background image

M

A X  

B

O R N

The statistical interpretation of quantum

mechanics

Nobel Lecture, December 11, 1954

The work, for which I have had the honour to be awarded the Nobel Prize
for 1954, contains no discovery of a fresh natural phenomenon, but rather

the basis for a new mode of thought in regard to natural phenomena. This
way of thinking has permeated both experimental and theoretical physics to

such a degree that it hardly seems possible to say anything more about it that

has not been already so often said. However, there are some particular
aspects which I should like to discuss on what is, for me, such a festive occa-
sion. The first point is this: the work at the Göttingen school, which I

directed at that time (1926-I927), contributed to the solution of an intellec-
tual crisis into which our science had fallen as a result of Planck’s discovery
of the quantum of action in 1900. Today, physics finds itself in a similar
crisis - I do not mean here its entanglement in politics and economics as a
result of the mastery of a new and frightful force of Nature, but I am con-
sidering more the logical and epistemological problems posed by nuclear
physics. Perhaps it is well at such a time to recall what took place earlier in
a similar situation, especially as these events are not without a definite dramat-
ic flavour.

The second point I wish to make is that when I say that the physicists had

accepted the concepts and mode of thought developed by us at the time, I

am not quite correct. There are some very noteworthy exceptions, partic-
ularly among the very workers who have contributed most to building up
the quantum theory. Planck, himself, belonged to the sceptics until he died.
Einstein, De Broglie, and Schrödinger have unceasingly stressed the unsatis-
factory features of quantum mechanics and called for a return to the con-
cepts of classical, Newtonian physics while proposing ways in which this
could be done without contradicting experimental facts. Such weighty views
cannot be ignored. Niels Bohr has gone to a great deal of trouble to refute
the objections. I, too, have ruminated upon them and believe I can make
some contribution to the clarification of the position. The matter concerns
the borderland between physics and philosophy, and so my physics lecture

256

background image

I N T E R P R E T A T I O N   O F   Q U A N T U M   M E C H A N I C S

257

will partake of both history and philosophy, for which I must crave your

indulgence.

First of all, I will explain how quantum mechanics and its statistical inter-

pretation arose. At the beginning of the twenties, every physicist, I think,

was convinced that Planck’s quantum hypothesis was correct. According to
this theory energy appears in finite quanta of magnitude hv in oscillatory
processes having a specific frequency v

 

(e.g. in light waves). Countless

experiments could be explained in this way and always gave the same value

of Planck’s constant 

h

Again, Einstein’s assertion that light quanta have

momentum hv/c (where c is the speed of light) was well supported by exper-
iment (e.g. through the Compton effect). This implied a revival of the
corpuscular theory of light for a certain complex of phenomena. The wave
theory still held good for other processes. Physicists grew accustomed to this
duality and learned how to cope with it to a certain extent.

In 1913 Niels Bohr had solved the riddle of line spectra by means of the

quantum theory and had thereby explained broadly the amazing stability
of the atoms, the structure of their electronic shells, and the Periodic System
of the elements. For what was to come later, the most important assumption

of his teaching was this: an atomic system cannot exist in all mechanically
possible states, forming a continuum, but in a series of discrete « stationary »

states. In a transition from one to another, the difference in energy E

m

 

-

 is emitted or absorbed as a light quantum 

 (according to whether

E

m

 

is greater or less than E

n

). This is an interpretation in terms of energy

of the fundamental law of spectroscopy discovered some years before by
W. Ritz. The situation can be taken in at a glance by writing the energy

levels of the stationary states twice over, horizontally and vertically. This
produces a square array

E

I

, E

2

, E

. . . .

E

1

 

11 12 13 

-

E

2

21 22 23 

   

-

E

3

31 32 33 

   

-

-   -   -   -

in which positions on a diagonal correspond to states, and non-diagonal
positions correspond to transitions.

