
also (Moritz and Hofmann–Wellenhof 1993, pp. 181–195).
Anyway, it is now customary to use as time coordinate, not x4 , but
x0 = ct , (3.44)
which is a simple real transformation of time scale, so that (3.39) be-
comes
ds2 = −dx20 + dx21 + dx22 + dx23 , (3.45)
which is almost as simple as (3.41), but in which all quantities are
real . It is called a pseudo–Euclidean form (mathematicians call it non–
positive–definite).
General relativity
Surface theory. For introductory purposes, let us first replace four–
dimensional space–time by a two–dimensional space, that is, a plane or
a curved surface. The line element in the plane is, of course,
ds2 = dx2 + dy 2 (3.46)
in Cartesian coordinates x and y. In polar coordinates r, φ with x =
r cos φ, y = r sin φ it becomes
ds2 = dr 2 + r 2 dφ2 . (3.47)
In general curvilinear coordinates (u, v) it may be written
Similarly to (3.49), however, the gij may also be functions of the coor-
dinates xi . The matrix [gij ] is generally called the metric tensor .
Similarly to the case n = 2, the gij may define a flat or a curved
space–time. The criterion, similar to the Gaussian curvature, is now
the Riemannian curvature tensor Rijkl , which is again formed of the gij
in a rather complicated manner. If
Rijkl ≡ 0 , (3.53)
where κ is a constant and Tij is the matter tensor , more precisely the
matter–stress–energy tensor. (The reader will have surmised in the
meantime that “tensor” is any quantity that carries indices. He is
3.4. THE THEORY OF RELATIVITY 97
for a few years. There I found eq. (3.54), and suddenly everything was
clear.
The Riemannian curvature tensor Rijkl does separate gravitation
and inertia, cf. (3.53) and (3.54). It is formed by the metric ten-
sor gij and its space–time derivatives of first and second order (non–
mathematicians, don’t get scared and forget it!). Such derivatives, how-
ever, need gij not only at one point, but also in a surrounding region,
however small. Thus gravitation and inertia cannot be separated at
one point only, but they can be separated in an arbitrarily small neigh-
borhood of it! (The reason is that inertia has a much more regular
space–time structure than gravitation.)
Components of the tensor Rijkl can be measured, at least in prin-
ciple, by instruments called gradiometers. Thus a combination of
gravimeters and gradiometers provides, at least theoretically, a rigorous
method for separating gravitation and inertia.
This is important, not only in aerial gravimetry, where gravitation
is the “signal” and inertia is the “noise”, but also in inertial navigation
which is now used in almost every airplane: here inertia is the signal and
gravitation is the noise. One person’s signal may be another person’s
noise, as every musician knows from experience with his (her) neighbors.
Suggested reading. There are good nonmathematical presentations
of the principles of relativity. My favorites are (Lanczos 1965) and
(Will 1986). An “easy” but mathematical introduction is (Moritz and
Hofmann–Wellenhof 1993). Remarkable as a very readable introduc-
tion to geometry in general is (Lanczos 1970). The treatment in (Lind-
say and Margenau 1957) is excellent as usual. Philosophical problems
of relativity are treated on a high level, also freely using physics and
mathematics where necessary, in (Treder 1974).
In the old small grocery stores it was possible to buy sugar, etc.,
in any reasonable amounts, e.g. 250 or 700 grams. In the modern
supermarkets, sugar comes only in packs of 1 kg, say.
It seems that nature furnishes energy also only in fixed packages of
a (of course, much smaller) constant size, called quantum of energy.
100 CHAPTER 3. PHYSICS
p2 =< ψ, ξ2 > ,
or for ξ9 , namely
p9 =< ψ, ξ9 > . (3.61)
If tonight we listen to the weather forecast for tomorrow, it might
be: p2 = 80% probability for rain and hence p9 = 20% for absence
of rain. (We are purposely using p2 and p9 , which is better for the
quantum analogy, rather than p1 and p2 as we should normally do.)
Next morning we look out of the window: sun is shining brightly. Thus
the event corresponding to the probability p9 has been realized (maybe,
St. Peter wanted to show the meteorologists who was who).
To return to our quantum measurement. We perform the measure-
ment of M and the outcome is µ9 , and hence the new state function is
ξ9 , for the very same reason that after measuring L with result λ3 the
state function was ψ3 .
Thus, as the result of our measurement, the state function, having
been ψ3 before the M –measurement, has suddenly become ξ9 after-
wards.
3.5. QUANTUM THEORY 103
Figure 3.12: A “fold” in the universe. Effects can propagate along the
“normal” space–time route AaB, but also “cut across”
the fold following route AbB
their validity, none of these theories have ever been contradicted by ex-
periment: Popper’s “falsification” (sec. 3.9) has still to come for them.
For the working physicist, the results hold independently of the
philosophical interpretation. The philosophical interpretation of quan-
tum theory is much more difficult and controversial (and therefore even
more interesting) than the interpretation of relativity. The interpreta-
tions of quantum mechanics are closely related to the various interpre-
tations of probability. All interpretations described above seem to be
logically impeccable; their preference may be a matter of taste. (I like
them all.)
The basically probabilistic character of quantum theory holds, of
course, independently of interpretation. The probabilities (3.60) of
transition from state ψ to state ξk exist and cannot be replaced by
something more definite. The radioactive decay (cf. the example of
Schrödinger’s cat) is intrinsically random: the exact instant in which
the radioactive atom decays is unpredictable; it is subject only to a
probabilistic law. Thus quantum theory seems to provide a universal
background of essentially random fluctuations (“vacuum fluctuations”).
Quantum theory particularly clearly shows the relevance of modern
physics to philosophical questions such as the nature of matter and
mind, reality, potentiality, and probability.
It shows, however, also the relevance of philosophy to physics: we
have met with ideas of Aristotle, Kant, Hegel, dialectic materialism,
and Eastern philosophy; and the overarching influence of Plato will
become evident later.
I cannot resist the temptation to call quantum theory a dialectic
synthesis of extreme (mathematical) simplicity and extreme (philosoph-
ical) complexity.
The practical relevance of quantum theory is remarkable. Among
many other achievements, it allows to reduce chemistry to physics.
Quantum mechanics explains the structure of the atoms of chemical
elements, the formation of molecules from such atoms, i.e., the nature
of chemical bonds, etc.
Besides the books already quoted, we mention some more general
works. It is recommended to first read (Davies and Brown 1986), a
wonderfully readable (Rabbi type 1) introduction and overview of the
major interpretations with interviews of the leading proponents. (Grib-
bin 1984) comes next in readability. Bohr’s (1958, 1963) books are basic
but rather difficult (Rabbi type 2). Interpretation problems of quantum
theory with emphasis on the Copenhagen interpretation are dealt with
extensively in the books by Heisenberg and Weizsäcker which are much
3.6. ELEMENTARY PARTICLES 111
(3) Hyperons: they are heavier than nucleons and extremely short-
lived (so–called Λ, Σ, Ξ, and Ω particles).
Nucleons and hyperons together are called baryons (Greek: heavy par-
ticles).
Hadrons consist of smaller particles called quarks (whimsically
called after the quotation by James Joyce which figures as the motto
of the present section). There are some 18 different types of quarks
with deliberately nonsensical names, of different masses and different
electric charges and other properties.
Mesons are formed by the combination of two quarks, and baryons
(nucleons and hyperons) are made up of three quarks. (Leptons seem
to be truly elementary, not made up of any smaller particles.)
3.6. ELEMENTARY PARTICLES 115
y = Ux (3.68)
are called spinors in the case of SU (2) and 3–spinors in the case of
SU (3); they have 2 and 3 complex components, respectively. Enough!
Symmetry and the Greeks. The theory of symmetry played a great
role in ancient Greece, starting at least with Pythagoras in the 6th
118 CHAPTER 3. PHYSICS
Space–time structure
As we have seen in sec. 3.4, the theory of relativity implies a unifica-
tion of space and time. By the line element (3.45) of special relativity
x0 = c t , (3.70)
It is the minus sign which makes all the difference, creating the light
cone separating past, present and future. We can travel in space but not
in time (see below). It is true that the Lorentz transformation (3.38)
also transforms time, but the change is usually negligible in practice
since v c; cf. also the “realistic” third picture in Fig. 3.16.
Nevertheless, if we wish the flow of time to be real (in a philosophical
sense), we should like time to have a unique direction, along which the
creative process of life and mind can proceed. The “present” should
consist of simultaneous events, which would require the two lines in the
third, “realistic” model in Fig. 3.16 to coincide exactly, forming a single
line separating past and present. In reality, however, the two lines in
this figure form an extremely small angle, instead of coinciding as they
do in classical mechanics: the present is “thin”, but not “infinitely
thin”.
Whitehead and, more recently, Havemann (1964, p. 92) have
pointed out a way out of this dilemma. Past and future must be rede-
fined in agreement with relativity: past is of what we already can have
knowledge, future is what we can still influence by our actions. Thus
creative action is possible. Time is not an illusion, history really goes
on, biological evolution exists, animals and men are born, live, and die.
We are actors and not merely passive spectators in the universe. Let
us call this model open universe.
3.7. SPACE AND TIME; COSMOLOGY 123
Now comes the surprise: for our galaxy, eqs. (3.77) and (3.78) give
v
ρ = arth = const. (3.79)
c
and for the “limit sphere”
R = rmax = c t (3.80)
which is the maximum r possible at time t since c = vmax , light velocity
as the maximum possible velocity. Now (3.79) gives
c
ρmax = arth = arth1 = ∞ . (3.81)
c
That means, all ρ’s of galaxies is constant, and space is infinite.
Thus simple mathematical equations transform an expanding uni-
verse of the type of Fig. 3.18 to a static model of type Fig. 3.17 (with
R = ∞).
But the expansion of the universe has been confirmed by observa-
tion! By which observation? By means of the Doppler shift of spectral
lines. If a police car, with its siren howling, approaches me, the sound
appears higher than when it has passed me and happily disappears in
the opposite direction. Thus for an approaching object, the frequency
appears higher than normal, and for a receding object, the frequency
appears lower. For light, red light has a lower frequency than blue light,
so the light of an approaching star would appear shifted towards blue.
What we observe, however, is a red shift which indicates that the stel-
lar object (star, galaxy, quasar) moves away from us, and the greater
the redshift, the greater the speed. This is the observation method by
which the expansion of the universe is measured.
But a frequency change could also be explained by assuming that
in the past, light has had a lower frequency than today. As we observe
more and more distant stars and galaxies, we also go back in time. A
star at a distance of one light year is observed as it was a year ago,
because light took one year to travel from the star to the observer. If
we observe a galaxy which is a million light years away, we see it, not
as it is now, but as it was a million years ago! Hence, if the observed
galaxy has a strong redshift, it means that it is far away in time (since
it is far away in space) and we see it as it was long ago when the clocks
(oscillating atoms or molecules are clocks!) were much slower than they
are now. Mathematically, this means that time is measured in terms of
τ rather than of t, and logically there is no difference between explaining
128 CHAPTER 3. PHYSICS
1000 m + 10 m + 0.1 m + · · · =
1 1 1
= 1000 m 1 + + + +··· =
100 1002 1003
1000 m
= 1000 m(1 + q + q 2 + q 3 + · · ·) =
1−q
1000 m 100000
= 1 = meters . (3.82)
1 − 100 99
loops of one day’s duration: the Ijon Tichy of Monday quarrels with
the Ijon Tichy of Tuesday and that of Wednesday and so forth.
The first major tale of time travel is the classical novel “The Time
Machine” by H.G. Wells published in 1895, where he anticipated several
ideas of the theory of relativity, especially the similarity of time and
space.
He made, however, a philosophically very important error. Taking
the identity of space and time too literally, his hero travels into the
future much more rapidly than the “ordinary rate of time”. Now, what
is the rate of time flow? What does it mean? What is the “speed of
time”? Ordinary “spatial” velocity is measured in meters per second,
m/sec, say (automobilists prefer km per hour). Thus, the “velocity” of
time would be sec/sec, and this is really the case. The velocity of time,
by definition, is
1 sec
=1 , (3.83)
1 sec
the “velocity” of time is dimensionless and always equal to unity ! It
simply cannot be 20 seconds per second. Still less can we travel back-
wards along our own world line, because that would involve negative
“velocities”.
This is an evident but essential distinction between space and time.
“Time travel” can only be along time–like world lines with “speed” equal
to unity.
All about time travel – physics, metaphysics and science fiction – can
be found in the comprehensive work (Nahin 1993) which also contains
an incredible number of references. Very nice is also (Rucker 1984).
Figure 3.21: The left chamber is filled by a gas, the right chamber is
initially empty
Now open the shutter: gas will flow from A to B until the gas is equally
distributed in A and B. If we assume that the gas molecules move
according to the reversible laws of classical mechanics, the opposite
process should also occur, at least from time to time: the whole amount
of gas now in B should flow back into A, and B should be empty, at
least for a moment.
This should happen, but it never happens. The explanation is sta-
tistical: the back flow does happen, but so seldom that there is no
chance whatsoever to watch such an event. The probability of such an
event is so small that it is practically zero.
To see what a time–reversal means for our practical life, run a
motion–picture backwards. Autumn leaves fall upwards, glass splinters
put themselves together to form a new bottle, two cars badly damaged
by a collision separate and become nice undamaged automobiles. The
biographical film of a famous man starts by showing him dead in his
grave. He resurrects from the grave and becomes progressively younger.
He goes to university, then to secondary school, to elementary school,
then to kindergarten. He becomes a baby and loses most of his hair.
His teeth become smaller and smaller and finally vanish. The end of
the film is an ugly little thing in a cradle, admired by his mother and,
somewhat more hesitantly, by his father.
What is the explanation of the arrow of time? I think, ultimately we
just have to accept it, just as we accept ourselves and our surroundings.
Nevertheless, explanations will help get a deeper understanding of this
mysterious phenomenon, which only looks natural because we have got
used to it.
So to speak, we have five main “arrows of time”:
Figure 3.23: Two electrons and a positron or one and the same electron
at three different places?
Figure 3.24: A single electron generates all electrons (•) and positrons (◦)
Introduction
In a poem, addressed to a physicist, Johann Wolfgang von Goethe
ascribes to him the opinion (which he himself rejects):
3.8. INVERSE PROBLEMS 135
g = Af (3.84)
f = A−1 g (3.85)
nature) from sense data g. This is the basic problem of the theory
of knowledge, or epistemology. There is no need to point out its dif-
ficulty, which has kept philosophers busy from Socrates and Plato to
Karl Popper.
Classification. Our mathematical (or pseudo–mathematical) sym-
bolism permits a useful classification:
(1) existence,
(2) uniqueness,
(3) stability.
This means that a solution must exist for arbitrary (within a certain
range) data, that there must be only one solution, and that this solu-
tion must depend continuously on the data. If one or more of these
requirements are violated, then we have an improperly posed, or ill–
posed problem. For a long time it was thought that only properly posed
problems are physically meaningful. In fact, deterministic processes, as
considered in classical mechanics, depend uniquely and continuously on
the initial data — this is the essence of causality — and thus correspond
to properly posed problems.
Only relatively recently it was recognized that there are important
problems that are not properly posed. There is now an extensive liter-
ature on improperly posed problems; cf. (Anger et al. 1993).
In fact, most inverse problems are ill–posed , an extreme example
being epistemology (or philosophy in general). But even some direct
problems are ill–posed, as we know from weather prediction and, more
generally, from chaos theory (sec. 3.2).
