You are on page 1of 394

Project Gutenberg's Five of Maxwell's Papers, by James Clerk Maxwell

This eBook is for the use of anyone anywhere at no cost and with
almost no restrictions whatsoever. You may copy it, give it away or
re-use it under the terms of the Project Gutenberg License included
with this eBook or online at www.gutenberg.org/license

Title: Five of Maxwell's Papers

Author: James Clerk Maxwell

Posting Date: February 25, 2014 [EBook #4908]

Release Date: January, 2004

[This file was first posted on March 24, 2002]

Language: English

*** START OF THIS PROJECT GUTENBERG EBOOK FIVE OF


MAXWELL'S PAPERS ***

Produced by Gordon Keener

This eBook includes 5 papers or speeches by James Clerk Maxwell.

The contents are:

Foramen Centrale
Theory of Compound Colours
1
Poinsot's Theory
Address to the Mathematical
Introductory Lecture

On the Unequal Sensibility of the Foramen Centrale to Light of


different Colours.

James Clerk Maxwell

[From the _Report of the British Association_, 1856.]

When observing the spectrum formed by looking at a long ve rtical slit


through a simple prism, I noticed an elongated dark spot running up
and down in the blue, and following the motion of the eye as it moved
_up and down_ the spectrum, but refusing to pass out of the blue into
the other colours. It was plain that the spot belonged both to the
eye and to the blue part of the spectrum. The result to which I have
come is, that the appearance is due to the yellow spot on the retina,
commonly called the _Foramen Centrale_ of Soemmering. The most
convenient method of observing the spot is by presenting to the eye in
not too rapid succession, blue and yellow glasses, or, still better,
allowing blue and yellow papers to revolve slowly before the eye. In
this way the spot is seen in the blue. It fades rapidly, but is
renewed every time the yellow comes in to relieve the effect of the
blue. By using a Nicol's prism along with this apparatus, the brushes
of Haidinger are well seen in connexion with the spot, and the fact of
the brushes being the spot analysed by polarized light becomes
evident. If we look steadily at an object behind a series of bright
bars which move in front of it, we shall see a curious bending of the
bars as they come up to the place of the yellow spot. The part which
comes over the spot seems to start in advance of the rest of the bar,
and this would seem to indicate a greater rapidity of sensation at the
yellow spot than in the surrounding retina. But I find the experiment
difficult, and I hope for better results from more accurate observers.

2
On the Theory of Compound Colours with reference to Mixtures of
Blue and Yellow Light.

James Clerk Maxwell

[From the _Report of the British Association_, 1856.]

When we mix together blue and yellow paint, we obtain green paint.
This fact is well known to all who have handled colours; and it is
universally admitted that blue and yellow make green. Red, yellow,
and blue, being the primary colours among painters, green is regarded
as a secondary colour, arising from the mixture of blue and yellow.
Newton, however, found that the green of the spectrum was not the same
thing as the mixture of two colours of the spectrum, for such a
mixture could be separated by the prism, while the green of the
spectrum resisted further decomposition. But still it was believed
that yellow and blue would make a green, though not that of the
spectrum. As far as I am aware, the first experiment on the subject
is that of M. Plateau, who, before 1819, made a disc with alternate
sectors of prussian blue and gamboge, and observed that, when
spinning, the resultant tint was not green, but a neutral gray,
inclining sometimes to yellow or blue, but never to green.
Prof. J. D. Forbes of Edinburgh made similar experiments in 1849, with
the same result. Prof. Helmholtz of Konigsberg, to whom we owe the
most complete investigation on visible colour, has given the true
explanation of this phenomenon. The result of mixing two coloured
powders is not by any means the same as mixing the beams of light
which flow from each separately. In the latter case we receive all
the light which comes either from the one powder or the other. In the
former, much of the light coming from one powder falls on particles of
the other, and we receive only that portion which has escaped
absorption by one or other. Thus the light coming from a mixture of
blue and yellow powder, consists partly of light coming directly from
blue particles or yellow particles, and partly of light acted on by
both blue and yellow particles. This latter light is green, since the
3
blue stops the red, yellow, and orange, and the yellow stops the blue
and violet. I have made experiments on the mixture of blue and yellow
light--by rapid rotation, by combined reflexion and transmission, by
viewing them out of focus, in stripes, at a great distance, by
throwing the colours of the spectrum on a screen, and by receiving
them into the eye directly; and I have arranged a portable apparatus
by which any one may see the result of this or any other mixture of
the colours of the spectrum. In all these cases blue and yellow do
not make green. I have also made experiments on the mixture of
coloured powders. Those which I used principally were "mineral blue"
(from copper) and "chrome-yellow." Other blue and yellow pigments gave
curious results, but it was more difficult to make the mixtures, and
the greens were less uniform in tint. The mixtures of these colours
were made by weight, and were painted on discs of paper, which were
afterwards treated in the manner described in my paper "On Colour as
perceived by the Eye," in the _Transactions of the Royal Society of
Edinburgh_, Vol. XXI. Part 2. The visible effect of the colour is
estimated in terms of the standard-coloured papers:--vermilion (V),
ultramarine (U), and emerald-green (E). The accuracy of the results,
and their significance, can be best understood by referring to the
paper before mentioned. I shall denote mineral blue by B, and
chrome-yellow by Y; and B3 Y5 means a mixture of three parts blue and
five parts yellow.

Given Colour. Standard Colours. Coefficient


V. U. E. of brightness.

B8 , 100 = 2 36 7 ............ 45

B7 Y1, 100 = 1 18 17 ............ 37

B6 Y2, 100 = 4 11 34 ............ 49

B5 Y3, 100 = 9 5 40 ............ 54

B4 Y4, 100 = 15 1 40 ............ 56

B3 Y5, 100 = 22 - 2 44 ............ 64

B2 Y6, 100 = 35 -10 51 ............ 76


4
B1 Y7, 100 = 64 -19 64 ............ 109

Y8, 100 = 180 -27 124 ............ 277

The columns V, U, E give the proportions of the standard colours which


are equivalent to 100 of the given colour; and the sum of V, U, E
gives a coefficient, which gives a general idea of the brightness. It
will be seen that the first admixture of yellow _diminishes_ the
brightness of the blue. The negative values of U indicate that a
mixture of V, U, and E cannot be made equivalent to the given colour.
The experiments from which these results were taken had the negative
values transferred to the other side of the equation. They were all
made by means of the colour-top, and were verified by repetition at
different times. It may be necessary to remark, in conclusion, with
reference to the mode of registering visible colours in terms of three
arbitrary standard colours, that it proceeds upon that theory of three
primary elements in the sensation of colour, which treats the
investigation of the laws of visible colour as a branch of human
physiology, incapable of being deduced from the laws of light itself,
as set forth in physical optics. It takes advantage of the methods of
optics to study vision itself; and its appeal is not to physical
principles, but to our consciousness of our own sensations.

On an Instrument to illustrate Poinsot's Theory of Rotation.

James Clerk Maxwell

[From the _Report of the British Association_, 1856.]

In studying the rotation of a solid body according to Poinsot's


method, we have to consider the successive positions of the
instantaneous axis of rotation with reference both to directions fixed
in space and axes assumed in the moving body. The paths traced out by
the pole of this axis on the _invariable plane_ and on the _central
5
ellipsoid_ form interesting subjects of mathematical investigation.
But when we attempt to follow with our eye the motion of a rotating
body, we find it difficult to determine through what point of the
_body_ the instantaneous axis passes at any time,--and to determine its
path must be still more difficult. I have endeavoured to render
visible the path of the instantaneous axis, and to vary the
circumstances of motion, by means of a top of the same kind as that
used by Mr Elliot, to illustrate precession*. The body of the
instrument is a hollow cone of wood, rising from a ring, 7 inches in
diameter and 1 inch thick. An iron axis, 8 inches long, screws into
the vertex of the cone. The lower extremity has a point of hard
steel, which rests in an agate cup, and forms the support of the
instrument. An iron nut, three ounces in weight, is made to screw on
the axis, and to be fixed at any point; and in the wooden ring are
screwed four bolts, of three ounces, working horizontally, and four
bolts, of one ounce, working vertically. On the upper part of the
axis is placed a disc of card, on which are drawn four concentric
rings. Each ring is divided into four quadrants, which are coloured
red, yellow, green, and blue. The spaces between the rings are white.
When the top is in motion, it is easy to see in which quadrant the
instantaneous axis is at any moment and the distance between it and
the axis of the instrument; and we observe,--1st. That the
instantaneous axis travels in a closed curve, and returns to its
original position in the body. 2ndly. That by working the vertical
bolts, we can make the axis of the instrument the centre of this
closed curve. It will then be one of the principal axes of inertia.
3rdly. That, by working the nut on the axis, we can make the order of
colours either red, yellow, green, blue, or the reverse. When the
order of colours is in the same direction as the rotation, it
indicates that the axis of the instrument is that of greatest moment
of inertia. 4thly. That if we screw the two pairs of opposite
horizontal bolts to different distances from the axis, the path of the
instantaneous pole will no longer be equidistant from the axis, but
will describe an ellipse, whose longer axis is in the direction of the
mean axis of the instrument. 5thly. That if we now make one of the
two horizontal axes less and the other greater than the vertical axis,
the instantaneous pole will separate from the axis of the instrument,
and the axis will incline more and more till the spinning can no
longer go on, on account of the obliquity. It is easy to see that, by
attending to the laws of motion, we may produce any of the above
6
effects at pleasure, and illustrate many different propositions by
means of the same instrument.

* _Transactions of the Royal Scottish Society of Arts_, 1855.

Address to the Mathematical and Physical Sections of the British


Association.

James Clerk Maxwell

[From the _British Association Report_, Vol. XL.]

[Liverpool, _September_ 15, 1870.]

At several of the recent Meetings of the British Association the


varied and important business of the Mathematical and Physical Section
has been introduced by an Address, the subject of which has been left
to the selection of the President for the time being. The perplexing
duty of choosing a subject has not, however, fallen to me.

Professor Sylvester, the President of Section A at the Exeter Meeting,


gave us a noble vindication of pure mathematics by laying bare, as it
were, the very working of the mathematical mind, and setting before
us, not the array of symbols and brackets which form the armoury of
the mathematician, or the dry results which are only the monuments of
his conquests, but the mathematician himself, with all his human
faculties directed by his professional sagacity to the pursuit,
apprehension, and exhibition of that ideal harmony which he feels to
be the root of all knowledge, the fountain of all pleasure, and the
condition of all action. The mathematician has, above all things, an
eye for symmetry; and Professor Sylvester has not only recognized the
symmetry formed by the combination of his own subject with those of
the former Presidents, but has pointed out the duties of his successor
in the following characteristic note:--

7
"Mr Spottiswoode favoured the Section, in his opening Address, with a
combined history of the progress of Mathematics and Physics; Dr.
Tyndall's address was virtually on the limits of Physical Philosophy;
the one here in print," says Prof. Sylvester, "is an attempted faint
adumbration of the nature of Mathematical Science in the abstract.
What is wanting (like a fourth sphere resting on three others in
contact) to build up the Ideal Pyramid is a discourse on the Relation
of the two branches (Mathematics and Physics) to, their action and
reaction upon, one another, a magnificent theme, with which it is to
be hoped that some future President of Section A will crown the
edifice and make the Tetralogy (symbolizable by _A+A'_, _A_, _A'_,
_AA'_) complete."

The theme thus distinctly laid down for his successor by our late
President is indeed a magnificent one, far too magnificent for any
efforts of mine to realize. I have endeavoured to follow Mr
Spottiswoode, as with far-reaching vision he distinguishes the systems
of science into which phenomena, our knowledge of which is still in
the nebulous stage, are growing. I have been carried by the
penetrating insight and forcible expression of Dr Tyndall into that
sanctuary of minuteness and of power where molecules obey the laws of
their existence, clash together in fierce collision, or grapple in yet
more fierce embrace, building up in secret the forms of visible
things. I have been guided by Prof. Sylvester towards those serene
heights

"Where never creeps a cloud, or moves a wind,


Nor ever falls the least white star of snow,
Nor ever lowest roll of thunder moans,
Nor sound of human sorrow mounts to mar
Their sacred everlasting calm."

But who will lead me into that still more hidden and dimmer region
where Thought weds Fact, where the mental operation of the
mathematician and the physical action of the molecules are seen in
their true relation? Does not the way to it pass through the very den
of the metaphysician, strewed with the remains of former explorers,
and abhorred by every man of science? It would indeed be a foolhardy
adventure for me to take up the valuable time of the Section by
leading you into those speculations which require, as we know,
8
thousands of years even to shape themselves intelligibly.

But we are met as cultivators of mathematics and physics. In our


daily work we are led up to questions the same in kind with those of
metaphysics; and we approach them, not trusting to the native
penetrating power of our own minds, but trained by a long-continued
adjustment of our modes of thought to the facts of external nature.