It was completely clear to Bohr that the law thus formulated is in conflict

with mechanics, and that therefore the use of the energy concept in this

background image

258 

1 9 5 4   M - B O R N

connection is problematical. He based this daring fusion of old and new on
his principle of correspondence. This consists in the obvious requirement that
ordinary classical mechanics must hold to a high degree of approximation
in the limiting case where the numbers of the stationary states, the so-called
quantum numbers, are very large (that is to say, far to the right and to the

lower part in the above array) and the energy changes relatively little from
place to place, in fact practically continuously.

Theoretical physics maintained itself on this concept for the next ten

years. The problem was this: an harmonic oscillation not only has a fre-

quency, but also an intensity. For each transition in the array there must be
a corresponding intensity. The question is how to find this through the
considerations of correspondence? It meant guessing the unknown from the
available information on a known limiting case. Considerable success was
attained by Bohr himself, by Kramers, Sommerfeld, Epstein, and many
others. But the decisive step was again taken by Einstein who, by a fresh

derivation of Planck’s radiation formula, made it transparently clear that the

classical concept of intensity of radiation must be replaced by the statistical

concept of transition probability. To each place in our pattern or array there

belongs (together with the frequency 

 = 

   

 a definite prob-

ability for the transition coupled with emission or absorption.

In Göttingen we also took part in efforts to distil the unknown mechanics

of the atom from the experimental results. The logical difficulty became ever
sharper. Investigations into the scattering and dispersion of light showed that
Einstein’s conception of transition probability as a measure of the strength of
an oscillation did not meet the case, and the idea of an amplitude of oscillation
associated with each transition was indispensable. In this connection, work by
Ladenburg

1

, Kramer

2

, Heisenberg

3

, Jordan and me

should be mentioned.

The art of guessing correct formulae, which deviate from the classical for-
mulae, yet contain them as a limiting case according to the correspondence
principle, was brought to a high degree of perfection. A paper of mine,

which introduced, for the first time I think, the expression 

quantum  

mechanics

in its title, contains a rather involved formula (still valid today) for the recip-
rocal disturbance of atomic systems.

Heisenberg, who at that time was my assistant, brought this period to a

sudden end

5

. He cut the Gordian knot by means of a philosophical prin-

ciple and replaced guess-work by a mathematical rule. The principle states
that concepts and representations that do not correspond to physically ob-
servable facts are not to be used in theoretical description. Einstein used the

background image

I N T E R P R E T A T I O N   O F   Q U A N T U M   M E C H A N I C S

259

same principle when, in setting up his theory of relativity, he eliminated the
concepts of absolute velocity of a body and of absolute simultaneity of two

events at different places. Heisenberg banished the picture of electron orbits
with definite radii and periods of rotation because these quantities are not
observable, and insisted that the theory be built up by means of the square
arrays mentioned above. Instead of describing the motion by giving a co-
ordinate as a function of time, x(t), an array of transition amplitudes 

should be determined. To me the decisive part of his work is the demand

to determine a rule by which from a given

array

the array for the square

can be found (or, more general, the multiplication rule for such arrays).

By observation of known examples solved by guess-work he found this

rule and applied it successfully to simple examples such as the harmonic and
anharmonic oscillator.

This was in the summer of 1925. Heisenberg, plagued by hay fever took

leave for a course of treatment by the sea and gave me his paper for publica-
tion if I thought I could do something with it.

The significance of the idea was at once clear to me and I sent the manu-

script to the Zeitschrift für Physik. I could not take my mind off Heisenberg’s

multiplication rule, and after a week of intensive thought and trial I suddenly
remembered an algebraic theory which I had learned from my teacher,
Professor Rosanes, in Breslau. Such square arrays are well known to math-
ematicians and, in conjunction with a specific rule for multiplication, are
called matrices. I applied this rule to Heisenberg’s quantum condition and
found that this agreed in the diagonal terms. It was easy to guess what the
remaining quantities must be, namely, zero; and at once there stood before

me the peculiar formula

p q      = 

This meant that coordinates and momenta cannot be represented by
figure values but by symbols, the product of which depends upon the order

of multiplication - they are said to be « non-commuting ».