Physically, well–posed problems are stable, and ill–posed problems
are unstable . Laplace’s demon, embodying classical determinism and
138 CHAPTER 3. PHYSICS
causality, operates in a stable fashion; St. Peter (if regarded as the saint
responsible for weather) has a rather unstable character.
ing” theory may hold as well within the limits of the accuracy of the
experiment.
For instance, the theory of special relativity reduces to classical me-
chanics for velocities v that are small as compared to the light velocity
c (in other terms, for c → ∞), and quantum theory reduces to classical
mechanics for h → 0 where h is Planck’s constant. Thus, in describing
ordinary (“macroscopic”) physical experiments, classical mechanics is
sufficient, but if we absolutely wish, we may describe the experiments
in a more complicated way also by special relativity or by quantum
mechanics: the results will be practically the same.
Thus experiments must be carefully and ingeniously devised in order
to distinguish between two theories, the so–called crucial experiments.
An example is the well–known Michelson–Morley experiment for spe-
cial relativity. It really permits to “falsify” classical mechanics in an
extreme case, cf. secs. 3.4 and 3.9.
Also the gravitational field of the Earth and of planets may be
correctly described by classical mechanics, if necessary with very small
“relativistic corrections”, although the description by Einstein’s theory
of general relativity is theoretically superior. This is a typical case of
a phenomenon pointed out by the well–known mathematician Henri
Poincaré: several different laws may fit equally well (if necessary with
small corrections). Thus the choice is “conventional” and may be done
by “esthetic” criteria such as “simplicity” or “mathematical elegance”.
This is called conventionalism.
The importance of measuring errors or “noise” in such considera-
tions is evident.
Thus the theory of inverse problems may provide a preliminary first
introduction to the theory of induction, verification and falsification
which will be considered in more detail in the following section. Con-
ventionalism will play a certain role in sec. 6.5.
Our standard Example 3 : Human perception, may also be consid-
ered from the point of view of inverse problems of second kind. Here
the operator A is the apparatus of human perception, called by Konrad
Lorenz (1973) the “mirror” (“Spiegel”) which mirrors our environment
for us. This “mirror” is human perception including neural “hardware”
and “software”, cf. secs. 1.2 and 1.3. Thus the investigation of this
“mirror”, looking with Konrad Lorenz at the “back side of the mirror”
(die Rückseite des Spiegels), may be regarded as an inverse problem of
second kind.
Thus “unreflected” sense perception is related to an inverse problem
of first kind, as we have seen above. On the other hand, physiological
3.8. INVERSE PROBLEMS 141
f = A−1 g . (3.90)
This time we have a well–posed problem since the inverse of a regular
square matrix A exists and is unique.
vT P v =⇒ minimum , (3.91)
the solution is
f = Q AT (A Q AT )−1 g (3.94)
quite similar to (3.92).
The free choice of the matrices P and Q expresses the fact that the
solution is not unique. Once P and Q have been fixed, however, the
solution is unique.
The generalized inverse A−1 is the product of matrices preceding g
at the right–hand sides of (3.92) and (3.94), respectively.
If you are not able to follow the mathematical argument, don’t
worry. The main result is that, by assuming definite matrices P and Q,
even overdetermined and underdetermined linear systems get a unique
solution.
Obviously this model is quite flexible because P and Q can be cho-
sen rather arbitrarily. As a matter of fact it is not claimed that the un-
derdetermined quantum problem or Whitehead’s overdetermined prob-
lem should be treated in this way. It is always good to know, however,
that well–defined models for the solution of overdetermined and under-
determined systems exist and can be used if necessity arises.
The present considerations also show the basic role of probability for
the solution of ill–posed inverse problems. Since a unique solution does
not exist, one tries to find the “best” solution on the base of statistical
considerations. Both least–squares principles (3.91) and (3.93) have
a statistical background: the “weight matrix” P and the “covariance
3.9. INDUCTION, VERIFICATION, FALSIFICATION 145
Induction
Frequently, induction is considered the inverse of deduction. Deduc-
tion proceeds from the general to the particular, using a general law to
compute particular observable quantities which then may be compared
with actual observations. Induction is said to proceed from the partic-
ular to the general, using particular observed data to derive the general
law. Thus induction is the inverse problem, in the sense of sec. 3.8, of
deduction.
This definition of induction is acceptable if we keep in mind that it
has a logical status completely different from that of deduction. Deduc-
tion is a precisely and uniquely defined, straightforward logical process
which can be formalized in terms of symbolic logic (sec. 2.1) and may
146 CHAPTER 3. PHYSICS
The problem of induction has been one of the most famous and
most difficult problems of philosophy, from David Hume (1711–1776)
to the present day. Let us start with some simple examples.
(1) Succession of day and night. This has been observed since
mankind came into existence, and there was never a single exception
(see also sec. 3.8). Can we conclude that tomorrow the sun will shine
again — at least above the clouds? Pragmatically we all believe that
there will be another day, but this cannot be proved logically. Induction
is not a purely logical problem. If logical procedures such as deduction
are called analytic, induction is not analytic. It is a physical problem:
there will be no tomorrow if the Earth or the Sun explode during the
night, or if the Earth has been destroyed by the impact of a huge me-
teorite. But still we may consider that with high probability there will
be another day.
(2) All swans are white. Let us assume, for the sake of argument,
that, so far, only white swans have been observed. Can we say (a) that
the next observed swan will also be white and (b) that all swans are
white? Obviously we can expect event (a) to occur with much higher
probability than the general law (b) to be true. Even if zoology claimed
that all swans are white (which it does not), a black swan could still
3.9. INDUCTION, VERIFICATION, FALSIFICATION 147
occur: a student might have painted the swan black in order to fool his
professor.
It is sometimes said that induction works if there is a certain uni-
formity of nature. This certainly applies to Example 1: the laws of
Earth rotation guarantee the succession of day and night if there is no
perturbation by a collision with a large meteorite or by an explosion as
mentioned above. But will these laws also hold tomorrow?
(3) Russell’s chicken. We quote from (Russell 1912, Chapter VI):
“Domestic animals expect food when they see the person who usually
feeds them . . . The man who has fed the chicken every day throughout
its life at last wrings its neck instead, showing that more refined views
as to the uniformity of nature would have been useful to the chicken.”
However, expectations govern our daily life as well as science. When
I come home and see and smell a nice meal, knowing that my wife is a
good cook, I expect the food to be good and healthy. When a physics
professor prepares a particularly showy experiment, knowing that he is
a skilful experimenter and trusting the laws of physics, he expects the
show to be successful. When a botanist plants the bulb of a tulip she
expects that, under normal circumstances, she will get the appropriate
flower. If a pianist starts a concert with Beethoven’s Appassionata and
touches the keys accordingly, he and his listeners expect to hear the
magnificent sounds of the sonata.
Still, one can never be sure: some jealous colleague may have put
the piano out of order right before the concert, etc.
But usually our, often unconscious, trust in the uniformity of nature
or its lawful behavior is justified, especially if we use experience and
common sense. If we discover a certain regularity and we expect an
underlying general law, then we may be justified to assume that the
regularity will persist in the near future, and if the regularity continues,
we may reasonably assume that we have discovered the corresponding
general law. Every new experiment will confirm the law and increase
its probability. This may be called induction by analogy.
Bayes’ theorem. Let H denote a hypothesis, and p(H) its initial “a
priori ” probability, or prior probability . Let E denote an event that
confirms the hypothesis. Then
p(E|H)
p(H|E) = p(H) = α p(H) . (3.95)
p(E)
This is the simplest form of Bayes’ theorem, cf. (Cohen 1989, p. 68).
It gives the “a posteriori ” probability, or posterior probability , of H
148 CHAPTER 3. PHYSICS
l = 124.327 m ± 2 mm .
falsify it. The more experiments are performed, the more the effect of
measuring errors will be reduced.
It is fundamental that physicists and natural philosophers put much
emphasis on Heisenberg’s uncertainty principle, but ordinary measuring
errors, which occur much more frequently and may be much larger, also
deserve their attention; cf. (Jeffreys 1961, pp. 13–14). Jeffreys’ books
merit particular respect because they were written by a scientist of
enormous experience with actual “dirty” data.
Many extreme logical or philosophical conclusions do not apply to
our real “fuzzy” world because of their very subtlety: they are used
as razors, not for splitting hairs (which would be appropriate) but for
cutting trees.
The results of deduction are logically true if the process of deduction
is done correctly. The results of induction, including verified or, better,
not–yet–falsified, theories, can, at best, be probable or acceptable on
a hypothetic basis. No physicist will consider relativity or quantum
theory “absolutely true”, in the same sense as 2+2=4. He regards them
as excellent and unsurpassed working tools, even as correct, but only in
the sense of exceptionally good approximations; cf. also sec. 6.5. In this
way, also Newtonian mechanics remains “correct”: for small velocities
and phenomena above the quantum level.
Crucial experiments. Nevertheless, there are crucial experiments
which really permit to decide between two theories or hypotheses. One
of the most famous crucial experiments is the Michelson–Morley ex-
periment which shows that the velocity of light is the same along all
directions on the moving Earth. This is incompatible with classical me-
chanics, eq. (3.36) on p. 91, but perfectly compatible with the special
theory of relativity.
It is instructive to study this case by means of Bayes’ formula. Gen-
eralizing (3.95) to the case of two competing hypotheses H1 (classical
mechanics ) and H2 (relativity ), we have
p(Hi )p(C|Hi )
p(Hi |C) = (3.100)
p(H1 )p(C|H1 ) + p(H2 )p(C|H2 )
= A p(Hi )p(C|Hi ) (3.101)
p(C|H1 ) = . (3.103)
so that
p(H2 ) = 0.1 (3.110)
in view of (3.106). Then (3.100) gives
0.9 9 .
p(H1 |C) = = = 9 , (3.111)
0.9 + 0.1(1 − δ) 1 − δ + 9
0.1(1 − δ) 1−δ .
p(H2 |C) = = = 1 − 9 . (3.112)
0.9 + 0.1(1 − δ) 1 − δ + 9
Thus, the crucial experiment has made to drop the probability of clas-
sical mechanics from 0.9 to 9. If = 10−6 (the experiment is very
reliable), then
.
p(H1 ) = 10−5 , (3.113)
.
p(H2 ) = 1 − 10−5 , (3.114)
which shows that relativity is confirmed also in this case: p(H2 ) has
risen from 0.1 to almost 1!
In fact, the assumption p(H2 ) = 0.1 is not too high because, even
before the Michelson–Morley experiment it was known that the Lorentz
transformation (3.38) on p. 92 holds for electrodynamics: Maxwell’s
equations are invariant with respect to (3.38); this was precisely what
Lorentz showed. So in view of the universality of physics, there was a
certain a priori probability that H2 would hold also for mechanics.
We see that basically the same result is obtained for very different
prior probabilities. This indicates that the choice of the prior probabil-
ities is not very essential.
We shall have much more to say on the laws of nature in following
sections. There is a long way from “Russell’s chicken” to the “eter-
nal inexorable laws of nature” of romantic poets and happy–minded
scientists, and a still longer way to “Schrödinger’s cat” . . .
The unreasonable effectiveness of mathematics. Already at this
point, however, we mention, for the first time in this book, a fact that
has intrigued physicists from Kepler to Einstein. The great quantum
physicist Eugene Wigner has called it the “unreasonable effectiveness
of mathematics in the natural sciences”. Penrose (1989, p. 430) puts it
as follows (“SUPERB” theories are, e.g., special and general relativity
and quantum mechanics, cf. p. 260):
It is hard for me to believe, as some have tried to maintain, that such
SUPERB theories could have arisen merely by some random natural selection
3.10. THE STRUCTURE OF SCIENTIFIC REVOLUTIONS 155
of ideas leaving only the good ones as survivors. The good ones are simply
much too good simply to be the survivors of ideas that have arisen in that
random way. There must, instead, be some deep underlying reason for the
accord between mathematics and physics, i.e. between Plato’s world and the
physical world.
Summary. Induction does exist. In simple cases, induction by anal-
ogy can be used with some care: Bayes’ theorem may help, also in the
estimation of parameters for simple laws given by a function contain-
ing several parameters. More general laws such as relativity require
the creative mind of a great scientist; verification or falsification by
experiment are necessary but can be expected almost automatically
to be performed by the scientific community. Crucial experiments are
especially important.
Additional reading. A nice introduction is the chapter on induction
in (Russell 1912). If you read what Russell (1948), Carnap (1950, 1966),
Popper (1977; Miller 1985), and Jeffreys (1961, 1973: if you don’t need
them, disregard the formulas but do read the text) have to say about
induction, and if you conclude with (Cohen 1989), then you should
know almost all the relevant present views on the topic.
motion and the mechanics of Newton, Legendre, and Laplace were the
consequences.
Another revolutionary change of paradigm was biological evolution.
The static system of botanical and zoological classification of Carl von
Linné (1707–1778) was made into a dynamic theory of evolution by
Jean Lamarck (1744–1829) and Charles Darwin (1809–1882).
Immanuel Kant (1724–1804) performed a “Copernican revolution”
in philosophy, emphasizing the role of the subject. In this way, he
founded the great school of German idealism (Fichte, Schelling, Hegel),
but he also influenced the contemporary philosophy of science.
The logical discoveries of George Boole (1815–1864), Gottlob Frege
(1848–1925), Giuseppe Peano (1858–1932) and Bertrand Russell (1872–
1970) lead to analytical philosophy and the modern theory of science,
or philosophy of science.
In physics, of course, we have the new paradigms of relativity and
quantum theory, and most recently, chaos theory.
The recent development in biochemistry and molecular biology
started modern genetics. Reductionism (life can be reduced to chem-
istry and physics) is not a logical consequence of these developments,
but is rather generally accepted at least as a working hypothesis.
Cybernetics, system theory, catastrophe theory, complexity theory,
and synergetics also have influenced modern scientific thinking, closely
related to the advent of electronic computers. These concepts will be
explained later (sec. 4.2).
In geosciences, we now have the paradigm of plate tectonics. The
Earth’s surface consists of a number of continental plates which move
with a speed on the order of 5 cm per year. Colliding with each other,
they pile up mountain chains such as the Rocky Mountains and the
Andes, but also the Himalayas and the Alps, accompanied by earth-
quakes and volcanism. It started with Alfred Wegener’s continental
drift (sec. 3.9) published in 1915, but it was generally recognized only
in the sixties.
A scientific revolution does not occur only because what Kuhn calls
“normal science” has been proved wrong or, in Popper’s terminology,
“falsified”. The old epicycle theory of planetary motion by Ptolemy
(2nd century A.D.) could have easily been adapted to the increasing
measuring accuracy by adding one more epicycle or two; it simply grew
too complicated to retain credibility. Kepler was motivated not only by
his belief in the simplicity and harmony of the world (which, in some
way or other, is shared by modern scientists as well) but also by mys-
tical speculations: he related the planetary orbits to the five Platonic
3.10. THE STRUCTURE OF SCIENTIFIC REVOLUTIONS 157
In sec. 1.1 we have seen that, in man and animals, the hypotha-
lamus serves, among other functions, as a thermostat regulating body
temperature in order to maintain this temperature at a constant level
within very narrow bounds. Feedback and regulation mechanism play
a fundamental though largely unconscious role in the working of the
human body, in movements such as walking, and may even be respon-
sible for what we call the action of mind on our body, such as reaching
for a book which we need when working at a problem. Such an action
from a higher level (thinking or willing) to a lower level (bodily move-
ment) is called downward causation. Also the activity of a computer
when we feed in a program and data, or instruct it to display a certain
information on the screen, are examples of downward causation (caus-
ing activity on the “lower” hardware level by providing input on the
“higher” software level).