As mathematicians, we perform certain mental operations on the symbols


of number or of quantity, and, by proceeding step by step from more
simple to more complex operations, we are enabled to express the same
thing in many different forms. The equivalence of these different
forms, though a necessary consequence of self-evident axioms, is not
always, to our minds, self-evident; but the mathematician, who by long
practice has acquired a familiarity with many of these forms, and has
become expert in the processes which lead from one to another, can
often transform a perplexing expression into another which explains
its meaning in more intelligible language.

As students of Physics we observe phenomena under varied


circumstances, and endeavour to deduce the laws of their relations.
Every natural phenomenon is, to our minds, the result of an infinitely
complex system of conditions. What we set ourselves to do is to
unravel these conditions, and by viewing the phenomenon in a way which
is in itself partial and imperfect, to piece out its features one by
one, beginning with that which strikes us first, and thus gradually
learning how to look at the whole phenomenon so as to obtain a
continually greater degree of clearness and distinctness. In this
process, the feature which presents itself most forcibly to the
untrained inquirer may not be that which is considered most
fundamental by the experienced man of science; for the success of any
physical investigation depends on the judicious selection of what is
to be observed as of primary importance, combined with a voluntary
abstraction of the mind from those features which, however attractive
they appear, we are not yet sufficiently advanced in science to
investigate with profit.

Intellectual processes of this kind have been going on since the first
formation of language, and are going on still. No doubt the feature
which strikes us first and most forcibly in any phenomenon, is the
9
pleasure or the pain which accompanies it, and the agreeable or
disagreeable results which follow after it. A theory of nature from
this point of view is embodied in many of our words and phrases, and
is by no means extinct even in our deliberate opinions.

It was a great step in science when men became convinced that, in


order to understand the nature of things, they must begin by asking,
not whether a thing is good or bad, noxious or beneficial, but of what
kind is it? and how much is there of it? Quality and Quantity were
then first recognized as the primary features to be observed in
scientific inquiry.

As science has been developed, the domain of quantity has everywhere


encroached on that of quality, till the process of scientific inquiry
seems to have become simply the measurement and registration of
quantities, combined with a mathematical discussion of the numbers
thus obtained. It is this scientific method of directing our
attention to those features of phenomena which may be regarded as
quantities which brings physical research under the influence of
mathematical reasoning. In the work of the Section we shall have
abundant examples of the successful application of this method to the
most recent conquests of science; but I wish at present to direct your
attention to some of the reciprocal effects of the progress of science
on those elementary conceptions which are sometimes thought to be
beyond the reach of change.

If the skill of the mathematician has enabled the experimentalist to


see that the quantities which he has measured are connected by
necessary relations, the discoveries of physics have revealed to the
mathematician new forms of quantities which he could never have
imagined for himself.

Of the methods by which the mathematician may make his labours most
useful to the student of nature, that which I think is at present most
important is the systematic classification of quantities.

The quantities which we study in mathematics and physics may be


classified in two different ways.

The student who wishes to master any particular science must make
10
himself familiar with the various kinds of quantities which belong to
that science. When he understands all the relations between these
quantities, he regards them as forming a connected system, and he
classes the whole system of quantities together as belonging to that
particular science. This classification is the most natural from a
physical point of view, and it is generally the first in order of
time.

But when the student has become acquainted with several different
sciences, he finds that the mathematical processes and trains of
reasoning in one science resemble those in another so much that his
knowledge of the one science may be made a most useful help in the
study of the other.

When he examines into the reason of this, he finds that in the two
sciences he has been dealing with systems of quantities, in which the
mathematical forms of the relations of the quantities are the same in
both systems, though the physical nature of the quantities may be
utterly different.

He is thus led to recognize a classification of quantities on a new


principle, according to which the physical nature of the quantity is
subordinated to its mathematical form. This is the point of view
which is characteristic of the mathematician; but it stands second to
the physical aspect in order of time, because the human mind, in order
to conceive of different kinds of quantities, must have them presented
to it by nature.

I do not here refer to the fact that all quantities, as such, are
subject to the rules of arithmetic and algebra, and are therefore
capable of being submitted to those dry calculations which represent,
to so many minds, their only idea of mathematics.

The human mind is seldom satisfied, and is certainly never exercising


its highest functions, when it is doing the work of a calculating
machine. What the man of science, whether he is a mathematician or a
physical inquirer, aims at is, to acquire and develope clear ideas of
the things he deals with. For this purpose he is willing to enter on
long calculations, and to be for a season a calculating machine, if he
can only at last make his ideas clearer.
11
But if he finds that clear ideas are not to be obtained by means of
processes the steps of which he is sure to forget before he has
reached the conclusion, it is much better that he should turn to
another method, and try to understand the subject by means of
well-chosen illustrations derived from subjects with which he is more
familiar.

We all know how much more popular the illustrative method of


exposition is found, than that in which bare processes of reasoning
and calculation form the principal subject of discourse.

Now a truly scientific illustration is a method to enable the mind to


grasp some conception or law in one branch of science, by placing
before it a conception or a law in a different branch of science, and
directing the mind to lay hold of that mathematical form which is
common to the corresponding ideas in the two sciences, leaving out of
account for the present the difference between the physical nature of
the real phenomena.

The correctness of such an illustration depends on whether the two


systems of ideas which are compared together are really analogous in
form, or whether, in other words, the corresponding physical
quantities really belong to the same mathematical class. When this
condition is fulfilled, the illustration is not only convenient for
teaching science in a pleasant and easy manner, but the recognition of
the formal analogy between the two systems of ideas leads to a
knowledge of both, more profound than could be obtained by studying
each system separately.

There are men who, when any relation or law, however complex, is put
before them in a symbolical form, can grasp its full meaning as a
relation among abstract quantities. Such men sometimes treat with
indifference the further statement that quantities actually exist in
nature which fulfil this relation. The mental image of the concrete
reality seems rather to disturb than to assist their contemplations.
But the great majority of mankind are utterly unable, without long
training, to retain in their minds the unembodied symbols of the pure
mathematician, so that, if science is ever to become popular, and yet
remain scientific, it must be by a profound study and a copious
12
application of those principles of the mathematical classification of
quantities which, as we have seen, lie at the root of every truly
scientific illustration.

There are, as I have said, some minds which can go on contemplating


with satisfaction pure quantities presented to the eye by symbols, and
to the mind in a form which none but mathematicians can conceive.

There are others who feel more enjoyment in following geometrical


forms, which they draw on paper, or build up in the empty space before
them.

Others, again, are not content unless they can project their whole
physical energies into the scene which they conjure up. They learn at
what a rate the planets rush through space, and they experience a
delightful feeling of exhilaration. They calculate the forces with
which the heavenly bodies pull at one another, and they feel their own
muscles straining with the effort.

To such men momentum, energy, mass are not mere abstract expressions
of the results of scientific inquiry. They are words of power, which
stir their souls like the memories of childhood.

For the sake of persons of these different types, scientific truth


should be presented in different forms, and should be regarded as
equally scientific whether it appears in the robust form and the vivid
colouring of a physical illustration, or in the tenuity and paleness
of a symbolical expression.

Time would fail me if I were to attempt to illustrate by examples the


scientific value of the classification of quantities. I shall only
mention the name of that important class of magnitudes having
direction in space which Hamilton has called vectors, and which form
the subject-matter of the Calculus of Quaternions, a branch of
mathematics which, when it shall have been thoroughly understood by
men of the illustrative type, and clothed by them with physical
imagery, will become, perhaps under some new name, a most powerful
method of communicating truly scientific knowledge to persons
apparently devoid of the calculating spirit.

13
The mutual action and reaction between the different departments of
human thought is so interesting to the student of scientific progress,
that, at the risk of still further encroaching on the valuable time of
the Section, I shall say a few words on a branch of physics which not
very long ago would have been considered rather a branch of
metaphysics. I mean the atomic theory, or, as it is now called, the
molecular theory of the constitution of bodies.

Not many years ago if we had been asked in what regions of physical
science the advance of discovery was least apparent, we should have
pointed to the hopelessly distant fixed stars on the one hand, and to
the inscrutable delicacy of the texture of material bodies on the
other.

Indeed, if we are to regard Comte as in any degree representing the


scientific opinion of his time, the research into what takes place
beyond our own solar system seemed then to be exceedingly unpromising,
if not altogether illusory.

The opinion that the bodies which we see and handle, which we can set
in motion or leave at rest, which we can break in pieces and destroy,
are composed of smaller bodies which we cannot see or handle, which
are always in motion, and which can neither be stopped nor broken in
pieces, nor in any way destroyed or deprived of the least of their
properties, was known by the name of the Atomic theory. It was
associated with the names of Democritus, Epicurus, and Lucretius, and
was commonly supposed to admit the existence only of atoms and void,
to the exclusion of any other basis of things from the universe.

In many physical reasonings and mathematical calculations we are


accustomed to argue as if such substances as air, water, or metal,
which appear to our senses uniform and continuous, were strictly and
mathematically uniform and continuous.

We know that we can divide a pint of water into many millions of


portions, each of which is as fully endowed with all the properties of
water as the whole pint was; and it seems only natural to conclude
that we might go on subdividing the water for ever, just as we can
never come to a limit in subdividing the space in which it is
contained. We have heard how Faraday divided a grain of gold into an
14
inconceivable number of separate particles, and we may see Dr Tyndall
produce from a mere suspicion of nitrite of butyle an immense cloud,
the minute visible portion of which is still cloud, and therefore must
contain many molecules of nitrite of butyle.

But evidence from different and independent sources is now crowding in


upon us which compels us to admit that if we could push the process of
subdivision still further we should come to a limit, because each
portion would then contain only one molecule, an individual body, one
and indivisible, unalterable by any power in nature.

Even in our ordinary experiments on very finely divided matter we find


that the substance is beginning to lose the properties which it
exhibits when in a large mass, and that effects depending on the
individual action of molecules are beginning to become prominent.

The study of these phenomena is at present the path which leads to the
development of molecular science.

That superficial tension of liquids which is called capillary


attraction is one of these phenomena. Another important class of
phenomena are those which are due to that motion of agitation by which
the molecules of a liquid or gas are continually working their way
from one place to another, and continually changing their course, like
people hustled in a crowd.

On this depends the rate of diffusion of gases and liquids through


each other, to the study of which, as one of the keys of molecular
science, that unwearied inquirer into nature's secrets, the late Prof.
Graham, devoted such arduous labour.

The rate of electrolytic conduction is, according to Wiedemann's


theory, influenced by the same cause; and the conduction of heat in
fluids depends probably on the same kind of action. In the case of
gases, a molecular theory has been developed by Clausius and others,
capable of mathematical treatment, and subjected to experimental
investigation; and by this theory nearly every known mechanical
property of gases has been explained on dynamical principles; so that
the properties of individual gaseous molecules are in a fair way to
become objects of scientific research.
15
Now Mr Stoney has pointed out[1] that the numerical results of
experiments on gases render it probable that the mean distance of
their particles at the ordinary temperature and pressure is a quantity
of the same order of magnitude as a millionth of a millimetre, and Sir
William Thomson has since[2] shewn, by several independent lines of
argument, drawn from phenomena so different in themselves as the
electrification of metals by contact, the tension of soap-bubbles, and
the friction of air, that in ordinary solids and liquids the average
distance between contiguous molecules is less than the
hundred-millionth, and greater than the two-thousand-millionth of a
centimetre.

[1] _Phil. Mag._, Aug. 1868.


[2] _Nature_, March 31, 1870.

These, of course, are exceedingly rough estimates, for they are


derived from measurements some of which are still confessedly very
rough; but if at the present time, we can form even a rough plan for
arriving at results of this kind, we may hope that, as our means of
experimental inquiry become more accurate and more varied, our
conception of a molecule will become more definite, so that we may be
able at no distant period to estimate its weight with a greater degree
of precision.

A theory, which Sir W. Thomson has founded on Helmholtz's splendid


hydrodynamical theorems, seeks for the properties of molecules in the
ring vortices of a uniform, frictionless, incompressible fluid. Such
whirling rings may be seen when an experienced smoker sends out a
dexterous puff of smoke into the still air, but a more evanescent
phenomenon it is difficult to conceive. This evanescence is owing to
the viscosity of the air; but Helmholtz has shewn that in a perfect
fluid such a whirling ring, if once generated, would go on whirling
for ever, would always consist of the very same portion of the fluid
which was first set whirling, and could never be cut in two by any
natural cause. The generation of a ring-vortex is of course equally
beyond the power of natural causes, but once generated, it has the
properties of individuality, permanence in quantity, and
indestructibility. It is also the recipient of impulse and of energy,
which is all we can affirm of matter; and these ring-vortices are
16
capable of such varied connexions and knotted self-involutions, that
the properties of differently knotted vortices must be as different as
those of different kinds of molecules can be.