I was as excited by this result as a sailor would be who, after a long voyage,

sees from afar, the longed-for land, and I felt regret that Heisenberg was not

background image

260

1 9 5 4   M . B O R N

there. I was convinced from the start that we had stumbled on the right path.
Even so, a great part was only guess-work, in particular, the disappearance
of the non-diagonal elements in the above-mentioned expression. For help

in this problem I obtained the assistance and collaboration of my pupil
Pascual Jordan, and in a few days we were able to demonstrate that I had
guessed correctly. The joint paper by Jordan and myself

6

 contains the most

important principles of quantum mechanics including its extension to elec-
trodynamics. There followed a hectic period of collaboration among the

three of us, complicated by Heisenberg’s absence. There was a lively ex-
change of letters; my contribution to these, unfortunately, have been lost in
the political disorders. The result was a three-author paper

7

 which brought

the formal side of the investigation to a definite conclusion. Before this paper

appeared, came the first dramatic surprise: Paul Dirac’s paper on the same
subject

8

. The inspiration afforded by a lecture of Heisenberg’s in Cambridge

had led him to similar results as we had obtained in Göttingen except that
he did not resort to the known matrix theory of the mathematicians, but
discovered the tool for himself and worked out the theory of such non-
commutating symbols.

The first non-trivial and physically important application of quantum

mechanics was made shortly afterwards by W. Pauli

who calculated the

stationary energy values of the hydrogen atom by means of the matrix method
and found complete agreement with Bohr’s formulae. From this moment
onwards there could no longer be any doubt about the correctness of the
theory.

What this formalism really signified was, however, by no means clear.

Mathematics, as often happens, was cleverer than interpretative thought.
While we were still discussing this point there came the second dramatic
surprise, 
the appearance of Schrödinger’s famous papers

10

. He took up quite

a different line of thought which had originated from Louis de Broglie

11

.

A few years previously, the latter had made the bold assertion, supported

by brilliant theoretical considerations, that wave-corpuscle duality, familiar
to physicists in the case of light, must also be valid for electrons. To each
electron moving free of force belongs a plane wave of a definite wavelength
which is determined by Planck’s constant and the mass. This exciting disser-
tation by De Broglie was well known to us in Göttingen. One day in

1925 I received a letter from C. J. Davisson giving some peculiar results on

the reflection of electrons from metallic surfaces. I, and my colleague on the

experimental side, James Franck, at once suspected that these curves of

background image

I N T E R P R E T A T I O N   O F   Q U A N T U M   M E C H A N I C S 2 6 1

Davisson’s were crystal-lattice spectra of De Broglie’s electron waves, and
we made one of our pupils,

 

Elsasser

 

1 2

,

 

to investigate the matter. His result

provided the first preliminary confirmation of the idea of De Broglie’s, and
this was later proved independently by Davisson and Germer

 

13 

and G. P.

Thomson

 

14 

by systematic experiments.

But this acquaintance with De Broglie’s way of thinking did not lead us

to an attempt to apply it to the electronic structure in atoms. This was left

to Schrödinger. He extended De Broglie’s wave equation which referred to
force-free motion, to the case where the effect of force is taken into account,
and gave an exact formulation of the subsidiary conditions, already suggested

by De Broglie, to which the wave function 

ψ 

must be subjected, namely that

it should be single-valued and finite in space and time. And he was successful

in deriving the stationary states of the hydrogen atom in the form of those

monochromatic solutions of his wave equation which do not extend to
infinity.

For a brief period at the beginning of 1926, it looked as though there

were, suddenly, two self-contained but quite distinct systems of explanation
extant: matrix mechanics and wave mechanics. But Schrödinger himself
soon demonstrated their complete equivalence.