An essential concept is feedback . A teacher gives a course, and
the students listen with more or less attention. If the teacher cares
159
160 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
course of the automobile is again fed back to the driver, deviations are
observed and corrected, etc.
This example also helps introduce two additional concepts. The
automobile acts as some kind of servomechanism for man, helping
him/her to cover large distances much faster than by walking. Ser-
vomechanisms in the narrower sense frequently replace human move-
ment and observation of deviations (from a goal) by automatic func-
tions. This happens, for instance, with a stabilized platform: deviations
from a fixed orientation are measured by gyroscopic sensors and acti-
vate servomotors which restore the desired fixed orientation. Modern
airplanes are frequently equipped with autopilots which, using inertial,
radar, altimeter, etc. information will relieve the pilot from consider-
able purely routine work, frequently do certain specific and well–defined
tasks more precisely and accurately than the human pilot could do.
A second important concept is the name, cybernetics or kybernetics,
for a whole new discipline: control and communication in man, animals,
and machines. This name was introduced in 1947 by Norbert Wiener,
cf. (Wiener 1961, p. 11). It is derived from the Greek word κυβερνη0 τ ης
(kybernetes ), which means steersman or pilot.
It shows that machines exist which are designed to perform a certain
task (Aristotle’s “final cause ”) through appropriate mechanisms oper-
ating according to the usual laws of causality in the physical sense (Aris-
totle’s “efficient cause ”). Thus the gap between physics (believed to be
subject to usual causality) and biology (where “goals”, “aims”, “pur-
poses” and similar “final causes” seem to govern the behavior of animals
and even the course of evolution) has at least partially been bridged.
Earlier, a “vital force” has been considered necessary to explain “fi-
nalistic” animal behavior, but, according to Wiener (1961, p. 44): “In
fact, the whole mechanist–vitalist controversy has been relegated to the
limbo of badly posed questions”.
Other examples of servomechanisms operated by man are bulldozers
and other excavating machines, replacing shovels and similar tools. In
fact, such tools and machines can be regarded as extensions of our body
to perform certain tasks much better than using our hands only.
Now it is basic that our hands, feet and other bodily organs may
also be considered instruments or tools (“servomechanisms”) directed
by our mind (or by the neural activity in our brains, if you prefer) to
perform certain operations, for instance reaching for a glass of water
and bringing it to our mouth. This operation is constantly controlled
by visual and other feedback, as we easily recognize if we try to perform
it in darkness.
162 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
shoots the mark, and then oscillates back and forth with rapidly in-
creasing amplitude, so that a person in the room would not even have
the choice between being heated and frozen to death: this would hap-
pen in rapid alternation. This corresponds to positive feedback which
clearly does not provide the desired stabilization of temperature, quite
similar to the cerebellar tremor mentioned above.
Mathematically, this is evident: negative feedback corresponds to
exponentially decaying cosine (or sine) functions
disc, that is, independently of the signal and its (zero) power. Input
(programs and data) contained in magnetic discs or tapes may cause a
computer to perform rapid mathematical and logical operations which
require considerable electric power (which is, however, again indepen-
dent of the size of the numbers in a particular kind of input) and its
degradation to thermic energy (remember the cooling towers of modern
supercomputers); and “powerless” books such as the Bible, the Koran
or, to a lesser extent, the works of Marx, Engels, and Lenin, have com-
pletely changed the face of the Earth.
This is possible through the use of some “reading equipment” such
as laser or magnetic scanners, or people reading those books, which
convert the “written” input into small electrical signals (including the
eye converting the optical information into nerve impulses!), which are
then greatly amplified and produce the results just described.
Electronic computation, hardware being activated by software, is
a beautiful example of downward causation: information on a higher
level (software) causes action on a lower level (hardware). Note that
this hardware motion is fully governed by physical laws, in particular
the laws of electronics! “Software laws” activate the appropriate “hard-
ware” (physical) laws; “final causes” act through activating the appro-
priate “efficient causes”. Software provides initial conditions causing
the computer to start working, as well as boundary conditions which
regulate its work during the computation. Generally, the term “bound-
ary conditions” is used as a compact synonym for both initial conditions
and boundary conditions in the narrower sense.
The interaction of mind with matter may well be of this kind. Ob-
jections that mind does not possess physical energy and cannot there-
fore act on matter, lose their force if we compare the action of mind
on matter to the action of software on hardware in a computer. For
more details cf. (Popper and Eccles 1977, Chapter E7), (Eccles 1994),
(Globus et al. 1976), and (Margenau 1984). Haken (1981, p. 196) dis-
cusses interaction at various levels.
A thorough discussion of mind–matter interaction can only be given
by quantum theory since quantum effects are believed to play a role
in neural activity. A comprehensive reference is (Stapp 1993); also
(Lockwood 1989) and (Penrose 1989) provide valuable insight. One
now even speaks of “quantum computers”!
Finalism in physics. Even in classical physics (mechanics and op-
tics) there are principles that seem to express some purpose and are
thus “final causes” in the sense of Aristotle. E.g., we have Fermat’s
principle: light moves from point A to point B along a path for which
4.1. FEEDBACK, REGULATION, DOWNWARD CAUSATION 165
Does not this remind you of a “survival of the fittest” in the sense
of Darwin’s evolution? And is it too c extravagant to regard Darwin’s
principle (survival of the fittest ) as an optimum principle which is some
kind of biological analogue to the finalistic principles of Fermat, Euler,
and Maupertuis in physics?
4.2 Self–organization
Self–constructibility is an emergent
property of a complex system.
Sir Alan Cottrell
A uniform wind will first slightly tilt the plane ocean surface
(Fig. 4.4 (b)). This tilted plane is unstable and hence breaks down.
Then the wave motion, which is a stable motion determined by pa-
rameters of wind and water and by atmospheric pressure, takes over.
What remains undetermined and “random”, however, is the position
of the wave crests (the “phase”): the same factors might as well have
formed a wave translated to the right (as shown by the broken line) or
to the left. But once the symmetry of the smooth plane surface has
been broken, the phase is determined.
Another well–known phenomenon are convection cells formed by
heating the lower surface of a fluid layer (Bénard problem), cf. Fig. 4.5.
The latter is clearly relevant for convection in the Earth’s viscous
mantle, causing the movement of continental plates in plate tectonics.
Also regular rolling cloud patterns, frequently visible in an otherwise
clear sky, are caused in a similar way.
Chemistry and biology. Self–organization also occurs in chemistry,
168 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
Figure 4.5: Convection (a) in a plane slab and (b) in a spherical shell
Figure 4.7: Interaction between the two cells leads to two new config-
urations of equilibrium)
troversial. We mention now only one of the simplest models for cell
differentiation due to the famous mathematical logician A.M. Turing
(Haken 1981, p. 104). Consider two separated identical cells (Fig. 4.6),
in which a substance A is contained in a certain equilibrium concen-
tration. Now we admit an exchange of substances between the two
cells. This may make the initial concentration of A unstable, and a
new equilibrium is reached, in which either cell 1 or cell 2 gets a higher
concentration, both possibilities being equally probable (Fig. 4.7).
Figure 4.8: Potential energy and symmetry breaking. The cases (a),
(b), (c) correspond to Figures 4.6 and 4.7
S = k log W (4.3)
dS
≥0 . (4.4)
dt
The standard example is a scientist’s desk on which papers and books
are accumulating. With normal use, the books and papers become
increasingly disordered (until the scientist gets disgusted and cleans
the desk). The disorder W and hence the entropy S increase.
Another example: if we take a glass of hot water and a glass of cold
water, and pour both into a larger glass, the temperature of the mixture
very soon settles at a uniform lukewarm state. The mixture is less
ordered than the originally separated amounts of cold and warm water
whose temperatures were different, exhibiting more structure than the
resulting uniform mixture. Again S has increased.
A similar case is the dissolution of a piece of sugar in a glass of
water: the sugar solution possesses higher entropy (less structure) than
the system consisting of the piece of sugar and the glass of water.
174 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
I = −∆S , ∆S = S − S0 . (4.5)
Genetic information
Life clearly also runs counter to the Second Law, both in an indi-
vidual living being and in the course of biological evolution. Think of a
newborn baby. It grows up, learns an enormous amount of things, and
becomes a highly developed man or woman. In this course, the “or-
der” or “organization” increases, entropy (disorder) decreases. Only
after death, the Second Law takes over again: the body decomposes,
disorder and hence entropy increase again.
Similarly, biological evolution has run counter to the Second Law:
animals and plants become more and more complex and highly orga-
nized, from amoebae and bacteria to roses and human beings.
All this is far from being fully understood. Certain facts appear
sure, however. First of all, the fact of evolution is not contested by any
serious scientist. What is controversial is whether evolution proceeds in
a completely random way, according to chance mutations and Darwin’s
“survival of the fittest”, or whether it follows more or less an existing
“cosmic blueprint” (Davies 1988). Probably the truth lies between
these extremes. As Penrose (1989, p. 416) has put it:
To my way of thinking, there is still something mysterious about evolution,
with its apparent “groping” towards some future purpose. Things at least
176 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
seem to organize themselves somewhat better than they “ought” to, just on
the basis of blind–chance evolution and natural selection.
A: adenine,
G: guanine,
C: cytosine, (4.7)
T: thymine.
U: uracil. (4.8)
DNA: A G C T
⇓ ⇓ ⇓ ⇓ ⇓ (4.9)
RNA: U C G A
gene =⇒ protein .
(1) Metabolism: the animal eats food (other animals, plants) which
furnish not only energy but also (negative) entropy, i.e., infor-
mation, since the food already has a rich structure. The body
uses this information for its own organization and releases the
food after having extracted energy and information, in a highly
degraded (high–entropy) state into its environment (Schrödinger
1944).
A “cycle” may copy part of a nucleic acid (e.g. RNA) and produce a
corresponding protein acting as an enzyme for increasing the activity
of the next cycle. Thus the action of the hypercycle is one of mutual
support of the work of the individual cycles. If Cycle 1 increases the
production of its enzymes (Enzyme 1), this enzyme will stimulate the
activity of Cycle 2, whose Enzyme 2 stimulates Cycle 3, etc.
Thus the cycles mutually help to increase their activity. It can
be shown that also the stability is mutually reinforced. The cycles
operate in an autocatalytic way, and also “assist each other” in a cross–
catalytic manner. The hypercycle is a model for cooperation rather than
competition in the organic world; cf. also (Jantsch 1980) and (Kauffman
1993). We also recognize a feedback acting within the hypercycle. As a
matter of fact, the hypercycle takes energy and “raw materials” from
the environment.
The great physicist John Archibald Wheeler was so fascinated by
the hypercycle that he spoke of the “life machine” of Manfred Eigen
(Wheeler 1994, p. 180)!
How the enzymes act on the structure of the cells and contribute to
their self–organization is a complicated and controversial issue. Cat-
alytic processes and hypercycles seem to play a role also here.
It thus appears that the organization of a living organism is of
the conductor–orchestra rather than of the “democratic” string quartet
type (see beginning of sec. 4.2).
The human genome seems to play the role of the over–all conductor,
but the musicians must be highly qualified to fill in the details of the
musical piece by improvisation (self–organization).
The genome thus determines the general features of the person, e.g.
intelligence on the one hand and genetic diseases on the other hand.
We are not, however, slaves of our genes, but can do much to make best
use of our capabilities and reduce our genetic limitations. We cannot
just shed our responsibilities and blame everything on our genes.
It is tempting to regard the human genome as the material basis
of the self–identity of every human person. This may well be, but the
material aspect is not the only one.
The stability of a personal organism is thus partly static. In part,
stability in nonlinear systems may also be dynamic, that of a fixed
point or point attractor (Fig. 4.11). (There are also more complicated
“strange attractors” such as the Lorenz attractor well known from me-
teorology, cf. (Abraham and Shaw 1992), (Briggs 1992), (Lorenz 1993).)
4.3. ENTROPY, INFORMATION, EVOLUTION 181
dS ≥ 0 . (4.10)
– they are open, i.e., they can react with their environment;
Still, all these questions are extremely difficult, and the partial an-
swers given so far are rather controversial. Fortunately, many of the
pioneers have written excellent popular books: (Eigen 1987), (Eigen
and Winkler 1975), (Haken 1981), (Kauffman 1993), (Prigogine and
Stengers 1984), (Nicolis and Prigogine 1989), (Thom 1975), so that
the reader can judge for himself. The extremes are marked by (Teil-
hard de Chardin 1955) and (Monod 1970): evolution according to a
“holistic” plan, and completely random and mechanistic evolution. An
excellent comparative treatment of all these various tendences and di-
rections of research is found in (Moser 1989, Chapter 4). Balanced and
concise presentations are given in (Davies 1988) and (Mayer–Kuckuk
1989, Chapter 9). The little pioneering book (Schrödinger 1944) has
retained its charm and freshness for over half a decade and is my fa-
vorite. Off the beaten track are the passages on the genetic code in
(Hofstadter 1979) and in (Cohen and Stewart 1994), other favorites of
mine. The presentation of Sheldrake (1981) is unconventional but very
readable and interesting. An excellent general introduction into evolu-
tion is (Edey and Johanson 1989), combining readability with a high
level. If you look for open and controversial problems, consult (Duncan
and Weston–Smith 1977). A monograph representing the current state
of research on complexity is (Zurek 1990), a publication of the Santa Fe
Institute, whose work is excitingly described by Waldrop (1992; gen-
eral introduction) and Lewin (1992; emphasis on biology). Cohen and
Stewart (1994) show how complex evolution really is.
nature, concepts are always inexact. In physics, the way out is to make
the experimental arrangement as exact as possible, and then, by a leap
of faith, idealize the situation and postulate that the measured values
obtained in this way are absolutely accurate. This may or may not
work.
One of the most accurately known physical constants is the speed
of light c in vacuum. Its value is
c = 299 792 458 m s−1 , (4.13)
that is, meters per second. The measuring accuracy is so high that the
last three numbers might be 459, but not, for instance, 464. We say that
c has been measured to an accuracy of ±1 m s−1 . This is the standard
error or r.m.s. (root mean square) error σ which is defined statistically
and known to everyone who has ever handled empirical data. Now,
σ = ±1 m s−1 does not mean that only 457 or 459 may be possible
outcomes of measurement; even an error of 3σ (possible outcomes 455
or 461) may occur, but with rather small probability. (Even larger
deviations are theoretically possible, but extremely improbable.)
σ 1 m s−1 .
= = 3 × 10−9 (dimensionless!) . (4.15)
c 299 792 458 m s −1
This accuracy is considered almost the highest accuracy that can con-
ceivably be achieved. So why not consider it “absolutely accurate”?
The chief limitation of the accuracy of
length
c= (4.16)
time
is the limited accuracy of measuring length or distance, that is, the
definition of the meter . The limited accuracy is due to the difficulties
mentioned at the beginning of sec. 2.4 regarding point definition and
distance measurement. The meter was defined as a certain multiple
of the wavelength of monochromatic light of a certain frequency, and
it was simply not possible to increase the accuracy of the frequency
definition. On the other hand, time definition by means of an atomic
clock is considerably more accurate.
So, some twenty years ago, it was in fact decided to give up the
old definition of the meter and define length as time × c. Thus the
defining constants were taken the second and the velocity of light, with
c according to (4.13) errorless and fixed once and for all. The meter
is now a derived quantity: it is the distance covered by light in c−1
seconds,
1
1m = ˆ s . (4.17)
c
Thus, in certain rare cases, it is permitted to consider an accurately
measured value as absolutely errorless, but then it may be necessary
to redefine other quantities in such a way that logical (and numerical!)
contradictions are avoided.