If a theory of this kind should be found, after conquering the


enormous mathematical difficulties of the subject, to represent in any
degree the actual properties of molecules, it will stand in a very
different scientific position from those theories of molecular action
which are formed by investing the molecule with an arbitrary system of
central forces invented expressly to account for the observed
phenomena.

In the vortex theory we have nothing arbitrary, no central forces or


occult properties of any other kind. We have nothing but matter and
motion, and when the vortex is once started its properties are all
determined from the original impetus, and no further assumptions are
possible.

Even in the present undeveloped state of the theory, the contemplation


of the individuality and indestructibility of a ring-vortex in a
perfect fluid cannot fail to disturb the commonly received opinion
that a molecule, in order to be permanent, must be a very hard body.

In fact one of the first conditions which a molecule must fulfil is,
apparently, inconsistent with its being a single hard body. We know
from those spectroscopic researches which have thrown so much light on
different branches of science, that a molecule can be set into a state
of internal vibration, in which it gives off to the surrounding medium
light of definite refrangibility--light, that is, of definite
wave-length and definite period of vibration. The fact that all the
molecules (say, of hydrogen) which we can procure for our experiments,
when agitated by heat or by the passage of an electric spark, vibrate
precisely in the same periodic time, or, to speak more accurately,
that their vibrations are composed of a system of simple vibrations
having always the same periods, is a very remarkable fact.

I must leave it to others to describe the progress of that splendid


series of spectroscopic discoveries by which the chemistry of the
heavenly bodies has been brought within the range of human inquiry. I
wish rather to direct your attention to the fact that, not only has
17
every molecule of terrestrial hydrogen the same system of periods of
free vibration, but that the spectroscopic examination of the light of
the sun and stars shews that, in regions the distance of which we can
only feebly imagine, there are molecules vibrating in as exact unison
with the molecules of terrestrial hydrogen as two tuning-forks tuned
to concert pitch, or two watches regulated to solar time.

Now this absolute equality in the magnitude of quantities, occurring


in all parts of the universe, is worth our consideration.

The dimensions of individual natural bodies are either quite


indeterminate, as in the case of planets, stones, trees, &c., or they
vary within moderate limits, as in the case of seeds, eggs, &c.; but
even in these cases small quantitative differences are met with which
do not interfere with the essential properties of the body.

Even crystals, which are so definite in geometrical form, are variable


with respect to their absolute dimensions.

Among the works of man we sometimes find a certain degree of


uniformity.

There is a uniformity among the different bullets which are cast in


the same mould, and the different copies of a book printed from the
same type.

If we examine the coins, or the weights and measures, of a civilized


country, we find a uniformity, which is produced by careful adjustment
to standards made and provided by the state. The degree of uniformity
of these national standards is a measure of that spirit of justice in
the nation which has enacted laws to regulate them and appointed
officers to test them.

This subject is one in which we, as a scientific body, take a warm


interest; and you are all aware of the vast amount of scientific work
which has been expended, and profitably expended, in providing weights
and measures for commercial and scientific purposes.

The earth has been measured as a basis for a permanent standard of


length, and every property of metals has been investigated to guard
18
against any alteration of the material standards when made. To weigh
or measure any thing with modern accuracy, requires a course of
experiment and calculation in which almost every branch of physics and
mathematics is brought into requisition.

Yet, after all, the dimensions of our earth and its time of rotation,
though, relatively to our present means of comparison, very permanent,
are not so by any physical necessity. The earth might contract by
cooling, or it might be enlarged by a layer of meteorites falling on
it, or its rate of revolution might slowly slacken, and yet it would
continue to be as much a planet as before.

But a molecule, say of hydrogen, if either its mass or its time of


vibration were to be altered in the least, would no longer be a
molecule of hydrogen.

If, then, we wish to obtain standards of length, time, and mass which
shall be absolutely permanent, we must seek them not in the
dimensions, or the motion, or the mass of our planet, but in the
wave-length, the period of vibration, and the absolute mass of these
imperishable and unalterable and perfectly similar molecules.

When we find that here, and in the starry heavens, there are
innumerable multitudes of little bodies of exactly the same mass, so
many, and no more, to the grain, and vibrating in exactly the same
time, so many times, and no more, in a second, and when we reflect
that no power in nature can now alter in the least either the mass or
the period of any one of them, we seem to have advanced along the path
of natural knowledge to one of those points at which we must accept
the guidance of that faith by which we understand that "that which is
seen was not made of things which do appear."

One of the most remarkable results of the progress of molecular


science is the light it has thrown on the nature of irreversible
processes--processes, that is, which always tend towards and never
away from a certain limiting state. Thus, if two gases be put into
the same vessel, they become mixed, and the mixture tends continually
to become more uniform. If two unequally heated portions of the same
gas are put into the vessel, something of the kind takes place, and
the whole tends to become of the same temperature. If two unequally
19
heated solid bodies be placed in contact, a continual approximation of
both to an intermediate temperature takes place.

In the case of the two gases, a separation may be effected by chemical


means; but in the other two cases the former state of things cannot be
restored by any natural process.

In the case of the conduction or diffusion of heat the process is not


only irreversible, but it involves the irreversible diminution of that
part of the whole stock of thermal energy which is capable of being
converted into mechanical work.

This is Thomson's theory of the irreversible dissipation of energy,


and it is equivalent to the doctrine of Clausius concerning the growth
of what he calls Entropy.

The irreversible character of this process is strikingly embodied in


Fourier's theory of the conduction of heat, where the formulae
themselves indicate, for all positive values of the time, a possible
solution which continually tends to the form of a uniform diffusion of
heat.

But if we attempt to ascend the stream of time by giving to its symbol


continually diminishing values, we are led up to a state of things in
which the formula has what is called a critical value; and if we
inquire into the state of things the instant before, we find that the
formula becomes absurd.

We thus arrive at the conception of a state of things which cannot be


conceived as the physical result of a previous state of things, and we
find that this critical condition actually existed at an epoch not in
the utmost depths of a past eternity, but separated from the present
time by a finite interval.

This idea of a beginning is one which the physical researches of


recent times have brought home to us, more than any observer of the
course of scientific thought in former times would have had reason to
expect.

But the mind of man is not, like Fourier's heated body, continually
20
settling down into an ultimate state of quiet uniformity, the
character of which we can already predict; it is rather like a tree,
shooting out branches which adapt themselves to the new aspects of the
sky towards which they climb, and roots which contort themselves among
the strange strata of the earth into which they delve. To us who
breathe only the spirit of our own age, and know only the
characteristics of contemporary thought, it is as impossible to
predict the general tone of the science of the future as it is to
anticipate the particular discoveries which it will make.

Physical research is continually revealing to us new features of


natural processes, and we are thus compelled to search for new forms
of thought appropriate to these features. Hence the importance of a
careful study of those relations between mathematics and Physics which
determine the conditions under which the ideas derived from one
department of physics may be safely used in forming ideas to be
employed in a new department.

The figure of speech or of thought by which we transfer the language


and ideas of a familiar science to one with which we are less
acquainted may be called Scientific Metaphor.

Thus the words Velocity, Momentum, Force, &c. have acquired certain
precise meanings in Elementary Dynamics. They are also employed in
the Dynamics of a Connected System in a sense which, though perfectly
analogous to the elementary sense, is wider and more general.

These generalized forms of elementary ideas may be called metaphorical


terms in the sense in which every abstract term is metaphorical. The
characteristic of a truly scientific system of metaphors is that each
term in its metaphorical use retains all the formal relations to the
other terms of the system which it had in its original use. The
method is then truly scientific--that is, not only a legitimate
product of science, but capable of generating science in its turn.

There are certain electrical phenomena, again, which are connected


together by relations of the same form as those which connect
dynamical phenomena. To apply to these the phrases of dynamics with
proper distinctions and provisional reservations is an example of a
metaphor of a bolder kind; but it is a legitimate metaphor if it
21
conveys a true idea of the electrical relations to those who have been
already trained in dynamics.

Suppose, then, that we have successfully introduced certain ideas


belonging to an elementary science by applying them metaphorically to
some new class of phenomena. It becomes an important philosophical
question to determine in what degree the applicability of the old
ideas to the new subject may be taken as evidence that the new
phenomena are physically similar to the old.

The best instances for the determination of this question are those in
which two different explanations have been given of the same thing.

The most celebrated case of this kind is that of the corpuscular and
the undulatory theories of light. Up to a certain point the phenomena
of light are equally well explained by both; beyond this point, one of
them fails.

To understand the true relation of these theories in that part of the


field where they seem equally applicable we must look at them in the
light which Hamilton has thrown upon them by his discovery that to
every brachistochrone problem there corresponds a problem of free
motion, involving different velocities and times, but resulting in the
same geometrical path. Professor Tait has written a very interesting
paper on this subject.

According to a theory of electricity which is making great progress in


Germany, two electrical particles act on one another directly at a
distance, but with a force which, according to Weber, depends on their
relative velocity, and according to a theory hinted at by Gauss, and
developed by Riemann, Lorenz, and Neumann, acts not instantaneously,
but after a time depending on the distance. The power with which this
theory, in the hands of these eminent men, explains every kind of
electrical phenomena must be studied in order to be appreciated.

Another theory of electricity, which I prefer, denies action at a


distance and attributes electric action to tensions and pressures in
an all-pervading medium, these stresses being the same in kind with
those familiar to engineers, and the medium being identical with that
in which light is supposed to be propagated.
22
Both these theories are found to explain not only the phenomena by the
aid of which they were originally constructed, but other phenomena,
which were not thought of or perhaps not known at the time; and both
have independently arrived at the same numerical result, which gives
the absolute velocity of light in terms of electrical quantities.

That theories apparently so fundamentally opposed should have so large


a field of truth common to both is a fact the philosophical importance
of which we cannot fully appreciate till we have reached a scientific
altitude from which the true relation between hypotheses so different
can be seen.

I shall only make one more remark on the relation between Mathematics
and Physics. In themselves, one is an operation of the mind, the
other is a dance of molecules. The molecules have laws of their own,
some of which we select as most intelligible to us and most amenable
to our calculation. We form a theory from these partial data, and we
ascribe any deviation of the actual phenomena from this theory to
disturbing causes. At the same time we confess that what we call
disturbing causes are simply those parts of the true circumstances
which we do not know or have neglected, and we endeavour in future to
take account of them. We thus acknowledge that the so-called
disturbance is a mere figment of the mind, not a fact of nature, and
that in natural action there is no disturbance.

But this is not the only way in which the harmony of the material with
the mental operation may be disturbed. The mind of the mathematician
is subject to many disturbing causes, such as fatigue, loss of memory,
and hasty conclusions; and it is found that, from these and other
causes, mathematicians make mistakes.

I am not prepared to deny that, to some mind of a higher order than


ours, each of these errors might be traced to the regular operation of
the laws of actual thinking; in fact we ourselves often do detect, not
only errors of calculation, but the causes of these errors. This,
however, by no means alters our conviction that they are errors, and
that one process of thought is right and another process wrong. I

One of the most profound mathematicians and thinkers of our time, the
23
late George Boole, when reflecting on the precise and almost
mathematical character of the laws of right thinking as compared with
the exceedingly perplexing though perhaps equally determinate laws of
actual and fallible thinking, was led to another of those points of
view from which Science seems to look out into a region beyond her own
domain.

"We must admit," he says, "that there exist laws" (of thought) "which
even the rigour of their mathematical forms does not preserve from
violation. We must ascribe to them an authority, the essence of which
does not consist in power, a supremacy which the analogy of the
inviolable order of the natural world in no way assists us to
comprehend."

Introductory Lecture on Experimental Physics.

James Clerk Maxwell

The University of Cambridge, in accordance with that law of its


evolution, by which, while maintaining the strictest continuity
between the successive phases of its history, it adapts itself with
more or less promptness to the requirements of the times, has lately
instituted a course of Experimental Physics. This course of study,
while it requires us to maintain in action all those powers of
attention and analysis which have been so long cultivated in the
University, calls on us to exercise our senses in observation, and our
hands in manipulation. The familiar apparatus of pen, ink, and paper
will no longer be sufficient for us, and we shall require more room
than that afforded by a seat at a desk, and a wider area than that of
the black board. We owe it to the munificence of our Chancellor,
that, whatever be the character in other respects of the experiments
which we hope hereafter to conduct, the material facilities for their
full development will be upon a scale which has not hitherto been
surpassed.