Wave mechanics enjoyed a very great deal more popularity than the

Göttingen or Cambridge version of quantum mechanics. It operates with a

wave function 

ψ, 

which in the case of one particle at least, can be pictured

in space, and it uses the mathematical methods of partial differential equations

which are in current use by physicists. Schrödinger thought that his wave
theory made it possible to return to deterministic classical physics. He propos-

ed (and he has recently emphasized his proposal anew’s), to dispense with the
particle representation entirely, and instead of speaking of electrons as par-
ticles, to consider them as a continuous density distribution 

 (or electric

density 

To us in Göttingen this interpretation seemed unacceptable in face of

well established experimental facts. At that time it was already possible to
count particles by means of scintillations or with a Geiger counter, and to
photograph their tracks with the aid of a Wilson cloud chamber.

It appeared to me that it was not possible to obtain a clear interpretation

of the 

ψ

-function, by considering bound electrons. I had therefore, as early

as the end of 1925, made an attempt to extend the matrix method, which
obviously only covered oscillatory processes, in such a way as to be

applicable to aperiodic processes. I was at that time a guest of the Mas-

background image

262

 

1 9 5 4   M . B O R N

sachusetts Institute of Technology in the USA, and I found there in Norbert
Wiener an excellent collaborator. In our joint paper

16 

we replaced the

matrix by the general concept of an operator, and thus made it possible to
describe aperiodic processes. Nevertheless we missed the correct approach.
This was left to Schrödinger, and I immediately took up his method since it

held promise of leading to an interpretation of the 

ψ

-function. Again an idea

of Einstein’s gave me the lead. He had tried to make the duality of particles -
light quanta or photons - and waves comprehensible by interpreting the

square of the optical wave amplitudes as probability density for the occur-
rence of photons. This concept could at once be carried over to the 

ψ

-func-

tion: 

 ought to represent the probability density for electrons (or other

particles). It was easy to assert this, but how could it be proved?

The atomic collision processes suggested themselves at this point. A swarm

of electrons coming from infinity, represented by an incident wave of

known intensity (i.e., 

 impinges upon an obstacle, say a heavy atom.

In the same way that a water wave produced by a steamer causes secondary
circular waves in striking a pile, the incident electron wave is partially
transformed into a secondary spherical wave whose amplitude of oscillation

ψ 

differs for different directions. The square of the amplitude of this wave at

a great distance from the scattering centre determines the relative probability
of scattering as a function of direction. Moreover, if the scattering atom it-

self is capable of existing in different stationary states, then Schrödinger’s
wave equation gives automatically the probability of excitation of these

states, the electron being scattered with loss of energy, that is to say, inelastic-
ally, as it is called. In this way it was possible to get a theoretical basis

17 

for

the assumptions of Bohr’s theory which had been experimentally confirmed
by Franck and Hertz. Soon Wentzel

18 

succeeded in deriving Rutherford’s

famous formula for the scattering of 

α

-particles from my theory.

However, a paper by Heisenberg

19

, containing his celebrated uncertainty

relationship, contributed more than the above-mentioned successes to the
swift acceptance of the statistical interpretation of the 

ψ

-function. It was

through this paper that the revolutionary character of the new conception
became clear. It showed that not only the determinism of classical physics
must be abandonded, but also the naive concept of reality which looked upon
the particles of atomic physics as if they were very small grains of sand. At
every instant a grain of sand has a definite position and velocity. This is not

the case with an electron. If its position is determined with increasing ac-
curacy, the possibility of ascertaining the velocity becomes less and vice

background image

I N T E R P R E T A T I O N   O F   Q U A N T U M   M E C H A N I C S

263

versa. I

 

shall return shortly to these problems in a more general connection,

but would first like to say a few words about the theory of collisions.