It would not be permissible to regard all three measured angles of
a triangle as errorless, since this results in a numerical contradiction:
the sum of the three angles in a triangle must be 180◦ or π:
α+β+γ =π , (4.18)
Thus the laws governing the behavior of living organisms may well
be regarded as some kind of software laws describing biological complex-
ity, following again (Davies 1988, p. 142). (This book is my favorite
reference for the present problems.)
Thus the “biological laws” define the initial or boundary conditions
for the work of our “biological” systems.
This comparison between animal and computer, imperfect as it is,
shows clearly not only the possibility, but even the necessity of laws
other than physics to describe the activity of a computer and, a for-
teriori , of a living organism. Note that in sec. 4.1 we have used the
computer in a similar way, as a model for mind–brain interaction.
The old fight between vitalists and reductionists is thus raised to an
objective and unemotional level. Any molecular biologist who speaks
of “vital forces” would commit scientific suicide, but if he (she) speaks
of “software laws”, nobody will pay attention to the fact that, possibly,
the same thing is expressed in two different forms.
Software contains information. Thus information must play a deci-
sive role for biological systems. Perhaps oversimplifying, we may thus
say
life = matter + information (4.19)
(Küppers 1987, p. 17). This is probably true, but what is information?
Unfortunately it is by no means a clearly and unambiguously defined
concept. The relation between entropy and information outlined in
sec. 4.3 is rather generally accepted, but neither gives a complete defini-
tion nor is understood by all scientists in the same way, cf. (Weizsäcker
1985, Chapter 5). It is one of the basic paradoxes of science and phi-
losophy that the most fundamental concepts such as matter, mind, or
information are so ill–defined.
By regarding a living organism as a complex system governed by
the laws of physics and “software laws” containing information we have
given the problem a simple (though probably oversimplified) structure
which may serve as a basis for more detailed investigations. The related
concept of self–organization (sec. 4.2) may serve for similar purposes.
Using the terminology of Niels Bohr, physical and biological laws
(if you don’t like this term, speak of software laws) are complemen-
tary (very much in the sense of complementary subspaces which are
orthogonal to each other, cf. Fig. 2.19 on p. 66 and Fig. 6.1 on p. 246).
This has been drastically formulated as follows (“Bohr’s para-
doxon ”): in order to determine whether a cat is fully governed by
physical laws, it is not sufficient to determine its weight or its bod-
ily temperature. One must use an X–ray equipment, which has to be
4.5. COMPLEXITY AND REDUCTIONISM 189
On reductionism
As we have said at the beginning of this section, it seems that the
laws of physics hold also for living organisms. Let us consider this
question more closely. Physical laws have an inherent inaccuracy as
we shall see in sec. 6.5. Experiments so far have shown that they
are satisfied in plants and animals within the measuring accuracy; this
assertion we shall call moderate reductionism. Furthermore, as we have
just seen, more precise physical measurements may interfere with life
(of a cat, say).
So the question whether physical laws are “really” and “absolutely”
the same in living and nonliving matter may well be meaningless. Nev-
ertheless, this has been frequently asserted, so we shall call it strong
reductionism. If someone asserts this, we can immediately retort: re-
duction to which physical laws? The “real” ones if they exist? The
laws found in our books of physics? But does Nature, or God, read
those books? Questions are becoming cynical, so let us stop them.
Strong reductionism is asserted mainly by biologists. Biologists of-
ten have a much stronger faith in physics than physicists themselves.
(Similarly, physicists have a much stronger faith in mathematics than
mathematicians, cf. the end of sec. 2.3. This seems to be due to inside
knowledge: it is said that Napoleon’s valet had much less respect of him
than most other people.) The book (Küppers 1987) contains highly in-
teresting articles by great contemporary physicists on this topic which
are very relevant to our discussion. Great physicists are usually rather
190 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
that the watch is nothing but cogwheels, springs, screws, pinions, and
other simple mechanical parts. The father is upset and tells the child
to put these “simple” mechanical parts together again to obtain the
original watch. The child is unable to do this. “Taking apart” was
a reductionist procedure, and “putting together” is a constructionist
problem which is much more difficult: the inverse problem is usually
essentially more difficult than the direct procedure.
In 1972 Philip Anderson asserted that “the reductionist procedure
does not by any means imply a ‘constructionist’ one: The ability to
reduce everything to simple fundamental laws does not imply the ability
to start from these laws and reconstruct the universe” (quoted after S.S.
Schweber, Physics Today, November 1993, p. 36).
H. Primas formulates it even more strongly: “Every machine relies
for its operation on the laws of physics and chemistry, but the machine’s
design is a higher–order principle. In this sense machines are irreducible
to physics” (quoted from Moser 1989, p. 137).
A rather generally prevailing compromise might be formulated as
follows. Physical laws hold throughout nature, living and lifeless alike.
What is characteristic for biology are not the laws but the boundary
conditions, which are responsible for the complexity of living organ-
isms, the physical laws being quite simple. You may call these bound-
ary conditions “biotonic” laws or “software laws”, depending on your
preferred way of thinking. Even the freedom of the will may be related
to boundary (or initial) conditions, cf. sec. 6.4 and Fig. 6.1.
192 CHAPTER 4. SYSTEMS, INFORMATION, EVOLUTION
Quine’s model
Some positivists may tend to consider logic as “true in all possible
worlds” (Leibniz), mathematics is reduced to logic, physics is expressed
in mathematical laws and hence reduced to mathematics, chemistry is
reduced to physics, and biology is reduced to physics and chemistry,
and hence to physics. This throughgoing reductionism” has been seen
to be endangered by logical paradoxes, Gödel’s theorem, Heisenberg’s
uncertainty relations, and by inaccuracies of both measurements and
theories in general, not to forget self–reference (Fig. 4.13).
Quine (1961, p. 42) has given a somewhat different and probably
more realistic picture (Fig. 4.14).
Thus, according to Quine, even logic and mathematics are not given
a priori once and for all, as gifts from heaven, perfect and changeless.
In fact, Weizsäcker (1985, p. 313–319) has worked on a “quantum
logic” in order to improve our understanding of quantum theory (cf.
end of sec. 2.6), and also Weyl’s opinion (the sentence following the
quotation at the end of sec. 2.3) fits well into this scheme:
A truly realistic mathematics should be conceived, in line with physics, as a
branch of the theoretical construction of the one real world, and we should
adopt the same sober and cautious attitude toward hypothetic extensions of
its foundations as is exhibited by physics.
Part C
Philosophy
Chapter 5
Philosophy for scientists
197
198 CHAPTER 5. PHILOSOPHY FOR SCIENTISTS
Hence what is primary is our mind , and we do not see the house, we
see some mental image, which could also be produced in another way,
by a motion picture or by a television program. Thus this leads to
Idealism. The primary perceptions are affections of the mind, sense
data, from which by thinking we reconstruct an external world which
may not even exist in the way it appears.
Now, however, we ask: where do these sense data come from? Unless
we suffer from an illusion, sense data must come from the external world
(of which also cinemas and television sets are part).
After naive realism as thesis and idealism as antithesis, we thus
arrive at a synthesis:
Critical realism, or scientific realism. Through our senses we get
information on the external world, which is certainly partial and im-
perfect (we do not see ultraviolet and infrared light, for instance) but
is essentially true. If this information is not true (illusions, cinema) we
at least understand why it is not true: science (physics, evolutionary
theory of knowledge, psychology, and psychiatry) tells us about this.
This is the scientific world picture, which essentially gives the same
results as naive realism, but in a refined way. It is maintained by
practically all scientists and philosophers, even by those who do not
admit it.
Two remarks are in order, however. This dialectic game could be
continued, getting a refined idealism, an even more sophisticated re-
alism, etc. (This is beautifully shown in Fichte’s Wissenschaftslehre
of 1904, cf. sec. 5.3.) We could stop at any stage of realism or ide-
alism, and many philosophers did stop at an idealistic position. We
shall be satisfied, however, with scientific realism, at least as a working
hypothesis.
The second remark is the following. Neither realism nor idealism,
at any stage, can be proved or refuted. If we take an extreme idealistic
position, we arrive at
Solipsism. Only my own ideas are real, my sense perceptions are
nothing but illusions, I live exactly like in a dream. (“Solus ipse ”
means: only I exist, all the rest is illusion.)
It is impossible to refute this position. (When I heard this for the
first time at the age of 16 years from our excellent philosophy teacher,
I wanted to test it. Waking up one morning, I decided that “in reality”
I continued to sleep and only dreamed that I had woken up. I per-
sisted in this attitude, behaving normally but considering everything a
dream: going to school, studying, taking examinations, playing etc. It
worked perfectly and was absolutely self–consistent. After a few weeks,
5.1. REALISM, IDEALISM AND DUALISM 199
d2 x
m =F , (5.1)
dt2
200 CHAPTER 5. PHILOSOPHY FOR SCIENTISTS
of the two “substances”. Thus there are two other possibilities: either
God has created both clocks so perfectly that they always remain abso-
lutely synchronized, or God is constantly watching them to keep them
continuously synchronized, immediately correcting any deviation. This
is a fine example of where “sharp” or “exact” philosophical reasoning
can lead to. (God as the “Universal Metaphysical Problem Solver”, cf.
sec. 1.2.)
The old philosophical fallacy was the belief that one has to start
from “clear”, “true”, “evident” and “precise” principles (“premises”)
and then follow sharp logical reasoning wherever it would lead you.
Only a slight error in the premises might lead to large errors in the
result (“ill–posed problem”, see below). Whitehead (1925, p. 75) calls
this unjustified sharpening of concepts (“noninteracting substances”!)
the “fallacy of misplaced concreteness ”. Ivan Supek (quoted in the
introduction) is more direct:
The old metaphysicians got caught in the trap of some absolutized words
and concepts, categories or principles; and wishing to construct a consistent
[philosophical] system, they locked themselves into a lifeless ivory tower.
all data that are coming from a given object). This definition, seem-
ingly circular, can be made logically acceptable, breaking the “vicious
circle”, in much the same way as in Russell’s definition of natural num-
bers (positive integers, cf. sec. 2.1).
Materialism and idealism both belong to the category of monism.
Therefore both Hegel (a dialectical idealist) and the dialectic materi-
alists thought very highly of Spinoza’s monism, which could easily be
considered a forerunner of materialism (in deus sive natura, consider
“God” just an example of extravagant terminology, retain nature and
call it matter!). On the other hand, explicit idealism and especially
dualism are much less acceptable to dialectic materialism, because the
supposedly Christian notion of “immortal soul” may be hidden below.
Panpsychism is another modern form of monism in the sense of
Spinoza. Mind is simply the inside, and matter the outside of every-
thing. This doctrine has been more or less explicitly favored by so
different philosophers as Whitehead, Teilhard de Chardin, and also in
a sense by Russell as we have seen. It appears very plausible if regarded
from the point of view of human perception and thinking: mind “looks”
from the inside at the outside material world. It seems to run into dif-
ficulties at the atomic level: does every atom (in a stone, for instance)
also have an “inside” of a mental character? Leibniz (1646–1716) af-
firms just this and calls these “mental atoms” monads (cf. sec. 5.3).
A similar but more modern version of such a theory was held by the
early Russell and greatly elaborated by Whitehead. To be sure, it looks
rather extravagant, but is fascinating and may contain an important el-
ement of truth (we already know that Whitehead once said: “It is more
important that a proposition be interesting than that it be true”). The
possibly false theory of monads is indeed more interesting than the
certainly true relation “1+2=3” . . .
Panlogism (Greek: all is logic) is a label frequently put on Hegel
because his monumental three–volume work “The Science of Logic ”
contains his basic philosophy. He tried to derive everything by means
of his dialectic logic, which is an idealistic approach.
A quite different form of “panlogism” was proposed by the great
physicist John Archibald Wheeler (Misner et al. 1973, pp. 1211–1212).
The basic “building blocks” of the universe, even more fundamental
than elementary particles, are logical propositions p, q, r, . . . of the form
defined in sec. 2.1 (“Logic of propositions”)! This is the most radical
form of panlogism, much more extreme even than Hegel’s dialectic.
5.1. REALISM, IDEALISM AND DUALISM 205
Later on, Wheeler modified this view by retaining only the “truth
values” (1 or 0) of propositional logic (sec. 2.1). In informatics, each
alternative, 1 or 0, represents a bit (bi nary unit ). So the world consists
of “bits”, or according to Wheeler (1994, p. 296): It from Bit. Remark-
ably enough, Wheeler’s bits are much the same as Weizsäcker’s (1985,
p. 392) Ure (Uralternativen, basic alternatives). The same idea seems
to have been found independently. (I am sure that both Weizsäcker
and Wheeler are much too great personalities as to start a quarrel of
priority.)
Dialectic materialism. As we have already mentioned in sec. 2.5,
Marx, Engels and Lenin applied dialectics to nature itself, considered
to consist of matter only. The dialectic contradictions lie in nature
(matter) itself , they are in our thinking only because we think about
nature. Mind is an emergent property of matter, it does not exist
independently of matter. Engels’ well–known example of a dialectic
process in nature may appear somewhat farfetched: a wheat grain falls
to the ground and ceases to exist as such (negation ), giving rise to a new
plant which, on dying (negation of negation ) gives rise to the grain “at a
higher level”, i.e., greatly multiplied (cf. also New Testament, St. John
12:24).
Hegel’s dialectic idealism appears more refined and sophisticated,
but dialectic materialism (Engels’ dialectic of nature) may be more
concrete, “down to earth”. The difference may, at least partly, be ver-
bal rather than real: whether one calls the basic substance “spirit” or
“matter” has enormous emotional significance, but is of little actual rel-
evance if the logical structures are identical. For instance, Lenin was an
enthusiastic and highly intelligent reader of Hegel. Both Hegel and the
dialectic materialists claim the monist and “pantheist” thinker Spinoza
as their intellectual ancestor. In fact, monism (one basic substance)
seems to be what matters most.
True, Findley (1958, p. 58) speaks of “the Marxists, who try to op-
erate Hegelian machinery with quite alien and unsuitable fuel”, and the
above example of the grain of wheat seems to confirm this opinion. On
the other hand, Findley (1958) says also that Hegel “is more nearly a
dialectic materialist than most Hegelians have realized”. Again: “There
is, however, as much materialism in Hegel as in Marx, since matter is
for him certainly a stage in the ‘Idea’. (Just as there is certainly also a
strong strain of teleological idealism in the supposedly scientific mate-
rialism of Marx.)” (Findley 1958, p. 23). (The historic and economic
theories of Marxism are not a subject of the present book.)
Let us look more closely to the question whether dialectics is a
206 CHAPTER 5. PHILOSOPHY FOR SCIENTISTS
World 3
Most controversial is the “reality” of World 3. Let us think of
mathematics, considered as a prototype of this world already by Plato.
Do mathematicians “invent” or “discover” their theorems? Most
non–mathematicians would speak of “inventing”, most mathematicians
speak of discovery. Only a year ago (June 1993), there was a math-
ematical sensation: a proof of Fermat’s Last Theorem (xn + y n = z n
is unsolvable for integers x, y, z, n, except for n = 2, cf. sec. 2.3) was
finally found, after Pierre de Fermat had formulated this theorem in
1637 and claimed to have found a proof but not given it. Since then,
generations of mathematicians have tried in vain to find such a proof.
(I just — April 1994 — learned that even the recent “proof” contains
a serious flaw.) Thus, the general feeling of mathematicians is that the
theorem is “already out there”, just waiting for a proof (or disproof).