The main feature, therefore, of Experimental Physics at Cambridge is


24
the Devonshire Physical Laboratory, and I think it desirable that on
the present occasion, before we enter on the details of any special
study, we should consider by what means we, the University of
Cambridge, may, as a living body, appropriate and vitalise this new
organ, the outward shell of which we expect soon to rise before us.
The course of study at this University has always included Natural
Philosophy, as well as Pure Mathematics. To diffuse a sound knowledge
of Physics, and to imbue the minds of our students with correct
dynamical principles, have been long regarded as among our highest
functions, and very few of us can now place ourselves in the mental
condition in which even such philosophers as the great Descartes were
involved in the days before Newton had announced the true laws of the
motion of bodies. Indeed the cultivation and diffusion of sound
dynamical ideas has already effected a great change in the language
and thoughts even of those who make no pretensions to science, and we
are daily receiving fresh proofs that the popularisation of scientific
doctrines is producing as great an alteration in the mental state of
society as the material applications of science are effecting in its
outward life. Such indeed is the respect paid to science, that the
most absurd opinions may become current, provided they are expressed
in language, the sound of which recals some well-known scientific
phrase. If society is thus prepared to receive all kinds of
scientific doctrines, it is our part to provide for the diffusion and
cultivation, not only of true scientific principles, but of a spirit
of sound criticism, founded on an examination of the evidences on
which statements apparently scientific depend.

When we shall be able to employ in scientific education, not only the


trained attention of the student, and his familiarity with symbols,
but the keenness of his eye, the quickness of his ear, the delicacy of
his touch, and the adroitness of his fingers, we shall not only extend
our influence over a class of men who are not fond of cold
abstractions, but, by opening at once all the gateways of knowledge,
we shall ensure the association of the doctrines of science with those
elementary sensations which form the obscure background of all our
conscious thoughts, and which lend a vividness and relief to ideas,
which, when presented as mere abstract terms, are apt to fade entirely
from the memory.

In a course of Experimental Physics we may consider either the Physics


25
or the Experiments as the leading feature. We may either employ the
experiments to illustrate the phenomena of a particular branch of
Physics, or we may make some physical research in order to exemplify a
particular experimental method. In the order of time, we should
begin, in the Lecture Room, with a course of lectures on some branch
of Physics aided by experiments of illustration, and conclude, in the
Laboratory, with a course of experiments of research.

Let me say a few words on these two classes of


experiments,--Experiments of Illustration and Experiments of Research.
The aim of an experiment of illustration is to throw light upon some
scientific idea so that the student may be enabled to grasp it. The
circumstances of the experiment are so arranged that the phenomenon
which we wish to observe or to exhibit is brought into prominence,
instead of being obscured and entangled among other phenomena, as it
is when it occurs in the ordinary course of nature. To exhibit
illustrative experiments, to encourage others to make them, and to
cultivate in every way the ideas on which they throw light, forms an
important part of our duty. The simpler the materials of an
illustrative experiment, and the more familiar they are to the
student, the more thoroughly is he likely to acquire the idea which it
is meant to illustrate. The educational value of such experiments is
often inversely proportional to the complexity of the apparatus. The
student who uses home-made apparatus, which is always going wrong,
often learns more than one who has the use of carefully adjusted
instruments, to which he is apt to trust, and which he dares not take
to pieces.

It is very necessary that those who are trying to learn from books the
facts of physical science should be enabled by the help of a few
illustrative experiments to recognise these facts when they meet with
them out of doors. Science appears to us with a very different aspect
after we have found out that it is not in lecture rooms only, and by
means of the electric light projected on a screen, that we may witness
physical phenomena, but that we may find illustrations of the highest
doctrines of science in games and gymnastics, in travelling by land
and by water, in storms of the air and of the sea, and wherever there
is matter in motion.

This habit of recognising principles amid the endless variety of their


26
action can never degrade our sense of the sublimity of nature, or mar
our enjoyment of its beauty. On the contrary, it tends to rescue our
scientific ideas from that vague condition in which we too often leave
them, buried among the other products of a lazy credulity, and to
raise them into their proper position among the doctrines in which our
faith is so assured, that we are ready at all times to act on them.

Experiments of illustration may be of very different kinds. Some may


be adaptations of the commonest operations of ordinary life, others
may be carefully arranged exhibitions of some phenomenon which occurs
only under peculiar conditions. They all, however, agree in this,
that their aim is to present some phenomenon to the senses of the
student in such a way that he may associate with it the appropriate
scientific idea. When he has grasped this idea, the experiment which
illustrates it has served its purpose.

In an experiment of research, on the other hand, this is not the


principal aim. It is true that an experiment, in which the principal
aim is to see what happens under certain conditions, may be regarded
as an experiment of research by those who are not yet familiar with
the result, but in experimental researches, strictly so called, the
ultimate object is to measure something which we have already seen--to
obtain a numerical estimate of some magnitude.

Experiments of this class--those in which measurement of some kind is


involved, are the proper work of a Physical Laboratory. In every
experiment we have first to make our senses familiar with the
phenomenon, but we must not stop here, we must find out which of its
features are capable of measurement, and what measurements are
required in order to make a complete specification of the phenomenon.
We must then make these measurements, and deduce from them the result
which we require to find.

This characteristic of modern experiments--that they consist


principally of measurements,--is so prominent, that the opinion seems
to have got abroad, that in a few years all the great physical
constants will have been approximately estimated, and that the only
occupation which will then be left to men of science will be to carry
on these measurements to another place of decimals.

27
If this is really the state of things to which we are approaching, our
Laboratory may perhaps become celebrated as a place of conscientious
labour and consummate skill, but it will be out of place in the
University, and ought rather to be classed with the other great
workshops of our country, where equal ability is directed to more
useful ends.

But we have no right to think thus of the unsearchable riches of


creation, or of the untried fertility of those fresh minds into which
these riches will continue to be poured. It may possibly be true
that, in some of those fields of discovery which lie open to such
rough observations as can be made without artificial methods, the
great explorers of former times have appropriated most of what is
valuable, and that the gleanings which remain are sought after, rather
for their abstruseness, than for their intrinsic worth. But the
history of science shews that even during that phase of her progress
in which she devotes herself to improving the accuracy of the
numerical measurement of quantities with which she has long been
familiar, she is preparing the materials for the subjugation of new
regions, which would have remained unknown if she had been contented
with the rough methods of her early pioneers. I might bring forward
instances gathered from every branch of science, shewing how the
labour of careful measurement has been rewarded by the discovery of
new fields of research, and by the development of new scientific
ideas. But the history of the science of terrestrial magnetism
affords us a sufficient example of what may be done by Experiments in
Concert, such as we hope some day to perform in our Laboratory.

That celebrated traveller, Humboldt, was profoundly impressed with the


scientific value of a combined effort to be made by the observers of
all nations, to obtain accurate measurements of the magnetism of the
earth; and we owe it mainly to his enthusiasm for science, his great
reputation and his wide-spread influence, that not only private men of
science, but the governments of most of the civilised nations, our own
among the number, were induced to take part in the enterprise. But
the actual working out of the scheme, and the arrangements by which
the labours of the observers were so directed as to obtain the best
results, we owe to the great mathematician Gauss, working along with
Weber, the future founder of the science of electro-magnetic
measurement, in the magnetic observatory of Gottingen, and aided by
28
the skill of the instrument-maker Leyser. These men, however, did not
work alone. Numbers of scientific men joined the Magnetic Union,
learned the use of the new instruments and the new methods of reducing
the observations; and in every city of Europe you might see them, at
certain stated times, sitting, each in his cold wooden shed, with his
eye fixed at the telescope, his ear attentive to the clock, and his
pencil recording in his note-book the instantaneous position of the
suspended magnet.

Bacon's conception of "Experiments in concert" was thus realised, the


scattered forces of science were converted into a regular army, and
emulation and jealousy became out of place, for the results obtained
by any one observer were of no value till they were combined with
those of the others.

The increase in the accuracy and completeness of magnetic observations


which was obtained by the new method, opened up fields of research
which were hardly suspected to exist by those whose observations of
the magnetic needle had been conducted in a more primitive manner. We
must reserve for its proper place in our course any detailed
description of the disturbances to which the magnetism of our planet
is found to be subject. Some of these disturbances are periodic,
following the regular courses of the sun and moon. Others are sudden,
and are called magnetic storms, but, like the storms of the
atmosphere, they have their known seasons of frequency. The last and
the most mysterious of these magnetic changes is that secular
variation by which the whole character of the earth, as a great
magnet, is being slowly modified, while the magnetic poles creep on,
from century to century, along their winding track in the polar
regions.

We have thus learned that the interior of the earth is subject to the
influences of the heavenly bodies, but that besides this there is a
constantly progressive change going on, the cause of which is entirely
unknown. In each of the magnetic observatories throughout the world
an arrangement is at work, by means of which a suspended magnet
directs a ray of light on a preparred sheet of paper moved by
clockwork. On that paper the never-resting heart of the earth is now
tracing, in telegraphic symbols which will one day be interpreted, a
record of its pulsations and its flutterings, as well as of that slow
29
but mighty working which warns us that we must not suppose that the
inner history of our planet is ended.

But this great experimental research on Terrestrial Magnetism produced


lasting effects on the progress of science in general. I need only
mention one or two instances. The new methods of measuring forces
were successfully applied by Weber to the numerical determination of
all the phenomena of electricity, and very soon afterwards the
electric telegraph, by conferring a commercial value on exact
numerical measurements, contributed largely to the advancement, as
well as to the diffusion of scientific knowledge.

But it is not in these more modern branches of science alone that this
influence is felt. It is to Gauss, to the Magnetic Union, and to
magnetic observers in general, that we owe our deliverance from that
absurd method of estimating forces by a variable standard which
prevailed so long even among men of science. It was Gauss who first
based the practical measurement of magnetic force (and therefore of
every other force) on those long established principles, which, though
they are embodied in every dynamical equation, have been so generally
set aside, that these very equations, though correctly given in our
Cambridge textbooks, are usually explained there by assuming, in
addition to the variable standard of force, a variable, and therefore
illegal, standard of mass.

Such, then, were some of the scientific results which followed in this
case from bringing together mathematical power, experimental sagacity,
and manipulative skill, to direct and assist the labours of a body of
zealous observers. If therefore we desire, for our own advantage and
for the honour of our University, that the Devonshire Laboratory
should be successful, we must endeavour to maintain it in living union
with the other organs and faculties of our learned body. We shall
therefore first consider the relation in which we stand to those
mathematical studies which have so long flourished among us, which
deal with our own subjects, and which differ from our experimental
studies only in the mode in which they are presented to the mind.

There is no more powerful method for introducing knowledge into the


mind than that of presenting it in as many different ways as we can.
When the ideas, after entering through different gateways, effect a
30
junction in the citadel of the mind, the position they occupy becomes
impregnable. Opticians tell us that the mental combination of the
views of an object which we obtain from stations no further apart than
our two eyes is sufficient to produce in our minds an impression of
the solidity of the object seen; and we find that this impression is
produced even when we are aware that we are really looking at two flat
pictures placed in a stereoscope. It is therefore natural to expect
that the knowledge of physical science obtained by the combined use of
mathematical analysis and experimental research will be of a more
solid, available, and enduring kind than that possessed by the mere
mathematician or the mere experimenter.

But what will be the effect on the University, if men Pursuing that
course of reading which has produced so many distinguished Wranglers,
turn aside to work experiments? Will not their attendance at the
Laboratory count not merely as time withdrawn from their more
legitimate studies, but as the introduction of a disturbing element,
tainting their mathematical conceptions with material imagery, and
sapping their faith in the formulae of the textbook? Besides this, we
have already heard complaints of the undue extension of our studies,
and of the strain put upon our questionists by the weight of learning
which they try to carry with them into the Senate-House. If we now
ask them to get up their subjects not only by books and writing, but
at the same time by observation and manipulation, will they not break
down altogether? The Physical Laboratory, we are told, may perhaps be
useful to those who are going out in Natural Science, and who do
not take in Mathematics, but to attempt to combine both kinds of study
during the time of residence at the University is more than one mind
can bear.

No doubt there is some reason for this feeling. Many of us have


already overcome the initial difficulties of mathematical training.
When we now go on with our study, we feel that it requires exertion
and involves fatigue, but we are confident that if we only work hard
our progress will be certain.

Some of us, on the other hand, may have had some experience of the
routine of experimental work. As soon as we can read scales, observe
times, focus telescopes, and so on, this kind of work ceases to
require any great mental effort. We may perhaps tire our eyes and
31
weary our backs, but we do not greatly fatigue our minds.

It is not till we attempt to bring the theoretical part of our


training into contact with the practical that we begin to experience
the full effect of what Faraday has called "mental inertia"--not only
the difficulty of recognising, among the concrete objects before us,
the abstract relation which we have learned from books, but the
distracting pain of wrenching the mind away from the symbols to the
objects, and from the objects back to the symbols. This however is
the price we have to pay for new ideas.