The mathematical approximation methods which I used were quite prim-

itive and soon improved upon. From the literature, which has grown to a
point where I cannot cope with, I would like to mention only a few of

the first authors to whom the theory owes great progress: Faxén in

Sweden, Holtsmark in Norway

20

, Bethe in Germany

21

, Mott and Massey

in England

22

.

Today, collision theory is a special science with its own big, solid text-

books which have grown completely over my head. Of course in the last
resort all the modern branches of physics, quantum electrodynamics, the
theory of mesons, nuclei, cosmic rays, elementary particles and their trans-
formations, all come within range of these ideas and no bounds could be set
to a discussion on them.

I

 

should also like to mention that in 1926 and 1927 I tried another way of

supporting the statistical concept of quantum mechanics, partly in collabora-

tion with the Russian physicist Fock

2 3

. In the above-mentioned three-

author paper there is a chapter which anticipates the Schrödinger function,
except that it is not thought of as a function 

 in space, but as a function

 of the discrete index n = 1

 

,

 2

, . . .

 

which enumerates the stationary states.

If the system under consideration is subject to a force which is variable with
time,   becomes also time-dependent, and  

 (t) 

 signifies the prob-

ability for the existence of the state at time t. Starting from an initial
distribution where there is only one state, transition probabilities are ob-
tained, and their properties can be examined. What interested me in partic-
ular at the time, was what occurs in the adiabatic limiting case, that is, for
very slowly changing action. It was possible to show that, as could have been

expected, the probability of transitions becomes ever smaller. The theory of

transition probabilities was developed independently by Dirac with great
success. It can be said that the whole of atomic and nuclear physics works
with this system of concepts, particularly in the very elegant form given to

them by Dirac

2 4

. Almost all experiments lead to statements about relative

frequencies of events, even when they occur concealed under such names as
effective cross section or the like.

How does it come about then, that great scientists such as Einstein, Schrö-

dinger, and De Broglie are nevertheless dissatisfied with the situation? Of
course, all these objections are levelled not against the correctness of the

formulae, but against their interpretation. Two closely knitted points of view

background image

264

 

1 9 5 4   M . B O R N

are to be distinguished: the question of determinism and the question of reality.

Newtonian mechanics is deterministic in the following sense:
If the initial state (positions and velocities of all particles) of a system is

accurately given, then the state at any other time (earlier or later) can be
calculated from the laws of mechanics. All the other branches of classical
physics have been built up according to this model. Mechanical determinism

gradually became a kind of article of faith: the world as a machine, an
automaton. As far as I can see, this idea has no forerunners in ancient and

medieval philosophy. The idea is a product of the immense success of
Newtonian mechanics, particularly in astronomy. In the 

19th

 

century it be-

came a basic philosophical principle for the whole of exact science. I asked
myself whether this was really justified. Can absolute predictions really be
made for all time on the basis of the classical equations of motion? It can
easily be seen, by simple examples, that this is only the case when the
possibility of absolutely exact measurement (of position, velocity, or other
quantities) is assumed. Let us think of a particle moving without friction on

a straight line between two end-points (walls), at which it experiences com-
pletely elastic recoil. It moves with constant speed equal to its initial speed
v

backwards and forwards, and it can be stated exactly where it will be at

a given time provided that v

is accurately known. But if a small inaccuracy

 is allowed, then the inaccuracy of prediction of the position at time is

     which increases with t. If one waits long enough until time   = 

where  l  is the distance between the elastic walls, the inaccuracy 

 will have

become equal to the whole space l. Thus it is impossible to forecast anything
about the position at a time which is later than t

c

Thus determinism lapses

completely into indeterminism as soon as the slightest inaccuracy in the data
on velocity is permitted. Is there any sense - and I mean any physical sense,
not metaphysical sense - in which one can speak of absolute data? Is one

justified in saying that the coordinate x = 

Ï€ 

cm where 

Ï€ 

3.1415. . 