Mathematicians are constantly busy to prove the “existence” of a solu-
tion to a complicated equation. Thus, in some way, the mathematical
objects exist, just waiting for being discovered. Primarily, π is not “in
the sky”, but in the strange but obviously real mathematical universe
(Barrow 1992).
All great mathematicians have believed in the reality of this mathe-
matical universe, from Pythagoras to Penrose (1989). Bertrand Russell,
who in his philosophy stressed empiricism, was a Platonist in logic and
mathematics. Gödel emphasized the need of Platonism in logic and
philosophy.
5.2. THE THREE–WORLD MODEL 209
three worlds differ mainly in their ontological status; see also secs. 1.2
and 5.1.
At any rate, the Three–World terminology is very useful. As we
have already said above, materialists may consider World 2 a subset of
World 1; Spinozists may regard World 1, on the one hand, and World 2
+ World 3 together, on the other hand, as the two “attributes” of the
one divine substance; Eastern philosophers together with Schrödinger
(1958) who hold that there is only one universal mind of which we
all participate, will put World 2 ≡ World 3, Russell, in his “neutral
monism”, may put World 1 ≡ World 2, etc.
To the medieval philosophers, the world of ideas, our World 3, was
the mind of God ; this terminology is alluded to in the title of the book
(Davies 1992). Hegel calls World 3 the “objective spirit” (not a bad
terminology, after all).
Less bombastically, we may say that World 1 is governed by “fuzzy”
logic and “fuzzy” mathematics (sec. 2.4), whereas World 3 is the realm
of exact logic and mathematics. This point merits further elaboration,
which will be done in the next subsection.
R. Bošković (1711–1787) gave an ingenious interpretation of
World 3 as a mathematical space which is governed by potentiality in
the sense of Aristotle (sec. 5.4). World 3 is the seat of the human mind,
whereas matter and the human body belong to World 1 (World 2 may
perhaps be considered a subset of World 3). The interaction between
World 3 and World 1 explains why nature is governed by mathematical
laws. Not all mathematical structures are realized in nature (not ev-
erything that is potential , i.e., possible, becomes actual , i.e., real), but
theoretical physics can make free use of the inexhaustible treasure of
mathematics that is contained in World 3. Furthermore, if the human
mind is located in World 3, the direct access we have to mathematics
according to Penrose (see below) is explained.
Another application of these ideas is the interpretation of quantum
mechanics in terms of potentiality and actuality (sec. 3.5). An excel-
lent book about Bošković is Ivan Supek: “Rudjer Bošković ”, Croatian
Academy of Sciences and Fine Arts, Zagreb, 1989. More about Bošković
will be found in sec. 6.6.
5.2. THE THREE–WORLD MODEL 211
holds at least for continuous objects; one might argue that integers
occur in World 1 more directly: a basket containing five apples may be
considered an “exact” realization of the integer “5”; cf. sec. 4.4.)
Thus we have the following scheme of objects:
in World 3: exact,
in World 2: exact (at least in principle),
in World 1: fuzzy (at least in general).
This seems to be a clear indication that World 1 and World 2 are
essentially different.
Analogy with computers. Now it is well known that by computers
(which belong to World 1), also exact logical and mathematical opera-
tions can be performed. The underlying computer programs (software),
however, belong to World 3. At the danger of overstretching the anal-
ogy, we may say that the program adapted so as to be able to serve as
input for a particular computer, “belongs to World 2 of that computer”.
Below we shall see that the computer analogy cannot be perfect.
Still we may say that World 1 and World 2 are essentially different, at
least as much as hardware and software in machine computation.
Mind vs. computer . It is claimed, however, by the most eminent
mathematicians that, in a way, they have direct access to World 3
which goes beyond formal (algorithmic) computation or logical deduc-
tion. This is most strongly and convincingly argued in (Penrose 1989,
pp. 416–423). From p. 418 we quote
Mathematical truth is not something that we ascertain merely by use of
an algorithm. I believe, also, that our consciousness is a crucial ingredient
in our comprehension of mathematical truth. We must “see” the truth of
a mathematical argument to be convinced of its validity. This “seeing” is
the very essence of consciousness. It must be present whenever we directly
perceive mathematical truth. When we convince ourselves of the validity of
Gödel’s theorem we not only “see” it, but by so doing we reveal the very
non–algorithmic nature of the “seeing” process itself.
What Penrose alludes to here, is the fact that in Gödel’s proof
(sec. 2.3) we can “see” by informal reasoning that Gödel’s basic propo-
sition G, though algorithmically unprovable, is nevertheless true. Cf.
the quotation of Findley at the beginning of sec. 2.5.
Lucas’ proof of free will . Lucas (1970, §25) has used Gödel’s theorem
in a similar way. The essentially non–algorithmic character of human
thought as exhibited by Gödel’s theorem shows that our thinking can-
not be the activity of a deterministic “thinking machine” because such
5.2. THE THREE–WORLD MODEL 213
a machine can only work algorithmically. More about this will be found
in sec. 6.4.
Thinking vs. acting. For those who believe that thinking is simply
an activity of the brain or the nervous system, essentially similar to its
activity in directing our bodily movements, we have a surprise in stock.
In its thinking function, our brain works (or is able to work) exactly,
as we have seen. Thus, by reasoning, we can establish rigorously valid
theorems about circles.
Now let us use the brain as a “steering unit” for a bodily movement,
e.g., the actual drawing (on paper or on the blackboard) of a circle. If
we hope to be able to draw a circle with equal exactness, we shall be
badly disappointed as we have seen above. Nobody is able to trace an
exact circle, cf. Fig. 6.2 on p. 255.
The reason is, of course, the inherent “fuzziness”, the inevitable
random background in World 1, cf. secs. 4.4 and 6.5.
Now comes the essential question. If this random background affects
our brain and nervous systems with respect to bodily movements, why
does it not affect our brain equally in its thinking activity? Comparing
the brain to a parallel–processing digital computer helps, but is only a
more or less perfect analogy because of the non–algorithmic character
of human thinking and because of the fact that the digital firing of
the neurons does not yet imply that the brain is a digital computer
(sec. 1.1).
As we have remarked at the beginning of sec. 4.4, only integers can
be determined exactly, whereas continuous quantities can only be de-
termined approximately. In computer language, this is the distinction
between digital and analog computers. It is thus tempting to relate
exact thinking to some digital operation of the brain; some analog
(continuous) operations of the brain would then be responsible for the
inexact bodily movements. There may be an element of truth in this
comparison, but almost certainly matters are not that simple. At any
rate, the “exactly thinking I” may be considered as a “center of com-
mand” for logical operations, which has no direct analogue in World 1
which is always “fuzzy”.
To me, this is another strong argument for the independent char-
acter of World 2: this “center of command” seems to be identical to
Penrose’s “consciousness” (1989, pp. 409–413) and to the “self” of Pop-
per and Eccles (1977) and to be essentially non–material in order not
to be subject to random fluctuations.
214 CHAPTER 5. PHILOSOPHY FOR SCIENTISTS
Sun may have exploded or the Earth destroyed by a giant comet. Not
even a falsification necessarily proves a law wrong: the experimental
setup may have been inappropriate or measuring errors may have pro-
vided the false impression that the theory was incorrect. This is the
problem of induction already discussed at some length in sec. 3.9.
All this is relatively trivial. A deeper question is whether reality
really “is” as it looks. For instance, is space “really” three–dimensional?
It may be five–dimensional or ten–dimensional, only we may not be
capable of perceiving more than three dimensions. Are there aspects of
reality which are inaccessible to science? Are there Kantian “things–
in–themselves” which are unknown as a matter of principle (p. 230)?
As a partial answer, let us consider the spectrum of electromagnetic
waves, with wave lengths from several kilometers down to 10−13 m. The
visible spectrum, from red down to violet, is on the order of 5 × 10−7 m,
above there are infrared and longer electromagnetic waves; below there
are ultraviolet waves, X–rays and gamma–rays. We know very well and
use also electromagnetic waves above and below the visible part of the
spectrum: they are accessible to indirect physical observation.
So it seems that we know practically everything which in principle
can be observed by present–day physical, chemical, etc. experiments.
Other things, e.g., some elementary particles in physics, have been pre-
dicted by theory but not yet been discovered (this is a nice illustration
of “knowledge of non–knowledge” mentioned above). Also higher di-
mensions, which are inaccessible to direct observation, are mathemati-
cally fully understood (sec. 2.6) and are under consideration as candi-
dates for certain “unified theories”. Infinite–dimensional spaces are, of
course, standard tools in quantum mechanics (without implying that
“ordinary” space is more than three–dimensional). (See, however, also
Bohm’s (1980) “enfolded reality” and “implicate order”, cf. sec. 3.5.)
It would be very unwise indeed to exclude that there are physical or
other phenomena which are completely different from all that we know
today and which cannot be observed by contemporary science, and for
which we do not even know where to look . As an example, think of
quantum phenomena before 1900.
What we can say is that our knowledge of the physical world is
essentially correct in corresponding to some external reality; this does
not exclude the possibility or even probability that it is incomplete.
Think of trying to catch a fly. Why is it so difficult? The fly’s
knowledge of the external world is certainly extremely rudimentary: it
probably comprises mainly the (instinctive) knowledge of how to get
food, how to escape enemies, etc. This limited knowledge, however, is
5.3. SUBJECT AND OBJECT 217
Monadic structures
Fig. 5.3 schematically illustrates a “monadic structure”, consisting
of a subject A and an object B. It is probably the simplest and most
fundamental structure in epistemology, relating an object, considered
to be at a location B, to a subject A (represented by the eye of the
observer). In order to avoid Whitehead’s “fallacy of simple location”
(sec. 5.1), we consider the object with respect to a certain location B
without excluding the possibility that the object is implicitly present
also at other locations, cf. Fig. 5.1 on p. 200.
l = eT L e (5.2)
where, in the observed average value l, the effect of the object (physi-
cal quantity) is the matrix L, and the effect of the observer is the unit
vector e (“state vector”); eT denotes the transpose of e. (For an even
more mathematical reader we shall not conceal the fact that the “ob-
servable” L represents an infinite matrix in the Heisenberg sense or,
equivalently, a linear operator in the Schrödinger sense; the vector e is
an infinite “state vector” or a “state function”.) The measured value l
thus is the projection of the “physical quantity” L onto the reference
system e of the observer. For more details see sec. 3.5.
Philosophy quite naturally starts with the subject, that is, with
idealism. Equally naturally, science starts with the object, that is with
realism or materialism.
As we have already remarked in sec. 5.1, realism is not absolutely
identical with materialism. Realism only emphasizes the priority and
existence of the external world, but does not deny that other persons
may have minds, as materialism does (as a matter of fact, materialism
also denies the reality of mind in the subject).
In sec. 5.1 we have seen that scientific realism, in the form of
the mathematical laws of physics, contains “non–material” objects of
World 3. This synthesis of “realism” and “idealism” was seen to make
“matter less material and mind less mental” (Russell).
In fact, like in the hen–egg problem, subject and object are insep-
arably connected in human knowledge, as already Plato, especially in
his dialogue Parmenides, has emphasized (e.g. Speiser 1952, p. 13): no
subject without object, no object without subject.
In science, this connection is particularly strong in quantum theory,
and very weak in sciences like paleontology. There the scientific subject
may come millions of years later: nobody will question that dinosaurs
have existed even without human observers watching them with awe
and fear, but nobody will question either that human observers are
necessary for the science of paleontology. Nature may very well exist
without human observers, but philosophy certainly not, and not even
science.
Fichte’s iteration
Matters become difficult in introspection: the mind is subject and
object at the same time: the thinking thinks the thinking. This has
already been mentioned in sec. 2.5; see the quotation of P.M. Møller
which was so dear to Niels Bohr, and the subsequent closely similar
quotation from Fichte. In fact, this does lead to an infinite regress
of “I’s” but that may not really matter: there are also many kinds
of mathematical infinities, so that World 3 (sec. 5.2) is certainly not
subject to overpopulation as our small Earth is.
Still, there are the well–known paradoxes of self–reference which we
have already discussed in secs. 2.3 and 2.5. As Sir John Kendrew put
it (Duncan and Weston–Smith 1977, p. 207):
Perhaps the most fundamental of all the difficulties encountered in biological
research is that the investigator cannot detach himself from the system under
study because he himself forms part of that system . . . in psychology the
investigator is himself one of the subjects under study.
5.3. SUBJECT AND OBJECT 221
Logical singularities
Fichte’s “Absolute” is clearly a logical singularity, as Fig. 5.5 shows.
This is a general situation, which is very well described by the quotation
from John Kendrew given above.
Let us recall some examples of such a subject–object interference:
part of the universe. The way in which Fichte and Hegel approach the
problem may, as a variation of Fig. 5.5, also be described as a spiral
(Fig. 5.7), corresponding to looking at the singularity S in Fig. 5.6 from
above. Fig. 5.7 clearly corresponds to the dialectic spiral of Figures 2.10
or better 2.12 on p. 53, and hence to Weizsäcker’s “Kreisgang”.
Aristotle
Aristotle (384–322 B.C.) was a student of Plato. He is the creator
of logic and systematic metaphysics. Substance consists of matter and
form. There are 4 causes:
– causa efficiens: causality in the modern sense (determinism),
physics;
– causa finalis: final cause, “downward causation”, biology (cf.
sec. 4.1);
– causa formalisis (the plan of a house); and
– causa materialis (matter, e.g. bricks of which the house is built).
The reader may wonder about these Latin names: they come from
medieval philosophy, especially St. Thomas Aquinas (see below). Of
course, one may also speak of “efficient cause”, “final cause”, “formal
cause”, and “material cause”.
The last two “causes” correspond to matter and form mentioned
above. In contrast to Plato, his emphasis is on classification and de-
scription rather than on mathematics. He was much more interested
in concrete science (physics, biology) and much more systematic than
Plato who was primarily a mathematical thinker. God, to him, is the
“prime mover”, the first cause. Aristotle was very influential in the
Middle Ages, where he was regarded as an authority in science as well
as in philosophy. The very apparent completeness of his scientific work
made it obsolete as soon as new discoveries were made which did not fit
into his system. His main importance in science today is his creation of
systematic logic, which was transcended only by symbolic logic (Peano,
Frege, Russell, Gödel etc.). Through the mathematization of modern
science, Plato won over Aristotle. With some pointed exaggeration we
may say that Plato is the philosopher of mathematical physics, whereas
Aristotle is the philosopher of biology.
Nevertheless, his distinction between potentiality and actuality is
important for the difficult problem of the relation between mathemat-
ics and the physical world (sec. 5.2, p. 210), for the interpretation of
quantum mechanics (sec. 3.5, p. 108), etc.
Neoplatonists
Plotinus (204–270 A.D.) took up and developed further the mystical
and mathematical aspects of Plato’s thoughts. His works are literary
228 CHAPTER 5. PHILOSOPHY FOR SCIENTISTS
Medieval philosophy
The main miracle of Western philosophy in the dark time between
500 and 800 is its sheer survival: all the principal works of classical
antiquity have been preserved. With Johannes Scotus Erigena (around
810–877) original work in the Platonist tradition started again, and
with Anselm of Canterbury (1033–1109) it reached a first culmination.
With Anselm, the school of medieval scholastic started. It flour-
ished in Paris and at other universities, and it is a school in which
theological and philosophical questions were discussed in a professional
way, based on more or less generally accepted premises; sometimes also
these premises were subjected to discussion. This kind of school philos-
ophy has contemporary analogues in the schools of Marxist philosophy
in the communist countries, and of logical positivism in the West.