But when we have overcome these difficulties, and successfully bridged


over the gulph between the abstract and the concrete, it is not a mere
piece of knowledge that we have obtained: we have acquired the
rudiment of a permanent mental endowment. When, by a repetition of
efforts of this kind, we have more fully developed the scientific
faculty, the exercise of this faculty in detecting scientific
principles in nature, and in directing practice by theory, is no
longer irksome, but becomes an unfailing source of enjoyment, to which
we return so often, that at last even our careless thoughts begin to
run in a scientific channel.

I quite admit that our mental energy is limited in quantity, and I


know that many zealous students try to do more than is good for them.
But the question about the introduction of experimental study is not
entirely one of quantity. It is to a great extent a question of
distribution of energy. Some distributions of energy, we know, are
more useful than others, because they are more available for those
purposes which we desire to accomplish.

Now in the case of study, a great part of our fatigue often arises,
not from those mental efforts by which we obtain the mastery of the
subject, but from those which are spent in recalling our wandering
thoughts; and these efforts of attention would be much less fatiguing
if the disturbing force of mental distraction could be removed.

This is the reason why a man whose soul is in his work always makes
more progress than one whose aim is something not immediately
connected with his occupation. In the latter case the very motive of
which he makes use to stimulate his flagging powers becomes the means
32
of distracting his mind from the work before him.

There may be some mathematicians who pursue their studies entirely for
their own sake. Most men, however, think that the chief use of
mathematics is found in the interpretation of nature. Now a man who
studies a piece of mathematics in order to understand some natural
phenomenon which he has seen, or to calculate the best arrangement of
some experiment which he means to make, is likely to meet with far
less distraction of mind than if his sole aim had been to sharpen his
mind for the successful practice of the Law, or to obtain a high place
in the Mathematical Tripos.

I have known men, who when they were at school, never could see the
good of mathematics, but who, when in after life they made this
discovery, not only became eminent as scientific engineers, but made
considerable progress in the study of abstract mathematics. If our
experimental course should help any of you to see the good of
mathematics, it will relieve us of much anxiety, for it will not only
ensure the success of your future studies, but it will make it much
less likely that they will prove injurious to your health.

But why should we labour to prove the advantage of practical science


to the University? Let us rather speak of the help which the
University may give to science, when men well trained in mathematics
and enjoying the advantages of a well-appointed Laboratory, shall
unite their efforts to carry out some experimental research which no
solitary worker could attempt.

At first it is probable that our principal experimental work must be


the illustration of particular branches of science, but as we go on we
must add to this the study of scientific methods, the same method
being sometimes illustrated by its application to researches belonging
to different branches of science.

We might even imagine a course of experimental study the arrangement


of which should be founded on a classification of methods, and not on
that of the objects of investigation. A combination of the two plans
seems to me better than either, and while we take every opportunity of
studying methods, we shall take care not to dissociate the method from
33
the scientific research to which it is applied, and to which it owes
its value.

We shall therefore arrange our lectures according to the


classification of the principal natural phenomena, such as heat,
electricity, magnetism and so on.

In the laboratory, on the other hand, the place of the different


instruments will be determined by a classification according to
methods, such as weighing and measuring, observations of time, optical
and electrical methods of observation, and so on.

The determination of the experiments to be performed at a particular


time must often depend upon the means we have at command, and in the
case of the more elaborate experiments, this may imply a long time of
preparation, during which the instruments, the methods, and the
observers themselves, are being gradually fitted for their work. When
we have thus brought together the requisites, both material and
intellectual, for a particular experiment, it may sometimes be
desirable that before the instruments are dismounted and the observers
dispersed, we should make some other experiment, requiring the same
method, but dealing perhaps with an entirely different class of
physical phenomena.

Our principal work, however, in the Laboratory must be to acquaint


ourselves with all kinds of scientific methods, to compare them, and
to estimate their value. It will, I think, be a result worthy of our
University, and more likely to be accomplished here than in any
private laboratory, if, by the free and full discussion of the
relative value of different scientific procedures, we succeed in
forming a school of scientific criticism, and in assisting the
development of the doctrine of method.

But admitting that a practical acquaintance with the methods of


Physical Science is an essential part of a mathematical and scientific
education, we may be asked whether we are not attributing too much
importance to science altogether as part of a liberal education.

Fortunately, there is no question here whether the University should


continue to be a place of liberal education, or should devote itself
34
to preparing young men for particular professions. Hence though some
of us may, I hope, see reason to make the pursuit of science the main
business of our lives, it must be one of our most constant aims to
maintain a living connexion between our work and the other liberal
studies of Cambridge, whether literary, philological, historical or
philosophical.

There is a narrow professional spirit which may grow up among men of


science, just as it does among men who practise any other special
business. But surely a University is the very place where we should
be able to overcome this tendency of men to become, as it were,
granulated into small worlds, which are all the more worldly for their
very smallness. We lose the advantage of having men of varied
pursuits collected into one body, if we do not endeavour to imbibe
some of the spirit even of those whose special branch of learning is
different from our own.

It is not so long ago since any man who devoted himself to geometry,
or to any science requiring continued application, was looked upon as
necessarily a misanthrope, who must have abandoned all human
interests, and betaken himself to abstractions so far removed from the
world of life and action that he has become insensible alike to the
attractions of pleasure and to the claims of duty.

In the present day, men of science are not looked upon with the same
awe or with the same suspicion. They are supposed to be in league
with the material spirit of the age, and to form a kind of advanced
Radical party among men of learning.

We are not here to defend literary and historical studies. We admit


that the proper study of mankind is man. But is the student of
science to be withdrawn from the study of man, or cut off from every
noble feeling, so long as he lives in intellectual fellowship with men
who have devoted their lives to the discovery of truth, and the
results of whose enquiries have impressed themselves on the ordinary
speech and way of thinking of men who never heard their names? Or is
the student of history and of man to omit from his consideration the
history of the origin and diffusion of those ideas which have produced
so great a difference between one age of the world and another?

35
It is true that the history of science is very different from the
science of history. We are not studying or attempting to study the
working of those blind forces which, we are told, are operating on
crowds of obscure people, shaking principalities and powers, and
compelling reasonable men to bring events to pass in an order laid
down by philosophers.

The men whose names are found in the history of science are not mere
hypothetical constituents of a crowd, to be reasoned upon only in
masses. We recognise them as men like ourselves, and their actions
and thoughts, being more free from the influence of passion, and
recorded more accurately than those of other men, are all the better
materials for the study of the calmer parts of human nature.

But the history of science is not restricted to the enumeration of


successful investigations. It has to tell of unsuccessful inquiries,
and to explain why some of the ablest men have failed to find the key
of knowledge, and how the reputation of others has only given a firmer
footing to the errors into which they fell.

The history of the development, whether normal or abnormal, of ideas


is of all subjects that in which we, as thinking men, take the deepest
interest. But when the action of the mind passes out of the
intellectual stage, in which truth and error are the alternatives,
into the more violently emotional states of anger and passion, malice
and envy, fury and madness; the student of science, though he is
obliged to recognise the powerful influence which these wild forces
have exercised on mankind, is perhaps in some measure disqualified
from pursuing the study of this part of human nature.

But then how few of us are capable of deriving profit from such
studies. We cannot enter into full sympathy with these lower phases
of our nature without losing some of that antipathy to them which is
our surest safeguard against a reversion to a meaner type, and we
gladly return to the company of those illustrious men who by aspiring
to noble ends, whether intellectual or practical, have risen above the
region of storms into a clearer atmosphere, where there is no
misrepresentation of opinion, nor ambiguity of expression, but where
one mind comes into closest contact with another at the point where
both approach nearest to the truth.
36
I propose to lecture during this term on Heat, and, as our facilities
for experimental work are not yet fully developed, I shall endeavour
to place before you the relative position and scientific connexion of
the different branches of the science, rather than to discuss the
details of experimental methods.

We shall begin with Thermometry, or the registration of temperatures,


and Calorimetry, or the measurement of quantities of heat. We shall
then go on to Thermodynamics, which investigates the relations between
the thermal properties of bodies and their other dynamical properties,
in so far as these relations may be traced without any assumption as
to the particular constitution of these bodies.

The principles of Thermodynamics throw great light on all the


phenomena of nature, and it is probable that many valuable
applications of these principles have yet to be made; but we shall
have to point out the limits of this science, and to shew that many
problems in nature, especially those in which the Dissipation of
Energy comes into play, are not capable of solution by the principles
of Thermodynamics alone, but that in order to understand them, we are
obliged to form some more definite theory of the constitution of
bodies.

Two theories of the constitution of bodies have struggled for victory


with various fortunes since the earliest ages of speculation: one is
the theory of a universal plenum, the other is that of atoms and void.

The theory of the plenum is associated with the doctrine of


mathematical continuity, and its mathematical methods are those of the
Differential Calculus, which is the appropriate expression of the
relations of continuous quantity.

The theory of atoms and void leads us to attach more importance to the
doctrines of integral numbers and definite proportions; but, in
applying dynamical principles to the motion of immense numbers of
atoms, the limitation of our faculties forces us to abandon the
attempt to express the exact history of each atom, and to be content
with estimating the average condition of a group of atoms large enough
37
to be visible. This method of dealing with groups of atoms, which I
may call the statistical method, and which in the present state of our
knowledge is the only available method of studying the properties of
real bodies, involves an abandonment of strict dynamical principles,
and an adoption of the mathematical methods belonging to the theory of
probability. It is probable that important results will be obtained
by the application of this method, which is as yet little known and is
not familiar to our minds. If the actual history of Science had been
different, and if the scientific doctrines most familiar to us had
been those which must be expressed in this way, it is possible that we
might have considered the existence of a certain kind of contingency a
self-evident truth, and treated the doctrine of philosophical
necessity as a mere sophism.

About the beginning of this century, the properties of bodies were


investigated by several distinguished French mathematicians on the
hypothesis that they are systems of molecules in equilibrium. The
somewhat unsatisfactory nature of the results of these investigations
produced, especially in this country, a reaction in favour of the
opposite method of treating bodies as if they were, so far at least as
our experiments are concerned, truly continuous. This method, in the
hands of Green, Stokes, and others, has led to results, the value of
which does not at all depend on what theory we adopt as to the
ultimate constitution of bodies.

One very important result of the investigation of the properties of


bodies on the hypothesis that they are truly continuous is that it
furnishes us with a test by which we can ascertain, by experiments on
a real body, to what degree of tenuity it must be reduced before it
begins to give evidence that its properties are no longer the same as
those of the body in mass. Investigations of this kind, combined with
a study of various phenomena of diffusion and of dissipation of
energy, have recently added greatly to the evidence in favour of the
hypothesis that bodies are systems of molecules in motion.

I hope to be able to lay before you in the course of the term some of
the evidence for the existence of molecules, considered as individual
bodies having definite properties. The molecule, as it is presented to
the scientific imagination, is a very different body from any of those
with which experience has hitherto made us acquainted.
38
In the first place its mass, and the other constants which define its
properties, are absolutely invariable; the individual molecule can
neither grow nor decay, but remains unchanged amid all the changes of
the bodies of which it may form a constituent.

In the second place it is not the only molecule of its kind, for there
are innumerable other molecules, whose constants are not
approximately, but absolutely identical with those of the first
molecule, and this whether they are found on the earth, in the sun, or
in the fixed stars.

By what process of evolution the philosophers of the future will


attempt to account for this identity in the properties of such a
multitude of bodies, each of them unchangeable in magnitude, and some
of them separated from others by distances which Astronomy attempts in
vain to measure, I cannot conjecture. My mind is limited in its power
of speculation, and I am forced to believe that these molecules must
have been made as they are from the beginning of their existence.

I also conclude that since none of the processes of nature, during


their varied action on different individual molecules, have produced,
in the course of ages, the slightest difference between the properties
of one molecule and those of another, the history of whose
combinations has been different, we cannot ascribe either their
existence or the identity of their properties to the operation of any
of those causes which we call natural.

Is it true then that our scientific speculations have really


penetrated beneath the visible appearance of things, which seem to be
subject to generation and corruption, and reached the entrance of that
world of order and perfection, which continues this day as it was
created, perfect in number and measure and weight?

We may be mistaken. No one has as yet seen or handled an individual


molecule, and our molecular hypothesis may, in its turn, be supplanted
by some new theory of the constitution of matter; but the idea of the
existence of unnumbered individual things, all alike and all
unchangeable, is one which cannot enter the human mind and remain
without fruit.
39
But what if these molecules, indestructible as they are, turn out to
be not substances themselves, but mere affections of some other
substance?

According to Sir W. Thomson's theory of Vortex Atoms, the substance of


which the molecule consists is a uniformly dense _plenum_, the
properties of which are those of a perfect fluid, the molecule itself
being nothing but a certain motion impressed on a portion of this
fluid, and this motion is shewn, by a theorem due to Helmholtz, to be
as indestructible as we believe a portion of matter to be.