is the

familiar transcendental number that determines the ratio of the circumfer-
ence of a circle to its diameter? As a mathematical tool the concept of a real
number represented by a nonterminating decimal fraction is exceptionally

important and fruitful. As the measure of a physical quantity it is nonsense. If

Ï€ 

is taken to the 20th or the 25th place of decimals, two numbers are ob-

tained which are indistinguishable from each other and the true value of 

Ï€

by any measurement. According to the heuristic principle used by Einstein
in the theory of relativity, and by Heisenberg in the quantum theory, con-
cepts which correspond to no conceivable observation should be eliminated

background image

I N T E R P R E T A T I O N   O F   Q U A N T U M   M E C H A N I C S

265

from physics. This is possible without difficulty in the present case also. It is
only necessary to replace statements like 

x

 

Ï€ 

cm by: the probability of

distribution of values of x

 

has a sharp maximum at 

x

 

Ï€ 

cm; and (if it is

desired to be more accurate) to add: of such and such a breadth. In short,
ordinary mechanics must also be statistically formulated. I have occupied
myself with this problem a little recently, and have realized that it is pos-
sible without difficulty. This is not the place to go into the matter more

deeply. I

 

should like only to say this: the determinism of classical physics

turns out to be an illusion, created by overrating mathematico-logical con-

cepts. It is an idol, not an ideal in scientific research and cannot, therefore,

be used as an objection to the essentially indeterministic statistical interpre-
tation of quantum mechanics.

Much more difficult is the objection based on reality. The concept of a

particle, e.g. a grain of sand, implicitly contains the idea that it is in a definite

position and has definite motion. But according to quantum mechanics it
is impossible to determine simultaneously with any desired accuracy both
position and velocity (more precisely : momentum, i.e. mass times velocity) .
Thus two questions arise: what prevents us, in spite of the theoretical asser-
tion, to measure both quantities to any desired degree of accuracy by refined

experiments? Secondly, if it really transpires that this is not feasible, are we

still justified in applying to the electron the concept of particle and therefore
the ideas associated with it?

Referring to the first question, it is clear that if the theory is correct - and

we have ample grounds for believing this - the obstacle to simultaneous
measurement of position and motion (and of other such pairs of so-called

conjugate quantities) must lie in the laws of quantum mechanics themselves.
In fact, this is so. But it is not a simple matter to clarify the situation. Niels
Bohr himself has gone to great trouble and ingenuity

25 

to develop a theory

of measurements to clear the matter up and to meet the most refined and

ingenious attacks of Einstein, who repeatedly tried to think out methods of

measurement by means of which position and motion could be measured

simultaneously and accurately. The following emerges: to measure space
coordinates and instants of time, rigid measuring rods and clocks are re-
quired. On the other hand, to measure momenta and energies, devices are
necessary with movable parts to absorb the impact of the test object and to
indicate the size of its momentum. Paying regard to the fact that quantum

mechanics is competent for dealing with the interaction of object and appa-
ratus, it is seen that no arrangement is possible that will fulfil both require-

background image

266

1 9 5 4   M . B O R N

ments simultaneously. There exist, therefore, mutually exclusive though
complementary experiments which only as a whole embrace everything
which can be experienced with regard to an object.

This idea of complementarity is now regarded by most physicists as the

key to the clear understanding of quantum processes. Bohr has generalized
the idea to quite different fields of knowledge, e.g. the connection be-

tween consciousness and the brain, to the problem of free will, and other
basic problems of philosophy. To come now to the last point: can we call
something with which the concepts of position and motion cannot be asso-
ciated in the usual way, a thing, or a particle? And if not, what is the reality
which our theory has been invented to describe?

The answer to this is no longer physics, but philosophy, and to deal with

it thoroughly would mean going far beyond the bounds of this lecture. I
have given my views on it elsewhere

2 6

. Here I will only say that I am em-

phatically in favour of the retention of the particle idea. Naturally, it is
necessary to redefine what is meant. For this, well-developed concepts are
available which appear in mathematics under the name of invariants in trans-
formations. Every object that we perceive appears in innumerable aspects.
The concept of the object is the invariant of all these aspects. From this point

of view, the present universally used system of concepts in which particles
and waves appear simultaneously, can be completely justified.