The most eminent scholastic philosopher was St. Thomas Aquinas
(1225–1274). Whereas his predecessors were Platonists, Thomas used
the work of Aristotle as the basis of his philosophy. This was his
strength and his weakness. He gained a rigorous and consistent sys-
tem, but he partly lost Plato’s inspiration. Thomas’ merit was also a
5.4. HISTORICAL LANDMARKS 229
(Stegmüller 1969).
Fichte
Johann Gottlieb Fichte (1762–1814) started out as a student of
Kant, transforming his transcendental idealism and enormously devel-
oping his dialectic approach. Originally, in his Wissenschaftslehre of
1794, he emphasized the (transcendental) I, posing the dialectic triad
I (subject, thesis), non–I (object, antithesis) together with their syn-
thesis. This obviously led Russell (1945, p. 718) to write that Fichte
“carried subjectivism to a point which seems almost to involve a kind
of insanity”. This is a classical example of a gross misjudgment of
a philosopher by a colleague; such misjudgments unfortunately occur
rather frequently in the history of philosophy (and it would be pre-
sumptuous to claim that the present book is free of them).
Later, in the Wissenschaftslehre of 1804, Fichte started with the
object (Fichte’s iteration, sec. 5.3). So, if Fichte is called an idealist, by
the same token he could be called a realist. This is another example of a
profound truth in the sense of Niels Bohr (sec. 2.5). The mathematician
Speiser (1959) relates Parmenides, Plato, and Fichte in a remarkable
“synthesis”.
Hegel
Georg Wilhelm Friedrich Hegel (1770–1831) is considered the great-
est dialectic philosopher. We have already discussed his dialectics to
considerable length in sec. 2.5. Although he is usually regarded as an
idealist philosopher, his thinking has become basic also for dialectic ma-
terialism: Karl Marx (1818– 1883) and Friedrich Engels (1820–1895).
Vladimir Ilich Lenin (1870–1924), the Russian revolutionary, studied
Hegel in great detail.
Needless to say, philosophers like Kant and especially Fichte and
Hegel are very difficult to read. Nevertheless, Fichte’s Wissenschafts-
lehre of 1804 and Hegel’s logic, which provide two alternative ways of
“ascent to the Absolute”, are extremely rewarding and well worth the
reading effort, just as the Rabbi’s third speech mentioned in the Preface.
(It is preferable, however, to start with the secondary literature, e.g.
(Findley 1958).)
Logical positivism and analytical philosophy
Logical positivism, also called neopositivism, started at Vienna, in-
spired by the physicist Ernst Mach (1836–1916), with Ludwig Wittgen-
stein (1889–1951), Rudolf Carnap (1891–1970), and others, and by
5.4. HISTORICAL LANDMARKS 233
which have retained their validity even today. Some problems formu-
lated by Parmenides, Plato, Aristotle, Descartes, or Leibniz have not
lost their actuality to the present day. The whole intellectual personal-
ity of a philosopher, in a sense, is much more important than the per-
sonality of a scientist. Newton’s or Einstein’s discoveries have merged
into the general stream of science. Plato’s views are discussed as Plato’s
views, Hegel’s views are commented or condemned as Hegel’s views. In
his charming little book, Jaspers (1953) says that philosophy is not a
science with more or less well established and recognized results, but an
ongoing process of thinking: “Philosophy means to be on the road. Its
questions are more essential than its answers, and each answer becomes
a new question.” Sometimes, walking on the road of philosophy, one
sees wonderful new vistas, but the true philosopher is not satisfied: he
moves on.
Thus all important philosophers from Plato to Wittgenstein are for
us, so to speak, contemporary philosophers. This may be compared to
music: judging from concert programs, available CD–recordings etc.,
our real contemporary composers are Bach, Mozart, Beethoven, and
Brahms! Also the best book on Beethoven is no substitute for listening
to his symphonies, and the best book on Plato’s philosophy cannot
dispense us from reading his dialogues. On the other hand, although
Einstein wrote fine books on relativity theory, there are contemporary
works such as (Misner et al. 1973) which are more modern and more
comprehensive and thus, in a certain sense, do supersede Einstein’s
books.
Let me conclude with a quotation from C.F. von Weizsäcker (1970,
p. 202), which I owe to Viktor Gutmann (Vienna):
[The physicist does not notice] . . . that by rejecting professional philosophy
he did not free himself from philosophy but became himself a dilettante
philosopher. Unconscious philosophy, however, is in general worse than a
conscious one, and thus just the most profound thinkers of modern physics
invariably return to philosophy.
236 CHAPTER 5. PHILOSOPHY FOR SCIENTISTS
Chapter 6
Philosophical implications of
science
237
238 CHAPTER 6. IMPLICATIONS OF SCIENCE
the passage of is real; at present, the future is not yet determined and
can still be influenced; thus there is room for creativity. The second
doctrine says that the block universe is essentially finished; everything
has been determined beforehand, and there is no place for freedom and
creativity; the passage of is only an illusion. According to Weyl (1949,
p. 116):
The objective world simply is, it does not happen. Only to the gaze of my
consciousness, crawling upward along the life line of my body, does a section
of this world come to life as a fleeting image in space which continuously
changes in time.
The problem of free will has a close relation to the mind–body prob-
lem (“downward causation”, cf. sec. 4.1; action of World 2 on World 1,
cf. sec. 5.2) and to the concept of an open, creative universe (cf. sec. 6.3).
Quantum theory
There is no doubt that quantum theory governs neural activity
rather than classical deterministic physics (Stapp 1993). Nonetheless
is it controversial whether the indeterminism of quantum jumps is di-
rectly responsible for free will, as some physicists think; cf. (Jordan
248 CHAPTER 6. IMPLICATIONS OF SCIENCE
1968, p. 338). The reason is that between quantum jumps, the quan-
tum state function evolves deterministically (according to Schrödinger’s
equation), and the quantum jumps themselves are governed by statis-
tical laws which seem to be no less rigid than deterministic laws, so
as to leave little room for freedom. At any rate, quantum fluctuations
provide an omnipresent random background which is important as we
have seen above.
Also, quantum theory mitigates the dualism between mind and
matter showing that the simplistic materialism of some modern “neu-
rophilosophers” and “neuroscientists” such as Churchland (1988), Den-
nett, and Edelman (1989) is inadequate: the adequate quantum theory,
in some way, always seems to involve mind; see sec. 3.5, (Lockwood
1989), (Margenau 1984), (Squires 1990), and (Stapp 1993).
Going an essential step further, the physicist F. Beck and J.C. Eccles
(Eccles 1994, sec. 9) have elaborated a detailed quantum–mechanical
model for mind–brain interaction. Such a model may not describe the
real situation, but it shows that such an interaction is possible.
This model is based on the plausible assumption that mind may
change the distribution of quantum probabilities of micro–events in the
brain, whose combined action produces the desired macro–event. This
is another example of amplification (sec. 4.1).
Each week, my mail contains a number of theories that people have sent me.
They are all different, and most are mutually inconsistent. Yet presumably
the grand unified theory has determined that the authors think they were
correct. So why should anything I say have any greater validity? Aren’t I
equally determined by the grand unified theory?
What does Epicurus’ argument really show? It does not directly
refute determinism. It shows, however, that universal determinism
(“everything is predetermined by a grand unified theory” (sec. 6.6)) is
non–arguable. If someone claims that he has found a proof of universal
determinism, we can immediately reply that his reasoning is inevitable
if universal determinism is true, regardless of the argument being log-
ically valid or not. So his argument is simply worthless, as the above
quotation from Hawking shows very clearly.
The argument of Epicurus is particularly malicious and insidious. It
turns determinism against itself: it beats determinism at its own game.
By the way, Gödel’s proof uses self–reference (reflexive logic, see
end of sec. 2.5), and also the argument of Epicurus is self–referential.
Otherwise, both arguments have little in common, but we see again the
importance of self–reference in philosophic reasoning.
Take the example of Kepler’s three laws for the motion of planets
around the Sun. The first and most important law is that planets
252 CHAPTER 6. IMPLICATIONS OF SCIENCE
move along ellipses whose focus is the center of the Sun. It is easy (I
do it regularly in my introductory course on mechanics for students in
the second semester) to derive, from Kepler’s laws, Newton’s laws of
gravitation.
Inversely, Newton’s law of gravitation permits the derivation of Ke-
pler’s laws, with two important qualifications.
(1) Possible orbits are not only ellipses, but also other conic sections:
parabolas and hyperbolas. This is relevant for comets.
(2) If, in addition to the Sun and the Earth (say), the attraction of
other planets is taken into account, then the orbits are slightly
perturbed: they no longer are exact ellipses.
The second fact is particularly important: Kepler’s laws do not hold
exactly, due to the perturbation of other planets! If the measurements
of Tyho de Brahe, from which Kepler derived his laws, had been more
accurate, then Kepler would never have arrived at his simple laws, and
Newton’s laws would not have followed so readily. In this case, as Alfred
North Whitehead has pointed out, too accurate measurements might
have hindered the progress of science!
This is another nice instance of a Hegelian triad (sec. 2.5). Thesis:
Kepler’s laws; antithesis: Newton’s laws; synthesis = thesis on a higher
level: Kepler’s laws with corrections for the effect of other planets.
Even Newton’s law of gravitation, however precise (to 10−7 ), is
not absolutely exact: Einstein’s theory of general relativity is better
(sec. 3.4). Einstein’s theory comprises Newtonian mechanics as a spe-
cial case for “small” velocities and “weak” gravitational fields, which is
sufficient for most applications.
Einstein’s general theory of relativity is extremely elegant, general,
and beautiful. So far, it has never been refuted by experiment. Is it
absolutely true?
Quantum mechanics (sec. 3.5) is equally elegant and general; it has
been confirmed by experiments better than any other theory. Is it
absolutely true?
Quantum theory holds for very small distances: between nucleus
and electrons in an atom, and between atoms in a molecule. At these
small distances, quantum effects must occur: all physicists agree on
that.
Now, no quantum effects can be derived from general relativity.
Therefore, general relativity cannot hold for very small distances. Thus
it cannot be absolutely true.
6.5. LAWS OF NATURE 253
exactly on the curve (Fig. 6.2). No matter how firm your hand is, it
will inevitably perform (very small) shaky movements. There seems to
be a randomizer in our nervous system which, however, works on the
physical level only and does not affect the mental level (sec. 5.2).
Conventionalism
Well, if the relativistic concepts differ so little from the correspond-
ing Euclidean concepts, why not assume Euclidean geometry to be
exactly valid, and treat the very small deviations (of order 10−8 ) as
“relativistic corrections”? In this spirit, why should we not say that
light rays are not Euclidean straight lines but are slightly curved, so
that α + β + γ = 180◦ + , where is a small quantity which is not zero
because of curvature effects?
6.5. LAWS OF NATURE 257
Positivism
Hipparchus (around 146–127 B.C.) and Ptolemy (2nd century A.D.)
explained the motion of the planets as seen from the Earth, by a su-
perposition of epicycles. Mathematically this is a series of cosines of
form
the planets are directly observable entities, but these arguments need
not be accepted by a stubborn opponent. Thus I should say that those
who regard the planets as metaphysical entities are still less in number
than those who regard in this way atoms, electrons or quarks (have you
ever seen a quark?), but the principle persists (have you ever been on
Mercury?).
In reality, of course, Kepler’s laws of planetary motion, and New-
ton’s and Einstein’s (sec. 3.4) theories of gravitation are universally
preferred by physicists. This shows that even physics contains unob-
servable quantities so that the positivist point of view is not carried
through completely in scientific practice.
Whitehead (1933, Chapter VIII) has given a beautiful example.
Earlier in this century it was remarked that a series (6.1) as known
before had to be supplemented by extremely small additional terms in
order to achieve best agreement with very precise contemporary ob-
servation. “Every Positivist must have been completely satisfied. A
simple description had been evolved which fitted the observed facts”
(ibid., p. 127). But, in fact, the American astronomer Percy Lowell
was not satisfied. He found out by computation that the new anoma-
lous terms corresponded to the attraction of an imaginary small planet,
an astronomer at the Lowell Oberservatory photographed in 1930 the
sky in this direction and found . . . the new planet Pluto! Whitehead
wrote
The civilized world has been interested at the thought of the newly discov-
ered planet, solitary and remote, for endless ages circling the sun and adding
its faint influence to the tide of affairs . . . The speculative extensions of laws,
baseless in the Positivist theory, are the obvious issue of speculative meta-
physical trust in the material permanences, such as telescopes, observatories,
mountains, planets, . . .
Thus positivism, with its justified emphasis on observability and
mathematical description, is important but should not be interpreted
too narrowly.
Here positivism, similar as conventionalism before, represents a cer-
tain general direction of thinking rather than a definite school of philos-
ophy such as logical positivism (sec. 5.4). Whitehead’s pointed remarks
should be understood in this sense.
SUPERB are, for instance, classical mechanics (within the usual limits
of applicability defined by the theories mentioned next), special and
general relativity, and quantum mechanics. USEFUL are, e.g., the
gauge theories in elementary–particle physics, in particular the “Stan-
dard Model” (sec. 3.6). TENTATIVE is, e.g., superstring theory as a
possible candidate for a “Theory of Everything” (sec. 6.6).
gravitation,
electromagnetism,
weak force,
strong force.
Thus three of the four forces have been unified. The great excep-
tion is gravitation, which somehow refuses being included in an ordi-
nary gauge theory. An extension of the latter, supersymmetry, may
provide a possibility to include gravitation as well. A final refinement
(and complication) is achieved by regarding elementary particles not
as pointlike but as an extremely small (∼ 10−13 cm) curve, or loop, a
“string”.
These strings can be combined with supersymmetry to give super-
string theory. This theory (or its variants) is so complicated that the
mathematical consequences have not yet been elaborated. Thus it has
not yet been possible to really understand whether and how well this
theory applies to the real world. With superstring theory we are now
(1994) so far as with Heisenberg’s unified theory in the sixties.
Excellent Rabbi type 1 references on these fascinating topics are
(Barrow 1991), (Davies 1984), (Davies and Brown 1988), (Hawking
1988, 1993), and (Kaku 1994). We again refer also to sec. 3.6.
Rudjer Bošković (1711–1787). “One of the most remarkable and
neglected figures in the history of modern European science was Roger
Boscovich” (Barrow 1991, p. 17). A Jesuit from Dubrovnik, he worked
in philosophy, mathematics, physics, geodesy, and similar sciences
which at that time were not yet so separated and specialized as to-
day. Recognized by Niels Bohr and Werner Heisenberg, he nevertheless
is little known outside Croatia. He proposed a grand unified force law,
generalizing Newton’s law of gravitation to include all other physical
forces. He believed in atoms and elementary particles and tried to find
a single law encompassing gravitation as well as the forces that would
hold his hypothetical elementary particles together. “His continuous
force law was the first scientific Theory of Everything” (Barrow 1991,
p. 18).
Bošković is important not so much for his particular contributions to
science, but he was a visionary whose ideas influenced great physicists
such as Faraday, Maxwell, Boltzmann, and Lord Kelvin. Some of his
philosophical ideas are relevant even today, cf. sec. 4.4 (p. 186) and
sec. 5.2 (p. 210).
Philosophical problems. A future TOE (Theory Of Everything )
poses very difficult and sometimes contradictory problems which are,
however, profound and fascinating.
(1) In the ideal case, it would describe and predict all physical
processes, however complex and “chaotic”: laws together with
boundary conditions.