If a theory of this kind is true, or even if it is conceivable, our


idea of matter may have been introduced into our minds through our
experience of those systems of vortices which we call bodies, but
which are not substances, but motions of a substance; and yet the idea
which we have thus acquired of matter, as a substance possessing
inertia, may be truly applicable to that fluid of which the vortices
are the motion, but of whose existence, apart from the vortical motion
of some of its parts, our experience gives us no evidence whatever.

It has been asserted that metaphysical speculation is a thing of the


past, and that physical science has extirpated it. The discussion of
the categories of existence, however, does not appear to be in danger
of coming to an end in our time, and the exercise of speculation
continues as fascinating to every fresh mind as it was in the days of
Thales.

End of Project Gutenberg's Five of Maxwell's Papers, by James Clerk Maxwell

*** END OF THIS PROJECT GUTENBERG EBOOK FIVE OF MAXWELL'S


PAPERS ***

40
***** This file should be named 4908.txt or 4908.zip *****
This and all associated files of various formats will be found in:
http://www.gutenberg.org/4/9/0/4908/

Produced by Gordon Keener

Updated editions will replace the previous one--the old editions


will be renamed.

Creating the works from public domain print editions means that no
one owns a United States copyright in these works, so the Foundation
(and you!) can copy and distribute it in the United States without
permission and without paying copyright royalties. Special rules,
set forth in the General Terms of Use part of this license, apply to
copying and distributing Project Gutenberg-tm electronic works to
protect the PROJECT GUTENBERG-tm concept and trademark. Project
Gutenberg is a registered trademark, and may not be used if you
charge for the eBooks, unless you receive specific permission. If you
do not charge anything for copies of this eBook, complying with the
rules is very easy. You may use this eBook for nearly any purpose
such as creation of derivative works, reports, performances and
research. They may be modified and printed and given away--you may do
practically ANYTHING with public domain eBooks. Redistribution is
subject to the trademark license, especially commercial
redistribution.

*** START: FULL LICENSE ***

THE FULL PROJECT GUTENBERG LICENSE


PLEASE READ THIS BEFORE YOU DISTRIBUTE OR USE THIS WORK

To protect the Project Gutenberg-tm mission of promoting the free


distribution of electronic works, by using or distributing this work
(or any other work associated in any way with the phrase "Project
Gutenberg"), you agree to comply with all the terms of the Full Project
Gutenberg-tm License (available with this file or online at
http://gutenberg.org/license).

41
Section 1. General Terms of Use and Redistributing Project Gutenberg-tm
electronic works

1.A. By reading or using any part of this Project Gutenberg-tm


electronic work, you indicate that you have read, understand, agree to
and accept all the terms of this license and intellectual property
(trademark/copyright) agreement. If you do not agree to abide by all
the terms of this agreement, you must cease using and return or destroy
all copies of Project Gutenberg-tm electronic works in your possession.
If you paid a fee for obtaining a copy of or access to a Project
Gutenberg-tm electronic work and you do not agree to be bound by the
terms of this agreement, you may obtain a refund from the person or
entity to whom you paid the fee as set forth in paragraph 1.E.8.

1.B. "Project Gutenberg" is a registered trademark. It may only be


used on or associated in any way with an electronic work by people who
agree to be bound by the terms of this agreement. There are a few
things that you can do with most Project Gutenberg-tm electronic works
even without complying with the full terms of this agreement. See
paragraph 1.C below. There are a lot of things you can do with Project
Gutenberg-tm electronic works if you follow the terms of this agreement
and help preserve free future access to Project Gutenberg-tm electronic
works. See paragraph 1.E below.

1.C. The Project Gutenberg Literary Archive Foundation ("the Foundation"


or PGLAF), owns a compilation copyright in the collection of Project
Gutenberg-tm electronic works. Nearly all the individual works in the
collection are in the public domain in the United States. If an
individual work is in the public domain in the United States and you are
located in the United States, we do not claim a right to prevent you from
copying, distributing, performing, displaying or creating derivative
works based on the work as long as all references to Project Gutenberg
are removed. Of course, we hope that you will support the Project
Gutenberg-tm mission of promoting free access to electronic works by
freely sharing Project Gutenberg-tm works in compliance with the terms of
this agreement for keeping the Project Gutenberg-tm name associated with
the work. You can easily comply with the terms of this agreement by
keeping this work in the same format with its attached full Project
Gutenberg-tm License when you share it without charge with others.
42
1.D. The copyright laws of the place where you are located also govern
what you can do with this work. Copyright laws in most countries are in
a constant state of change. If you are outside the United States, check
the laws of your country in addition to the terms of this agreement
before downloading, copying, displaying, performing, distributing or
creating derivative works based on this work or any other Project
Gutenberg-tm work. The Foundation makes no representations concerning
the copyright status of any work in any country outside the United
States.

1.E. Unless you have removed all references to Project Gutenberg:

1.E.1. The following sentence, with active links to, or other immediate
access to, the full Project Gutenberg-tm License must appear prominently
whenever any copy of a Project Gutenberg-tm work (any work on which the
phrase "Project Gutenberg" appears, or with which the phrase "Project
Gutenberg" is associated) is accessed, displayed, performed, viewed,
copied or distributed:

This eBook is for the use of anyone anywhere at no cost and with
almost no restrictions whatsoever. You may copy it, give it away or
re-use it under the terms of the Project Gutenberg License included
with this eBook or online at www.gutenberg.org/license

1.E.2. If an individual Project Gutenberg-tm electronic work is derived


from the public domain (does not contain a notice indicating that it is
posted with permission of the copyright holder), the work can be copied
and distributed to anyone in the United States without paying any fees
or charges. If you are redistributing or providing access to a work
with the phrase "Project Gutenberg" associated with or appearing on the
work, you must comply either with the requirements of paragraphs 1.E.1
through 1.E.7 or obtain permission for the use of the work and the
Project Gutenberg-tm trademark as set forth in paragraphs 1.E.8 or
1.E.9.

1.E.3. If an individual Project Gutenberg-tm electronic work is posted


with the permission of the copyright holder, your use and distribution
must comply with both paragraphs 1.E.1 through 1.E.7 and any additional
terms imposed by the copyright holder. Additional terms will be linked
43
to the Project Gutenberg-tm License for all works posted with the
permission of the copyright holder found at the beginning of this work.

1.E.4. Do not unlink or detach or remove the full Project Gutenberg-tm


License terms from this work, or any files containing a part of this
work or any other work associated with Project Gutenberg-tm.

1.E.5. Do not copy, display, perform, distribute or redistribute this


electronic work, or any part of this electronic work, without
prominently displaying the sentence set forth in paragraph 1.E.1 with
active links or immediate access to the full terms of the Project
Gutenberg-tm License.

1.E.6. You may convert to and distribute this work in any binary,
compressed, marked up, nonproprietary or proprietary form, including any
word processing or hypertext form. However, if you provide access to or
distribute copies of a Project Gutenberg-tm work in a format other than
"Plain Vanilla ASCII" or other format used in the official version
posted on the official Project Gutenberg-tm web site (www.gutenberg.org),
you must, at no additional cost, fee or expense to the user, provide a
copy, a means of exporting a copy, or a means of obtaining a copy upon
request, of the work in its original "Plain Vanilla ASCII" or other
form. Any alternate format must include the full Project Gutenberg-tm
License as specified in paragraph 1.E.1.

1.E.7. Do not charge a fee for access to, viewing, displaying,


performing, copying or distributing any Project Gutenberg-tm works
unless you comply with paragraph 1.E.8 or 1.E.9.

1.E.8. You may charge a reasonable fee for copies of or providing


access to or distributing Project Gutenberg-tm electronic works provided
that

- You pay a royalty fee of 20% of the gross profits you derive from
the use of Project Gutenberg-tm works calculated using the method
you already use to calculate your applicable taxes. The fee is
owed to the owner of the Project Gutenberg-tm trademark, but he
has agreed to donate royalties under this paragraph to the
Project Gutenberg Literary Archive Foundation. Royalty payments
must be paid within 60 days following each date on which you
44
prepare (or are legally required to prepare) your periodic tax
returns. Royalty payments should be clearly marked as such and
sent to the Project Gutenberg Literary Archive Foundation at the
address specified in Section 4, "Information about donations to
the Project Gutenberg Literary Archive Foundation."

- You provide a full refund of any money paid by a user who notifies
you in writing (or by e-mail) within 30 days of receipt that s/he
does not agree to the terms of the full Project Gutenberg-tm
License. You must require such a user to return or
destroy all copies of the works possessed in a physical medium
and discontinue all use of and all access to other copies of
Project Gutenberg-tm works.

- You provide, in accordance with paragraph 1.F.3, a full refund of any


money paid for a work or a replacement copy, if a defect in the
electronic work is discovered and reported to you within 90 days
of receipt of the work.

- You comply with all other terms of this agreement for free
distribution of Project Gutenberg-tm works.

1.E.9. If you wish to charge a fee or distribute a Project Gutenberg-tm


electronic work or group of works on different terms than are set
forth in this agreement, you must obtain permission in writing from
both the Project Gutenberg Literary Archive Foundation and Michael
Hart, the owner of the Project Gutenberg-tm trademark. Contact the
Foundation as set forth in Section 3 below.

1.F.

1.F.1. Project Gutenberg volunteers and employees expend considerable


effort to identify, do copyright research on, transcribe and proofread
public domain works in creating the Project Gutenberg-tm
collection. Despite these efforts, Project Gutenberg-tm electronic
works, and the medium on which they may be stored, may contain
"Defects," such as, but not limited to, incomplete, inaccurate or
corrupt data, transcription errors, a copyright or other intellectual
property infringement, a defective or damaged disk or other medium, a
computer virus, or computer codes that damage or cannot be read by
45
your equipment.

1.F.2. LIMITED WARRANTY, DISCLAIMER OF DAMAGES - Except for the


"Right
of Replacement or Refund" described in paragraph 1.F.3, the Project
Gutenberg Literary Archive Foundation, the owner of the Project
Gutenberg-tm trademark, and any other party distributing a Project
Gutenberg-tm electronic work under this agreement, disclaim all
liability to you for damages, costs and expenses, including legal
fees. YOU AGREE THAT YOU HAVE NO REMEDIES FOR NEGLIGENCE,
STRICT
LIABILITY, BREACH OF WARRANTY OR BREACH OF CONTRACT
EXCEPT THOSE
PROVIDED IN PARAGRAPH 1.F.3. YOU AGREE THAT THE
FOUNDATION, THE
TRADEMARK OWNER, AND ANY DISTRIBUTOR UNDER THIS
AGREEMENT WILL NOT BE
LIABLE TO YOU FOR ACTUAL, DIRECT, INDIRECT, CONSEQUENTIAL,
PUNITIVE OR
INCIDENTAL DAMAGES EVEN IF YOU GIVE NOTICE OF THE
POSSIBILITY OF SUCH
DAMAGE.

1.F.3. LIMITED RIGHT OF REPLACEMENT OR REFUND - If you discover a


defect in this electronic work within 90 days of receiving it, you can
receive a refund of the money (if any) you paid for it by sending a
written explanation to the person you received the work from. If you
received the work on a physical medium, you must return the medium with
your written explanation. The person or entity that provided you with
the defective work may elect to provide a replacement copy in lieu of a
refund. If you received the work electronically, the person or entity
providing it to you may choose to give you a second opportunity to
receive the work electronically in lieu of a refund. If the second copy
is also defective, you may demand a refund in writing without further
opportunities to fix the problem.

1.F.4. Except for the limited right of replacement or refund set forth
in paragraph 1.F.3, this work is provided to you 'AS-IS' WITH NO OTHER
WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED, INCLUDING BUT
NOT LIMITED TO
46
WARRANTIES OF MERCHANTABILITY OR FITNESS FOR ANY PURPOSE.

1.F.5. Some states do not allow disclaimers of certain implied


warranties or the exclusion or limitation of certain types of damages.
If any disclaimer or limitation set forth in this agreement violates the
law of the state applicable to this agreement, the agreement shall be
interpreted to make the maximum disclaimer or limitation permitted by
the applicable state law. The invalidity or unenforceability of any
provision of this agreement shall not void the remaining provisions.

1.F.6. INDEMNITY - You agree to indemnify and hold the Foundation, the
trademark owner, any agent or employee of the Foundation, anyone
providing copies of Project Gutenberg-tm electronic works in accordance
with this agreement, and any volunteers associated with the production,
promotion and distribution of Project Gutenberg-tm electronic works,
harmless from all liability, costs and expenses, including legal fees,
that arise directly or indirectly from any of the following which you do
or cause to occur: (a) distribution of this or any Project Gutenberg-tm
work, (b) alteration, modification, or additions or deletions to any
Project Gutenberg-tm work, and (c) any Defect you cause.