The latest research on nuclei and elementary particles has led us, how-

ever, to limits beyond which this system of concepts itself does not appear to
suffice. The lesson to be learned from what I have told of the origin of
quantum mechanics is that probable refinements of mathematical methods

will not suffice to produce a satisfactory theory, but that somewhere in our
doctrine is hidden a concept, unjustified by experience, which we must elim-

inate to open up the road.

1. 

R. Ladenburg, Z. Physik

4 (192.1) 451

R. Ladenburg and F. Reiche, 

Naturwiss.,

11 (1923 ) 584.

2.

 

H. A. Kramers,

 

Nature , 113 (1924) 673.

3. H. A. Kramers and W. Heisenberg, Z

Physik, 

31

 

(1925) 681.

4. M. Born, Z. 

Physik, 

26 (1924)

 

379; M. Born and P. Jordan, Z. 

Physik, 33 

(1925)

479.

5.

 

W. Heisenberg,

 

Z. 

Physik, 

33 (1925) 879.

6.

 

M. Born and P.

 

Jordan, Z. 

Physik,  34 

(1925)

      

858.

background image

I N T E R P R E T A T I O N   O F   Q U A N T U M   M E C H A N I C S

267

7. M. Born, W. Heisenberg, and P. Jordan, Z. Physik, 35 (1926) 557.

8. 

P. A. M. 

Dirac, Proc. Roy. Soc. (London), A 109 (1925) 642.

9. W. Pauli, Z. Physik, 36 (1926) 336.

10. E. Schrödinger, Ann. Physik, [4] 79 (1926) 361,489,734; 80 (1926) 437; 81(1926)

109.

11. L. de Broglie, Thesis Paris, 1924; Ann. Phys. (Paris), [10]

 

3 (1925) 22.

12.

 

W. 

Elasser, 

Naturwiss., 13 (1925) 711.

13. C. J. Davisson and L. H. Germer, Phys. Rev., 30 (1927) 707.
14. G. P. Thomson and A. Reid, Nature, 119 (1927) 890; G. P. Thomson, Proc. Roy.

Soc. (London), A 117 (1928) 600.

15. E. Schrödinger, Brit. J. Phil. Sci., 3 (1952) 109, 233.
16. M. Born and N. Wiener, Z. Physik, 36 (1926) 174.
17. M. Born, Z. Physik, 37 (1926) 863 ; 38 (1926) 803 ; Göttinger Nachr. Math. Phys.

Kl., (1926) 146.

18. G. Wentzel, Z. Physik, 40 (1926) 590.
19. W. Heisenberg, Z. Physik, 43 (1927) 172.

20.

 

H. Faxén and J. Holtsmark, Z. Physik, 45 (1927) 307.

21. H. Bethe, Ann. Physik, 5 (1930) 325.
22.

 

N. F. Mott, Proc. Roy. Soc. (London), A 124 (1929) 422, 425; Proc. Cambridge Phil.

Soc., 25 (1929) 304.

23. M. Born, Z. Physik, 40 (1926) 167; M. Born and V. Fock, Z. Physik, 51 (1928)

165.

24. P. A. M. Dirac, Proc. Roy. Soc. (London), A 109 (1925) 642; 110

 

(1926) 561;

 

111

(1926) 281; 112 (26) 674.

25. N. Bohr, Naturwiss., 16 (1928) 245; 17 (1929) 483; 21 (1933)

 

13

 

«Kausalität und

Komplementarität» (Causality and Complementarity), Die Erkenntnis, 6 (1936) 293.

26. M. Born, Phil. Quart., 3 (1953) 134;

 

Physik. Bl., I0 (1954) 49.