266 CHAPTER 6. IMPLICATIONS OF SCIENCE
(2) If chemistry and biology are, in fact, nothing but theories of par-
ticularly complex physical systems, a TOE must also fully explain
chemistry and biology, including “self–organization”.
(3) It must incorporate the uncertainty principles of Gödel and
Heisenberg.
(4) If mental brain activities were, in reality, only physico–chemical
processes, then TOE would also describe human thinking; the
“freedom of the will” would then be nothing but an illusion.
(5) The TOE, predicting everything, must also derive itself .
It is clear that all this appears quite improbable. We would then
have Laplace’s demon (sec. 3.1) again, in contradiction to chaos theory
and Heisenberg’s uncertainty relation.
As we have seen in sec. 4.5, a reduction of biology to physics may
work in this “analytic” direction. In the opposite (“synthetic”) direc-
tion it probably will not be possible because information is needed, in
agreement with the “equation”
life = matter + information .
As a TOE, at least as it appears now, does not contain this “vital”
information, “synthesis” and prediction appear more or less impossible.
So the freedom of the will does not seem to be in danger, not even
from a TOE. Even if theoretically, provided we knew the required ini-
tial conditions fully and with absolute precision, we could calculate
human thinking, this cannot be done practically because the amount
of calculation exceeds all that is humanly possible. Furthermore, if we
could exactly predict our future decisions, they would be known, and
this knowledge might change our future thinking and acting! This in-
genious argument was given by Hawking (1993, p. 135). In fact, this
“pre–cognition” would constitute an inadmissible time loop (sec. 3.7,
Fig. 3.20 on p. 129).
Finally, even if a TOE were able to derive itself, how would we know
that it is true? Of course, in this case, the TOE would be internally
consistent, but how do we know that it corresponds to external reality?
Hawking’s answer is Darwin’s theory of evolution: animals behaving
not in agreement with reality, and early man having wrong “theories”
about reality are unfit to survive and will be eliminated by evolution (cf.
sec. 1.4). Interesting as this argument is (we have used it frequently),
I should not consider it a philosophically fully satisfactory answer.
6.6. THEORIES OF EVERYTHING 267
This is related to the argument of Epicurus (sec. 6.4); and also the
argument of Lucas (ibid. ) is of direct relevance here.
Let us therefore ask again: “Is everything predetermined?” Assum-
ing that a TOE exists, Hawking (1993, p. 139) finally says: “The answer
is yes, it is. But it might as well not be, because we can never know
what is determined.” Isn’t this a beautiful example of a profound truth
(of which the logical contrary is also a profound truth) in the sense of
Bohr (sec. 2.5)?
A final, and presumably conclusive, argument against complete pre-
determination comes from the Heisenberg uncertainty relation. Since a
TOE must comprise quantum theory, it must also contain this uncer-
tainty relation which is an ineluctable consequence of quantum theory
(sec. 3.5). Cf. also (Gell–Mann 1994) and (Weinberg 1993).
Relation to dialectics. As we have seen, a true TOE must derive it-
self . This is another instance of the important concept of self–reference:
As we have seen at the end of sec. 2.5, the dialectic system of Hegel’s
logic provides such a self–reference or self–derivation; cf. also (Speiser
1952, p. 110). This is possible in Hegel’s informal thinking, but not so
readily in the formal structures of contemporary logic and mathemat-
ics. It would have to be a self–referential deductive system (Wheeler
1994, p. 309).
In the light of these remarks it seems unlikely that we shall ever
have a true TOE in the strict sense of the contradictory requirements
(1) to (5) given above. What can happen if no TOE exists at all?
Hawking (1988, p. 166) gives two alternatives.
(1) There is no final theory of the universe, only an infinite sequence
of theories which better and better describe the universe. This is the
present view of those physicists who do not believe in a universal TOE.
(2) There is no theory of the universe. The physical laws are valid
only to a certain amount of accuracy; below this accuracy, events occur
in a spontaneous and random fashion. This would be a general instance
of the dialectics of freedom and necessity (sec. 2.5); this possibility has
also been envisaged by Whitehead (1933, Chapter VII). We are also
reminded of Wheeler’s “no laws” (sec. 6.5).
Bootstrapping. Remarkably enough, superstring theory seems in
fact to be able to “derive itself” in the form of “bootstrapping”, cf. end
of sec. 2.5: provided the system of elementary particles (as predicted
by superstring theory) exists, it produces itself. This is the bootstrap
principle of self–consistency (Gell–Mann 1994, Chapters 10 and 14); cf.
also (Capra 1976, Chapter 18).
268 CHAPTER 6. IMPLICATIONS OF SCIENCE
Logical singularities
Self–reference and self–derivation may lead to logical singularities.
Examples have been given at the end of sec. 5.3 (especially Examples
1 to 3).
Logical singularities occur if the observer (or theoretician) A
(α) coincides with the object B or
(β) is part of the object B.
Case (α) has been seen to apply Plotin’s definition of human think-
ing: The thinking thinks the thinking (sec. 2.5). Case (β) corresponds
to B being the whole universe, of which A necessarily is a part (from
Kant’s antinomies to TOE). It is illustrated by Fig. 6.4.
Figure 6.4: Subject A lies (a) outside object B, and (b) inside uni-
verse B
(For specialists only.)A simple model for such a logical singularity is
provided by the mathematical singularity of the gravitational potential
of an extended body B, well known to physicists and geodesists. The
Newtonian potential V at point A is
ZZ Z
ρ
V (A) = G dv , (6.2)
l
B
Figure 6.5: The Newtonian potential (a) for an external point A, and
(b) for an internal point A.
also certain that God exists. But lest believers rejoice too early, the
argument can be inverted: if God’s nonexistence is in any way possible,
it is certain that God does not exist!
This is indeed a logically fascinating paradox of the Infinite, and
the little book (Plantinga 1965) is as intellectually rewarding as (Nagel
and Newman 1958).
So far we have considered the concept, or concepts, of God mainly
from a philosophical and logical point of view. So far it has been
independent of revealed religion. This distinction between philosophy
and theology (religion) was made already by St. Thomas Aquinas. An
excellent, comprehensive and objective treatment of both points of view
(philosophy and religion) was given by Küng (1978); an outstanding
concise, clear and understandable summary is found in the last chapter
of (Bocheński 1959).
Science and religion. It is a commonplace that science and religion
do not contradict each other because they are dealing with different
subject matters. Nevertheless, they have a common boundary, and
wars are frequently waged about boundaries.
What should be rejected is the doctrine of two separate truths, one
for science and one for religion. In fact, the synthesis between science
and religion is not a cheap compromise, but requires considerations on
a rather high level.
Such high–level efforts can be found, for example, in the writings
of Whitehead and of Weizsäcker, and on the Catholic side of Teilhard
de Chardin (1955), and in particular of Karl Rahner. Cheap popular-
izations frequently do more harm than good. Tipler’s (1994) ingenious
and ingenuous attempt to develop a theology on the basis of physics
is probably a singularity, although an interesting one. (Davies 1983)
and (Ferguson 1994) are more realistic. (Kolakowski 1982) is a fine
counterpoint to (Küng 1978) by a former Marxist.
A chief stumbling–block has been the theory of biological evolution.
The theory that man has evolved from animal predecessors was consid-
ered incompatible with man’s immortal soul and hence with Christian-
ity. This apparent contradiction was used by both sides to fight the
other, with most unfortunate results.
In fact, even if man has evolved in this way (which no serious sci-
entist would doubt nowadays), man’s intellect is such that he is clearly
separated from the other animals. Thus, he seems to be the only ani-
mal capable of thinking about himself. This “thinking about thinking”
is self–consciousness, cf. (Eccles 1989).
Another really stupid stumbling–block has been the apparent con-
272 CHAPTER 6. IMPLICATIONS OF SCIENCE
BIBLE SCIENCE
creation “big bang”
light (1st day) energy
Heaven and Earth (2nd day) stars, planets, Earth
oceans, oceans,
vegetation (3rd day) vegetation
Sun and Moon, day Earth rotation,
and night (4th day) formation of the Moon
fish and birds (5th day) animals
mammals, man (6th day) mammals, man
With the possible exception of the 4th day, the order is almost exact;
cf. (Weizsäcker 1973).
As we believe today, evolution is not restricted to biology, but
started right after the “big bang”. First hydrogen was formed, then
helium was synthesized, and heavier elements followed. Galaxies and
stars were formed. Stellar evolution is now standard in astronomy. The
evolution of the planetary system was discussed already by Laplace and
Kant.
The history of mankind and the story of the bible fit beautifully into
this picture of a thoroughly evolutionary universe. Christianity was
always proud of its historical evolution, from the first man (“Adam”
simply means “man”!) to Moses, King David, and Jesus. Not to see
that this historical development fits perfectly into general evolution
seems, in hindsight, an almost unbelievable shortsightedness. Teilhard
de Chardin (1955) was one of the first Catholics to work this into a
consistent picture, and he had to pay for it. Nowadays, of course, this
is no problem any more.
Can science help prove the existence of God ? When Napoleon asked
Laplace why he did not mention God in his famous book on celestial
mechanics, Laplace answered: “Sire, I do not need this hypothesis”. Of
course, he was right from the point of view of classical mechanics: God
6.7. THE ABSOLUTE 273
may have installed the giant mechanical clockwork of the universe, but
since then it continued to run automatically according to the rigorous
laws of mechanics. Divine interference was no longer needed in classical
mechanics.
Other scientists and philosophers thought differently: Kepler ad-
mired the divine harmony of the universe, and so did Newton. Leibniz
tried hard to reconcile science and religion. Kant deduced the existence
of God and the immortality of the soul from the moral law to which all
men are subject. Goethe, following Spinoza, sought God in nature and
art, Einstein in the wonderful mathematical laws of nature (“Science
without religion is lame, religion without science is blind” (Einstein
1954, p. 46). Heisenberg and Weizsäcker are Evangelic Christians.
Also to me it is impossible to believe that a beautiful mountain
flower or Riemann’s formula for the distribution of prime numbers are
pure products of “chance” or “natural selection”.
Biologists, discussing evolution, frequently speak of the “creativ-
ity” or the “inventiveness” of evolution. The physicist Freeman Dyson
(quoted from (Kaku 1994, p. 258)) writes:
As we look out into the Universe and identify the many accidents of physics
and astronomy that have worked together to our benefit, it almost seems as
if the Universe must in some sense have known that we were coming.
Here obviously to evolution and to the Universe we find ascribed prop-
erties of the Absolute . . .
It is generally believed that the increasing complexity of evolution
implies an increase of information (or a decrease of entropy). Does this
imply a continuous supply of information “from the outside”, some kind
of “continuous creation”? Tresmontant (1976) and others think so.
The old clash between religion and science was largely due to White-
head’s “fallacy of misplaced concreteness”. The exaggerated “spiritual-
ism” of religion seemed to be incompatible with the exaggerated “ma-
terialism” of science. As we have seen, materialism no longer reigns
supreme in modern physics. In relativity and quantum theory, science
has proved to be infinitely more complex and counterintuitive than the
most daring speculations of theologians and metaphysicists.
In fact, the main importance of modern science for religious thinking
may be the fact that it has provided highly counterintuitive and strange
models which make the old difficulties of religion (miracles etc.) look
harmless in comparison. No one any longer believes that God really
stopped the movement of the Sun in order to permit Joshua to win his
battle against the Amorites in daylight: the bible is a religious work and
274 CHAPTER 6. IMPLICATIONS OF SCIENCE
6.8 Pluralism
Die Wahrheit ist symphonisch.
Hans Urs von Balthasar
to fit into one single system, even too complicated to fit into a common
physical theory.
A high mountain looks different as we regard it from north, south,
east, or west. Which one is the “correct” perspective? Is the view from
north “true” and the view from east “false”? Should the observer from
south critize the viewer from the west?
In philosophy, however, it is customary for the adherents of one
system to critize their colleagues from the other systems. Such criti-
cism is necessary in order to make the reasoning logically stringent and
convincing. It should be an objective and fair criticism, preceded by
intense and sympathetic study of the other’s complete reasoning.
Superficial criticism based on sentence–by–sentence refutation is
easy and cheap. It is not difficult to condemn existentialism on the
basis of logical positivism and vice versa. A particularly shocking mis-
understanding regarding Fichte was mentioned in sec. 5.4.
It is quite improbable that great philosophical systems are com-
pletely erroneous. They at least contain some important truth or in-
sight.
My general experience with philosophical criticism is that positive
opinions of philosophers about other philosophers seem usually right,
and that negative judgments of philosophers about other philosophers
seem usually wrong. Why? I don’t know, it might be related to human
psychology.
All this speaks in favor of tolerance and scientific and philosoph-
ical pluralism. The plurality of philosophical systems is as necessary
for human understanding as the plurality of musical instruments is for
an orchestra. Figuratively speaking, Plato’s violin is, of course, om-
nipresent. But Fichte’s flute and Hegel’s trumpet add color, and the
cello of Kant and the contrabass of logical positivism provide a reliable
basis. The exotic instruments of Gödel and Wittgenstein are used if
the orchestra is to play Schönberg and Webern . . .
Thus Hans Urs von Balthasar says (see the motto of this section):
“TRUTH IS SYMPHONIC”.
276 CHAPTER 6. IMPLICATIONS OF SCIENCE
Selected additional reading
It is quite clear to me that you will probably neither be able to read all the
books given, nor even wish to do so. It is a rich menu, from which you are
invited to select at your convenience.
You are asking for a suggestion? If you can read only one book, I suggest
(Davies 1988), if you have time for two books, take also (Whitehead 1925),
and as a third, (Penrose 1989), which is tougher but extremely rewarding.
My next candidates on the reading list would be (Hofstadter 1979), (Cohen
and Stewart 1994), and (Popper 1982). Further recommendations are found
in the text of the book.
A triad of very readable and mutually complementary introductory book-
lets on “pure” philosophy is (Bocheński 1959), (Jaspers 1953) and (Russell
1912).
As far as available to the author, book titles have been given in English.
Several German and French books are available in an English translation
(and vice versa).
Abraham R H and Shaw C D (1992) Dynamics: The Geometry of Behavior,
2nd ed, Addison–Wesley, Redwood City, Cal.
Anger G, Gorenflo R, Jochmann H, Moritz H, and Webers W eds (1993)
Inverse Problems: Principles and Applications in Geophysics, Tech-
nology, and Medicine, Akademie Verlag, Berlin (mathematical).
Barrow J D (1988) The World Within the World, Oxford Univ. Press.
Barrow J D (1991) Theories of Everything, Oxford Univ. Press.
Barrow J D (1992) Pi in the Sky, Oxford Univ. Press.
Blokhintsev D I (1968) The Philosophy of Quantum Mechanics, Reidel,
Dordrecht, Holland (mathematical but also of general interest).
Bocheński J M (1959) Wege zum philosophischen Denken, Herder,
Freiburg.
Bohm D (1980) Wholeness and the Implicate Order, Routledge, London.
Bohr N (1934) Atomic Theory and the Description of Nature, Cambridge
Univ. Press.
Bohr N (1958) Atomic Physics and Human Knowledge, Wiley, New York.
Bohr N (1963) Essays 1958–1962 on Atomic Physics and Human Knowl-
edge, Wiley, London. (All three volumes reprinted 1987 by Ox Bow
Press, Woodbridge, Connecticut.)
277
278 ADDITIONAL READING
Briggs J (1992) Fractals: the Patterns of Chaos, Simon and Schuster, New
York.
Čapek M ed (1976) The Concepts of Space and Time, Reidel, Dordrecht,
Holland.
Capra F (1975) The Tao of Physics, Wildwood House, London.