Section 2. Information about the Mission of Project Gutenberg-tm

Project Gutenberg-tm is synonymous with the free distribution of


electronic works in formats readable by the widest variety of computers
including obsolete, old, middle-aged and new computers. It exists
because of the efforts of hundreds of volunteers and donations from
people in all walks of life.

Volunteers and financial support to provide volunteers with the


assistance they need, are critical to reaching Project Gutenberg-tm's
goals and ensuring that the Project Gutenberg-tm collection will
remain freely available for generations to come. In 2001, the Project
Gutenberg Literary Archive Foundation was created to provide a secure
and permanent future for Project Gutenberg-tm and future generations.
To learn more about the Project Gutenberg Literary Archive Foundation
and how your efforts and donations can help, see Sections 3 and 4
and the Foundation web page at http://www.pglaf.org.

47
Section 3. Information about the Project Gutenberg Literary Archive
Foundation

The Project Gutenberg Literary Archive Foundation is a non profit


501(c)(3) educational corporation organized under the laws of the
state of Mississippi and granted tax exempt status by the Internal
Revenue Service. The Foundation's EIN or federal tax identification
number is 64-6221541. Its 501(c)(3) letter is posted at
http://pglaf.org/fundraising. Contributions to the Project Gutenberg
Literary Archive Foundation are tax deductible to the full extent
permitted by U.S. federal laws and your state's laws.

The Foundation's principal office is located at 4557 Melan Dr. S.


Fairbanks, AK, 99712., but its volunteers and employees are scattered
throughout numerous locations. Its business office is located at
809 North 1500 West, Salt Lake City, UT 84116, (801) 596-1887, email
business@pglaf.org. Email contact links and up to date contact
information can be found at the Foundation's web site and official
page at http://pglaf.org

For additional contact information:


Dr. Gregory B. Newby
Chief Executive and Director
gbnewby@pglaf.org

Section 4. Information about Donations to the Project Gutenberg


Literary Archive Foundation

Project Gutenberg-tm depends upon and cannot survive without wide


spread public support and donations to carry out its mission of
increasing the number of public domain and licensed works that can be
freely distributed in machine readable form accessible by the widest
array of equipment including outdated equipment. Many small donations
($1 to $5,000) are particularly important to maintaining tax exempt
status with the IRS.

The Foundation is committed to complying with the laws regulating


charities and charitable donations in all 50 states of the United
48
States. Compliance requirements are not uniform and it takes a
considerable effort, much paperwork and many fees to meet and keep up
with these requirements. We do not solicit donations in locations
where we have not received written confirmation of compliance. To
SEND DONATIONS or determine the status of compliance for any
particular state visit http://pglaf.org

While we cannot and do not solicit contributions from states where we


have not met the solicitation requirements, we know of no prohibition
against accepting unsolicited donations from donors in such states who
approach us with offers to donate.

International donations are gratefully accepted, but we cannot make


any statements concerning tax treatment of donations received from
outside the United States. U.S. laws alone swamp our small staff.

Please check the Project Gutenberg Web pages for current donation
methods and addresses. Donations are accepted in a number of other
ways including checks, online payments and credit card donations.
To donate, please visit: http://pglaf.org/donate

Section 5. General Information About Project Gutenberg-tm electronic


works.

Professor Michael S. Hart is the originator of the Project Gutenberg-tm


concept of a library of electronic works that could be freely shared
with anyone. For thirty years, he produced and distributed Project
Gutenberg-tm eBooks with only a loose network of volunteer support.

Project Gutenberg-tm eBooks are often created from several printed


editions, all of which are confirmed as Public Domain in the U.S.
unless a copyright notice is included. Thus, we do not necessarily
keep eBooks in compliance with any particular paper edition.

Most people start at our Web site which has the main PG search facility:

http://www.gutenberg.org
49
This Web site includes information about Project Gutenberg-tm,
including how to make donations to the Project Gutenberg Literary
Archive Foundation, how to help produce our new eBooks, and how to
subscribe to our email newsletter to hear about new eBooks.

% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %

% %

% The Project Gutenberg EBook of On the Quantum Theory of Line-Spectra, Part


1 and 2, by Niels Henrik David Bohr

% %

% This eBook is for the use of anyone anywhere in the United States and most

% other parts of the world at no cost and with almost no restrictions %

% whatsoever. You may copy it, give it away or re-use it under the terms of

% the Project Gutenberg License included with this eBook or online at %

% www.gutenberg.org. If you are not located in the United States, you'll have

% to check the laws of the country where you are located before using this ebook.

% %

% %

% %

% Title: On the Quantum Theory of Line-Spectra, Part 1 and 2 %

% %

% Author: Niels Henrik David Bohr %

% %

% Release Date: November 22, 2014 [EBook #47167] %


50
% %

% Language: EN %

% %

% Character set encoding: ISO-8859-1 %

% %

% *** START OF THIS PROJECT GUTENBERG EBOOK QUANTUM


THEORY OF LINE-SPECTRA ***

% %

% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %

\def\ebook{47167}

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

%% %%

%% Packages and substitutions: %%

%% %%

%% book: Required. %%

%% inputenc: Latin-1 text encoding. Required. %%

%% %%

%% ifthen: Logical conditionals. Required. %%

%% %%

%% amsmath: AMS mathematics enhancements. Required. %%

%% amssymb: Additional mathematical symbols. Required. %%

51
%% %%

%% alltt: Fixed-width font environment. Required. %%

%% %%

%% indentfirst: Indent first paragraph of each section. Optional. %%

%% %%

%% footmisc: Start footnote numbering on each page. Required. %%

%% %%

%% calc: Length calculations. Required. %%

%% %%

%% fancyhdr: Enhanced running headers and footers. Required. %%

%% %%

%% geometry: Enhanced page layout package. Required. %%

%% hyperref: Hypertext embellishments for pdf output. Required. %%

%% %%

%% %%

%% Producer's Comments: %%

%% %%

%% OCR text for this ebook was obtained on Sept. 10, 2014, from %%

%% http://www.archive.org/details/onquantumtheoryo01bohruoft. %%

%% %%

%% Footnote marks have been moved to after punctuation for %%

%% consistency. (Individual instances not nted.) %%

%% %%
52
%% Minor changes to the original are noted in this file in three %%

%% ways: %%

%% 1. \Typo{}{} for typographical corrections, showing original %%

%% and replacement text side-by-side. %%

%% 2. \Chg{}{} and \Add{}, for inconsistent/missing punctuation,%%

%% italicization, and capitalization. %%

%% 3. [** TN: Note]s for lengthier or stylistic comments. %%

%% %%

%% %%

%% Compilation Flags: %%

%% %%

%% The following behavior may be controlled by boolean flags. %%

%% %%

%% ForPrinting (false by default): %%

%% If false, compile a screen optimized file (one-sided layout, %%

%% blue hyperlinks). If true, print-optimized PDF file: Larger %%

%% text block, two-sided layout, black hyperlinks. %%

%% %%

%% %%

%% PDF pages: 212 (if ForPrinting set to false) %%

%% PDF page size: 4.5 x 6.5" (non-standard) %%

%% %%

%% Summary of log file: %%


53
%% * Two underfull hboxes (no visual problem) %%

%% * One overfull vbox (< 0.321 pt) %%

%% %%

%% Compile History: %%

%% %%

%% November, 2014: (Andrew D. Hwang) %%

%% texlive2012, GNU/Linux %%

%% %%

%% Command block: %%

%% %%

%% pdflatex x3 %%

%% %%

%% %%

%% November 2014: pglatex. %%

%% Compile this project with: %%

%% pdflatex 47167-t.tex ..... THREE times %%

%% %%

%% pdfTeX, Version 3.1415926-2.5-1.40.14 (TeX Live 2013/Debian) %%

%% %%

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

\listfiles

\documentclass[12pt]{book}[2005/09/16]

54
%%%%%%%%%%%%%%%%%%%%%%%%%%%%% PACKAGES %%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%

\usepackage[latin1]{inputenc}[2006/05/05]

\usepackage{ifthen}[2001/05/26] %% Logical conditionals

\usepackage{amsmath}[2000/07/18] %% Displayed equations

\usepackage{amssymb}[2002/01/22] %% and additional symbols

\usepackage{alltt}[1997/06/16] %% boilerplate, credits, license

\IfFileExists{indentfirst.sty}{%

\usepackage{indentfirst}[1995/11/23]

}{}

\usepackage[perpage]{footmisc}[2005/03/17]

\usepackage{calc}[2005/08/06]

\usepackage{fancyhdr} %% For running heads

55
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%

%%%% Interlude: Set up PRINTING (default) or SCREEN VIEWING %%%%

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%

% ForPrinting=true false (default)

% Asymmetric margins Symmetric margins

% 1 : 1.5 text block aspect ratio 3 : 4 text block aspect ratio

% Black hyperlinks Blue hyperlinks

% Start major marker pages recto No blank verso pages

\newboolean{ForPrinting}

%% UNCOMMENT the next line for a PRINT-OPTIMIZED VERSION of the


text %%

%\setboolean{ForPrinting}{true}

%% Initialize values to ForPrinting=false

\newcommand{\Margins}{hmarginratio=1:1} % Symmetric margins

\newcommand{\HLinkColor}{blue} % Hyperlink color

\newcommand{\PDFPageLayout}{SinglePage}

\newcommand{\TransNote}{Transcriber's Note}

\newcommand{\TransNoteCommon}{%

56
The camera-quality files for this public-domain ebook may be

downloaded \textit{gratis} at

\begin{center}

\texttt{www.gutenberg.org/ebooks/\ebook}.

\end{center}

This ebook, comprising Parts I~and~II only, was produced using scanned

images and OCR text generously provided by the University of Toronto

Gerstein Library through the Internet Archive.

\bigskip

Minor typographical corrections and presentational changes have been

made without comment.

\bigskip

\newcommand{\TransNoteText}{%

\TransNoteCommon

This PDF file is optimized for screen viewing, but may be recompiled

for printing. Please consult the preamble of the \LaTeX\ source file

for instructions and other particulars.

}
57
%% Re-set if ForPrinting=true

\ifthenelse{\boolean{ForPrinting}}{%

\renewcommand{\Margins}{hmarginratio=2:3} % Asymmetric margins

\renewcommand{\HLinkColor}{black} % Hyperlink color

\renewcommand{\PDFPageLayout}{TwoPageRight}

\renewcommand{\TransNote}{Transcriber's Note}

\renewcommand{\TransNoteText}{%

\TransNoteCommon

This PDF file is optimized for printing, but may be recompiled for

screen viewing. Please consult the preamble of the \LaTeX\ source

file for instructions and other particulars.

}{% If ForPrinting=false, don't skip to recto

\renewcommand{\cleardoublepage}{\clearpage}

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%

%%%% End of PRINTING/SCREEN VIEWING code; back to packages %%%


%

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%

\ifthenelse{\boolean{ForPrinting}}{%

58
\setlength{\paperwidth}{8.5in}%

\setlength{\paperheight}{11in}%

% 1:1.6

\usepackage[body={5.75in,8.5in},\Margins]{geometry}[2002/07/08]

}{%

\setlength{\paperwidth}{4.5in}%

\setlength{\paperheight}{6.5in}%

\raggedbottom

% 3:4

\usepackage[body={4.25in,5.6in},\Margins,includeheadfoot]{geometry}
[2002/07/08]

\providecommand{\ebook}{00000} % Overridden during white-washing

\usepackage[pdftex,

hyperfootnotes=false,

pdfauthor={Niels Henrik David Bohr},

pdftitle={The Project Gutenberg eBook \#\ebook: On the Quantum Theory of


Line-Spectra, Part II.},

pdfkeywords={University of Toronto, The Internet Archive, Andrew D. Hwang},

pdfstartview=Fit, % default value

pdfstartpage=1, % default value

pdfpagemode=UseNone, % default value

59
bookmarks=true, % default value

linktocpage=false, % default value

pdfpagelayout=\PDFPageLayout,

pdfdisplaydoctitle,

pdfpagelabels=true,

bookmarksopen=true,

bookmarksopenlevel=0,

colorlinks=true,

linkcolor=\HLinkColor]{hyperref}[2007/02/07]

%% Fixed-width environment to format PG boilerplate %%

\newenvironment{PGtext}{%

\begin{alltt}

\fontsize{8.1}{10}\ttfamily\selectfont}%

{\end{alltt}}

% Errors found during digitization

\newcommand{\Typo}[2]{#2}

% Stylistic changes made for consistency

\newcommand{\Chg}[2]{#2}

%\newcommand{\Chg}[2]{#1} % Use this to revert inconsistencies in the original


60
\newcommand{\Add}[1]{\Chg{}{#1}}

% [** TN: Add comma before final item in a list]