Carnap R (1950) Logical Foundations of Probability, Chicago Univ. Press
(for specialists).
Carnap R (1958) Introduction to Symbolic Logic and Its Applications,
Dover, New York (for specialists).
Carnap R (1966) Philosophical Foundations of Physics, Basic Books, New
York.
Churchland P M (1988) Matter and Consciousness, 2nd ed., MIT Press.
Cohen L J (1986) The Dialogue of Reason, Oxford Univ. Press.
Cohen L J (1989) An Introduction to the Philosophy of Induction and
Probability, Oxford Univ. Press.
Cohen J and Stewart I (1994) The Collapse of Chaos: Discovering Sim-
plicity in a Complex World, Viking–Penguin, New York.
Copleston F (1946) A History of Philosophy, 9 vols., Burns and Oates,
London.
Davies P (1983) God and the New Physics, Simon and Schuster, New York.
Davies P (1984) Superforce: The Search for a Great Unified Theory of
Nature, Simon and Schuster, New York.
Davies P (1988) The Cosmic Blueprint, Simon and Schuster, New York.
Davies P (1992) The Mind of God: The Scientific Basis for a Rational
World, Simon and Schuster, New York.
Davies P and Brown J R (1986) The Ghost in the Atom, Cambridge Univ.
Press.
Davies P and Brown J R (1988) Superstrings: A Theory of Everything?
Cambridge Univ. Press.
Ditfurth H von (1976) Der Geist fiel nicht vom Himmel, Hoffmann und
Campe, Hamburg.
Duncan R and Weston–Smith M eds (1977) The Encyclopedia of Ignorance,
Pergamon, Oxford.
Eccles J C (1973) The Understanding of the Brain, McGraw–Hill, New
York.
Eccles J C (1989) Evolution of the Brain: Creation of the Self, Routledge,
London.
Eccles J C (1994) How the Self Controls Its Brain, Springer, Berlin.
Eddington A (1939) The Philosophy of Physical Science, Cambridge Univ.
Press.
Eddington A (1959) New Pathways in Science, Ann Arbor Paperback, Univ.
of Michigan Press.
Edelman G M (1989) The Remembered Present: A Biological Theory of
Consciousness, Basic Books, New York.
ADDITIONAL READING 279
Hawking S (1993) Black Holes and Baby Universes and Other Essays,
Bantam Press, London.
Hawking S W and Ellis G F R (1973) The Large Scale Structure of Space
Time, Cambridge Univ. Press (mathematical).
Heintel E (1990) Die Stellung der Philosophie in der “Universitas Litter-
arum”, Publishing House of the Austrian Academy of Sciences, Wien.
Heisenberg W (1955) Das Naturbild der heutigen Physik, Rowohlt, Ham-
burg.
Heisenberg W (1958) Physics and Philosophy, Harper & Row, New York.
Heisenberg W (1973) Der Teil und das Ganze, Piper, München.
Hempel C G (1966) Philosophy of Natural Science, Prentice–Hall, Engle-
wood Cliffs, N.J.
Hofstadter D R (1979) Gödel, Escher, Bach: an Eternal Golden Braid,
Basic Books, New York.
Hofstadter D R (1985) Metamagical Themas: Questing for the Essence of
Mind and Pattern, Basic Books, New York.
Hofstadter D R and Dennett D C, eds (1981) The Mind’s I, Basic Books,
New York.
Holzmüller W (1984) Information in Biological Systems: The Role of
Macromolecules, Cambridge Univ. Press.
Hubel D H (1988) Eye, Brain, and Vision, Scientific American Library, New
York.
Huxley A (1970) The Perennial Philosophy, Harper & Row, New York.
Jantsch E (1980) The Self–Organizing Universe, Pergamon, Oxford.
Jaspers K (1953) Einführung in die Philosophie, Piper, München.
Jaspers K (1962–1993) The Great Philosophers, Harcourt Brace Jo-
vanovich, New York.
Jeffreys H (1961) Theory of Probability, 3rd ed., Oxford Univ. Press (math-
ematical).
Jeffreys H (1973) Scientific Inference, 3rd ed., Cambridge Univ. Press.
Jordan P (1968) Der Naturwissenschaftler vor der religiösen Frage, 5th ed.,
Stalling, Oldenburg.
Kaku M (1994) Hyperspace: A Scientific Odyssey Through Parallel Uni-
verses, Time Warps, and the 10th Dimension, Oxford Univ. Press.
Kauffman S A (1993) The Origins of Order: Self–Organization and Selec-
tion in Evolution, Oxford Univ. Press.
Kohonen T (1988) Self–Organization and Associative Memory, 2nd ed.,
Springer, Berlin.
Kolakowski L (1982) Religion — If there is no God, Fontana Paperbacks,
London.
Koltermann R (1994) Grundzüge der modernen Naturphilosophie, Knecht,
Frankfurt a. M.
Kosko B (1993) Fuzzy Thinking, Hyperion, New York.
ADDITIONAL READING 281
Weizsäcker C F von (1970) Zum Weltbild der Physik, 11th ed., Hirzel,
Stuttgart.
Weizsäcker C F von (1971) Die Einheit der Natur, Hanser, München.
Weizsäcker C F von (1973) Die Tragweite der Wissenschaft, 4th ed., Hirzel,
Stuttgart.
Weizsäcker C F von (1977) Der Garten des Menschlichen, Hanser,
München.
Weizsäcker C F von (1985) Aufbau der Physik, Hanser, München.
Weizsäcker C F von (1992) Zeit und Wissen, Hanser, München.
Weyl H (1949) Philosophy of Mathematics and Natural Science, Princeton
Univ. Press.
Weyl H (1952) Symmetry, Princeton Univ. Press.
Wheeler J A (1994) At Home in the Universe. American Institute of
Physics, Woodbury, New York.
Whitehead A N (1925) Science and the Modern World, Macmillan, New
York.
Whitehead A N (1929) Process and Reality, Macmillan, New York (diffi-
cult!).
Whitehead A N (1933) Adventures of Ideas, Macmillan, London.
Whitrow G J (1980) The Natural Philosophy of Time, 2nd ed., Oxford
Univ. Press.
Wiener N (1961) Cybernetics, 2nd ed., Wiley, New York.
Will C M (1986) Was Einstein Right? Basic Books, New York.
Young J Z (1987) Philosophy and the Brain, Oxford Univ. Press.
Zurek W H (1990) Complexity, Entropy and the Physics of Information,
Addison–Wesley, Redwood City, California.
Index
285
286 INDEX
161, 165, 172, 173, 175– fuzzy logic, 38, 40, 57, 210, 254
177, 183, 214, 217, 224, fuzzy proposition, 40
239, 241, 245, 260, 262, fuzzy set, 39, 40, 42
266, 271–273
exact, 38, 57, 78, 81, 83, 90, 183, Galilei G , 73, 150
200, 202, 210–213, 231, Galilei transformation, 91, 92
252–256, 263 Gardner M , 132, 134
gauge theory, 118, 119, 192, 261,
fallacy of misplaced concreteness, 264, 265
12, 202, 206, 229, 237, 273 Gauss C F , 19, 40, 76, 95, 151,
fallacy of simple location, 201, 217 184, 186, 255
false, 28, 29, 36, 40, 41, 43 Gauss principle of least constraint,
falsification, 20, 110, 139, 140, 145, 75, 165
150, 151, 155, 215, 216, Gaussian curvature, 95, 96
233, 254, 255, 257, 261, Gaussian error curve, 184, 254
262 Gell–Mann M , 107, 117, 119, 171,
FBI, 193 172, 192, 267
feedback, 5, 159–163, 180, 250, 251 gene, 169, 176, 177, 180
Ferguson K , 271 general covariance, 46, 150, 157
Fermat P de , 35, 208 general relativity, 46, 55, 94–99,
Fermat’s principle, 164, 165, 246 123, 129, 140, 150, 157,
fermion, 114, 115 201, 231, 237, 241, 252,
Feynman R P , 33, 107, 132, 165 253, 257, 260, 261, 264
Fichte J G , 11, 45, 51, 55, 57, 220– generalized inverse, 144
222, 224, 232 genetic information, 175, 176
final causation, 79, 246 genome, 176, 179, 180
finalism, 79, 164, 165 Gentzen G , 35
finalistic, 79, 161, 165 geodesic, 75, 76
Findley J N , 44, 50, 205, 212, 232, geodesy, 6, 39, 41, 87, 149, 150,
240, 270 255, 269
Fischer H , 274 geophysics, 135, 149, 150
Folse H J , 51, 59 German idealism, 156, 225
formal logic, 29, 36, 43–45, 50, 51, Gibbs J W , 80, 89
53, 54, 57, 58 Gilson E , 14, 239
formalized thinking, 44 Glashow S , 118, 264
fractals, 83 Glass L , 83
free will, freedom of the will, 121, Gleick J , 82, 261
191, 212, 228, 230, 241, Globus G G , 13, 53, 164, 219, 237
245–251, 266 Gnedenko B V , 90
Frege G , 27, 34, 156 God, 11, 14, 60, 158, 189, 202, 204,
frequency theory of probability, 87 206, 210, 214, 215, 218,
future, 120–123, 130, 132, 142, 226–230, 240, 241, 261,
170, 241, 242, 266 269–274
fuzzy, 38–43, 50, 57, 58, 152, 210– Gödel K , 34, 59, 128, 208, 233,
213, 254, 259 241, 242
290 INDEX
Gödel’s theorem, 4, 33, 34, 36, 44, Hilbert space, 62, 69, 87, 100, 105,
56, 58, 128, 158, 193, 212, 200
222, 248, 254 Hipparchus , 257
Gödelian uncertainty, 36, 266 Hittmayr O , 234
Goethe J W von , 134, 228, 230, Hofmann–Wellenhof B , 94, 99,
273 101, 257, 269
Gold T , 241 Hofstadter D R , 8, 14, 36, 58, 59,
Götschl J , 181 183, 189, 192, 238
grandmother cells, 16, 17 holism, 30, 43, 46, 108, 189, 190,
gravitation, 74, 75, 97–99, 115, 192
118, 119, 256, 264, 265 holistic, 108, 109, 183, 203, 222
Gribbin J , 110 Holzmüller W , 49, 177
group, 33, 116–118 homeostasis, 5
Gulyga A , 45, 225 Hubel D H , 17
Gutmann V , 235 human perception, 14, 138–140,
204, 214
Hadamard J , 30, 149, 242 Hume D , 146, 233
Haken H , 164, 167–169, 171, 183 Huxley A , 270
Haldane J B S , 13, 49, 56 Huygens C , 166
Halmos P R , 26 hypercycle, 171, 179, 180, 182
Hamilton W , 76, 77 hypothalamus, 5, 159
Hamilton’s equations, 76, 77, 82, hypothesis, 20, 21, 260
100 hypothetic realism, 21
hardware, 8, 9, 138, 140, 159, 164,
187, 212, 238, 246 idealism, 10, 13, 45, 55, 158, 197,
Hartmann N , 21, 45, 217, 225, 231 198, 200, 201, 204, 220–
Hartshorne Ch , 270 222, 232, 237, 239, 240
Havemann R , 57, 107, 108, 122 identity theory, 10
Hawking S , 107, 112, 132, 134, ill–posed problem, 137, 143, 144,
233, 234, 248, 265–267, 146, 149, 202, 203
269 image processing, 15
Hegel G W F , 11, 45, 50, 51, 55, immanence, 269, 274
58, 69, 110, 204, 205, 221, immortality, 9, 12, 121, 230, 238,
222, 224, 226, 229, 232 273
Heidegger M , 274 improperly posed problem, 137,
Heintel E , 234 149
Heisenberg W , 100, 105, 106, 110, indeterminism, 242, 247, 249, 251
226, 234, 263, 264 induction, 90, 139, 140, 145–155,
Heisenberg uncertainty relation, 216, 233, 261, 262
37, 103, 152, 193, 223, inertial force, 96–98
254, 266, 267 inertial navigation, 99
hemisphere (brain), 5, 15 inexact concepts, 38, 39, 184
Heraclitus , 226 infinity, 28, 37, 38, 59, 62, 270
Hilbert D , 32, 34, 37, 59, 209 informal reasoning, 43, 44, 212
INDEX 291
symmetry breaking, 113, 169, 170 time, 74, 93, 94, 102, 120–124,
synapse, 7 126–134, 228, 240–242
synergetics, 156, 166, 167, 171, 179 time loop, 129, 130, 266
Synge J L , 98 time travel, 128, 130, 242
synthesis, 45, 47–49, 51, 53–55, 57, time–reversible, 80, 130
58, 98, 110, 123, 171, 198, Tipler F J , 126, 271
200, 220–222, 232, 241, TOE (Theory of Everything), 58,
252, 271, 274 146, 192, 261, 263–269
synthetic, 5, 18, 215, 230, 231, 266 tomography, 135, 136, 138
system identification, 141, 142 transcendence, 269
Szent–Györgyi A , 193 Treder H J , 99, 134, 221
Tresmontant C , 238, 273
triad
tautology, 30, 230 dialectic, 45, 51, 53, 98, 232,
Teilhard de Chardin P , 126, 183, 252, 274
204, 219, 271, 272 truth, 28–30, 32, 215, 262, 270,
TENTATIVE, 260, 261 271, 274, 275
thalamus, 5, 15 Turing A M , 169, 170
Thales , 225
theology, 126, 229, 271, 274 uncertainty relation, 37, 103, 152,
theorem of Pythagoras, 60, 61, 93 193, 223, 254, 266, 267
theoria, 155, 231, 256, 274 underdetermined problem, 142–
theory, 149–152, 154–158, 251–263 145
Theory of Everything, 58, 146, unified theory, 112, 113, 118, 119,
192, 261, 263–269 216, 249, 264, 265
theory of ideas, 226 union, 24, 25, 66, 67, 85
theory of knowledge, 3, 19, 137, unitary transformation, 103, 116
214, 217 unity of contraries, 53
evolutionary, 17, 20, 138 Universal Metaphysical Problem
thermodynamics, 171, 173, 175, Solver, 13, 202, 215, 229
177–179, 181, 182, 192, universe, 30, 78, 121–129, 223,
254, 260 224, 240–243, 267, 268,
thermostat, 5, 79, 159, 162, 165, 272, 273
246, 247, 250, 251 as an object, 223, 224
thermostat analogy, 250 creative, 123, 241, 242, 245
thesis, 45–49, 53, 54, 98, 123, 126, open, 123, 240–242, 245, 249
198, 232, 252, 274 unprovable, 35, 36, 44, 55, 212
unreasonable effectiveness of
thinking
mathematics (Wigner),
nonverbal, 30, 32
150, 154, 157, 255, 262
thinking about thinking, 50, 51, unstable, 81, 83, 137, 138, 167,
58, 220, 222, 268, 271, 274 169, 203, 242–244, 250
Thom R , 79, 170, 183 Urs von Balthasar H , 275
Thompson R F , 7 USEFUL, 260, 261
Thomson J J , 111
three–world model, 207 Varela F J , 21, 165, 225, 262
298 INDEX
verification, 139, 140, 145, 150, working hypothesis, 20, 156, 198,
151, 155, 215, 233, 261, 206, 215, 260
262 world–line, 120, 121, 129, 130, 218
vicious circle, 26, 36, 52, 204 World 1, 207–213, 242, 245, 250
vital force, 161, 182, 186, 188, 190 World 2, 207–213, 242, 245, 250
vitalism, 157, 186, 192 World 3, 207–212, 220, 222, 226,
Vollmer G , 21 242, 250