\newcommand{\MyDots}{\dots\Add{,}\ }

\newcommand{\eps}{\varepsilon}

\renewcommand{\theta}{\vartheta}

\renewcommand{\phi}{\varphi}

\newcommand{\Note}[1]{}

%% Miscellaneous global parameters %%

% No hrule in page header

\renewcommand{\headrulewidth}{0pt}

% Footnote markers followed by a parenthesis

\makeatletter

\renewcommand\@makefnmark%

{\mbox{\@textsuperscript{\normalfont\@thefnmark})}}

\renewcommand\@makefntext[1]%

{\noindent\makebox[2.4em][r]{\@makefnmark\,}#1}

\makeatother

61
% Loosen spacing

\setlength{\emergencystretch}{1em}

%% Running heads %%

\newcommand{\FlushRunningHeads}{\clearpage\fancyhf{}}

\newcommand{\InitRunningHeads}{%

\setlength{\headheight}{15pt}

\pagestyle{fancy}

\thispagestyle{empty}

\fancyhead[C]{\thepage}

% Uniform style for running heads

\newcommand{\RHeads}[1]{\textsc{#1}}

\newcommand{\SetRunningHeads}[1]{%

\ifthenelse{\boolean{ForPrinting}}{%

\fancyhead[LE,RO]{\RHeads{#1}}%

}{

\fancyhead[L]{\RHeads{#1}}%

62
\newcommand{\BookMark}[2]{\phantomsection\pdfbookmark[#1]{#2}{#2}}

%% Major document divisions %%

\newcommand{\PGBoilerPlate}{%

\pagenumbering{Alph}

\pagestyle{empty}

\BookMark{0}{PG Boilerplate.}

\newcommand{\FrontMatter}{%

\cleardoublepage

\frontmatter

\BookMark{-1}{Front Matter.}

\newcommand{\MainMatter}{%

\FlushRunningHeads

\InitRunningHeads

\mainmatter

\BookMark{-1}{Main Matter.}

\newcommand{\BackMatter}{%

\FlushRunningHeads

\InitRunningHeads

\backmatter
63
\BookMark{-1}{Back Matter.}

\newcommand{\PGLicense}{%

\FlushRunningHeads

\pagenumbering{Roman}

\InitRunningHeads

\BookMark{-1}{PG License.}

\SetRunningHeads{License.}

% For internal use by equation tags and PDF bookmarks

\newcounter{PartNo}

\newcounter{SecNo}

%% Sectional units %%

\newcommand{\Part}[2]{%

\FlushRunningHeads

\InitRunningHeads

\BookMark{0}{Part #1}

\stepcounter{PartNo}

\setcounter{SecNo}{0}

\section*{\centering\large\normalfont\textsc{Part #1}}

\subsection*{\centering\large #2}
64
}

\newcommand{\Section}[2]{

\subsubsection*{\centering\normalsize§\;#1 #2}

\refstepcounter{SecNo}

\label{section:\Roman{PartNo}.\theSecNo}

\BookMark{1}{Section \Roman{PartNo}-#1}

\newcommand{\Introduction}{

\subsubsection*{\centering\normalsize Introduction.}

\BookMark{0}{Introduction.}

% Optional first argument is the Part (if not "II")

\newcommand{\SecRef}[2][]{%

\ifthenelse{\equal{#1}{}}{%

\hyperref[section:\Roman{PartNo}.#2]{{\upshape §\;#2}}%

}{%

\hyperref[section:#1.#2]{{\upshape §\;#2}}%

}%

65
\newenvironment{Hypothesis}[1]{\par#1\ \itshape}{\par\upshape}

\newenvironment{Remark}{\medskip\par\small}{\par\normalsize}

% Page separators

\newcommand{\PageSep}[1]{\ignorespaces}

\newcommand{\Pagelabel}[1]{\phantomsection\label{page:#1}}

% [** TN: Use this macro to duplicate the original uses of "p." versus "page"]

%\newcommand{\PageRef}[2][p.]{\hyperref[page:#2]{#1~\pageref*{page:#2}}}

% [** TN: Convert "p." to "page"]

\newcommand{\PageRef}[2][p.]{\hyperref[page:#2]{page~\pageref*{page:#2}}}

\newcommand{\PageRefI}[2][p.]{\PageRef[#1]{#2}}

% [** TN: If Part II is compiled separately from Part I, use the

% following to flag original (incorrect) page numbers with a \dag]

%\newcommand{\PageRefI}[2][p.]{\ifthenelse{\equal{#1}{}}{#2}{#1~#2}${}^{\
dag}$}

% Miscellaneous textual conveniences

\newcommand{\First}[1]{\textsc{#1}}

\newcommand{\Name}[1]{\textsc{#1}}

\newcommand{\Emph}[1]{\emph{#1}} % [** TN: Gesperrt in the original]

\newcommand{\Ord}[2]{#1\textsuperscript{\textit{#2}}}

66
\newcommand{\eg}{e.\;g.}

\newcommand{\ie}{i.\;e.}

% Sometimes need to break long lines for screen formatting

\newcommand{\BreakMath}[2]{\ifthenelse{\boolean{ForPrinting}}{#1}{#2}}

\newcommand{\MathBack}{\hspace*{-2em}}

%% Miscellaneous mathematical formatting %%

\DeclareMathOperator{\Div}{div}

\DeclareMathOperator{\curl}{curl}

\renewcommand{\div}{\Div}

\newcommand{\dd}{\partial}

\newcommand{\Frak}[1]{\mathfrak{#1}}

\newcommand{\frakA}{\Frak{A}}

\newcommand{\frakB}{\Frak{B}}

\newcommand{\frakC}{\Frak{C}}

\newcommand{\frakE}{\Frak{E}}

\newcommand{\frakH}{\Frak{H}}

\newcommand{\frakI}{\Frak{I}}

\newcommand{\frakn}{\Frak{n}}

\newcommand{\fraka}{\Frak{a}}

\newcommand{\frakb}{\Frak{b}}

\newcommand{\frakc}{\Frak{c}}
67
\newcommand{\fraks}{\Frak{s}}

\newcommand{\frakt}{\Frak{t}}

\newcommand{\frakv}{\Frak{v}}

\newcommand{\frakw}{\Frak{w}}

% Cross-ref-able equation tags

% Optional first argument of each is the Part (if necessary for disambiguation)

\newcommand{\Tag}[2][]{\phantomsection\label{eqn:#1#2}\tag*{\
ensuremath{#2}}}

\newcommand{\Eq}[2][]{\hyperref[eqn:#1#2]{\ensuremath{#2}}}

\newcommand{\EqI}[1]{\Eq[]{#1}}

%%%%%%%%%%%%%%%%%%%%%%%% START OF DOCUMENT %%


%%%%%%%%%%%%%%%%%%%%%%%%

\begin{document}

%% PG BOILERPLATE %%

\PGBoilerPlate

\begin{center}

\begin{minipage}{\textwidth}

\small

\begin{PGtext}

The Project Gutenberg EBook of On the Quantum Theory of Line-Spectra, Part 1


and 2, by Niels Henrik David Bohr

68
This eBook is for the use of anyone anywhere in the United States and most

other parts of the world at no cost and with almost no restrictions

whatsoever. You may copy it, give it away or re-use it under the terms of

the Project Gutenberg License included with this eBook or online at

www.gutenberg.org. If you are not located in the United States, you'll have

to check the laws of the country where you are located before using this ebook.

Title: On the Quantum Theory of Line-Spectra, Part 1 and 2

Author: Niels Henrik David Bohr

Release Date: November 22, 2014 [EBook #47167]

Language: EN

Character set encoding: ISO-8859-1

*** START OF THIS PROJECT GUTENBERG EBOOK QUANTUM THEORY


OF LINE-SPECTRA ***

\end{PGtext}

\end{minipage}

69
\end{center}

\newpage

%% Credits and transcriber's note %%

\begin{center}

\begin{minipage}{\textwidth}

\begin{PGtext}

Produced by Andrew D. Hwang

\end{PGtext}

\end{minipage}

\vfill

\end{center}

\begin{minipage}{0.85\textwidth}

\small

\BookMark{0}{Transcriber's Note.}

\subsection*{\centering\normalfont\scshape%

\normalsize\MakeLowercase{\TransNote}}%

\raggedright

\TransNoteText

\end{minipage}

%%%%%%%%%%%%%%%%%%%%%%%%%%% FRONT MATTER %%%


%%%%%%%%%%%%%%%%%%%%%%%

70
\PageSep{i}

\FrontMatter

\iffalse % [** TN: Publisher's titlepage with series header and price]

\begin{center}

Mémoires de l'Académie Royale des Sciences et des Lettres de Danemark.


Copenhague.

Section des Sciences, 3\ieme~série, t.~IV, \numero~1, fasc.~1.

\Large ON \\

\LARGE THE QUANTUM THEORY \\

OF LINE-SPECTRA \\

\footnotesize

BY \\

\large

N. BOHR \\

\vfill

\textsf{PART I} \\

\vfill

\footnotesize

71
\textsc{D. Kgl.\ Danske Vidensk.\ Selsk.\ Skripter, naturvidensk.\ og mathem.\
Afd.,~8. Række,~IV.}

\vfill\vfill

\large

KØBENHAVN \\

\textsc{hovedkommissionær}: ANDR. FRED. HØST \&~SON, \textsc{kgl., hof-


boghandel} \\

BIANCO LUNOS BOGTRYKKERI \\

1918

Pris: 2~Kr.\ 25~Ore.

\end{center}

\cleardoublepage

\fi % [** TN: End of publisher's titlepage]

\PageSep{ii}

%[Blank page]

\PageSep{iii}

\begin{center}

\large ON

\vfill

\Huge THE QUANTUM THEORY \\

OF LINE-SPECTRA

72
\vfill

\footnotesize

BY

\vfill

\large

N. BOHR \\

\vfill\vfill\vfill

\scriptsize

\textsc{D. Kgl.\ Danske Vidensk.\ Selsk.\ Skripter, naturvidensk.\ og mathem.\


Afd.,~8. Række,~IV.}

\vfill\vfill\vfill

\normalsize

KØBENHAVN \\

\scriptsize

\textsc{hovedkommissionær}: ANDR. FRED. HØST \&~SON, \textsc{kgl., hof-


boghandel} \\

\textsc{bianco lunos bogtrykkeri} \\

1918

\end{center}

\newpage

73
\PageSep{iv}

\null

\vfill

\begin{center}

\itshape

\footnotesize

DEDICATED TO THE MEMORY OF~MY~VENERATED~TEACHER \\[2\


baselineskip]

%[** TN: Small caps in the original]

\normalsize

Professor C. CHRISTIANSEN \\[\baselineskip]

\footnotesize

OCTOBER~9, 1843\qquad \dag\ NOVEMBER~28, 1917

\upshape\vfill

UDGIVET PAA CARLSBERGFONDETS BEKOSTNING

\end{center}

\PageSep{v}

\MainMatter

\Introduction

74
\First{In} an attempt to develop certain outlines of a theory of line-spectra based
on

a suitable application of the fundamental ideas introduced by Planck in his theory

of temperature-radiation to the theory of the nucleus atom of Sir \Name{Ernest


Rutherford},

the writer has shown that it is possible in this way to obtain a simple interpretation

of some of the main laws governing the line-spectra of the elements, and

especially to obtain a deduction of the well known Balmer formula for the
hydrogen

spectrum\footnote

{\Name{N. Bohr}, Phil.\ Mag., XXVI, pp.\ 1, 476, 857~(1913), XXVII, p.~506
(1914), XXIX. p.~332 (1915), XXX.

p.~394 (1915).}

The theory in the form given allowed of a detailed discussion

only in the case of periodic systems, and obviously was not able to account in

detail for the characteristic difference between the hydrogen spectrum and the

spectra of other elements, or for the characteristic effects on the hydrogen spectrum

of external electric and magnetic fields. Recently, however, a way out of this
difficulty

has been opened by \Name{Sommerfeld}\footnote

{\Name{A.~Sommerfeld}, Ber.\ Akad.\ München, 1915, pp.~425, 459, 1916,


p.~131. 1917. p.~83. Ann.\ de Phys.,

LI. p.~1 (1916).}

who, by introducing a suitable generalisation

of the theory to a simple type of non-periodic motions and by taking the

75
small variation of the mass of the electron with its velocity into account, obtained

an explanation of the fine-structure of the hydrogen lines which was found to be

in brilliant conformity with the measurements. Already in his first paper on this

subject, \Name{Sommerfeld} pointed out that his theory evidently offered a clue to
the

interpretation of the more intricate structure of the spectra of other elements.


Briefly

afterwards \Name{Epstein}\footnote

{\Name{P.~Epstein}, Phys.\ Zeitschr.\ XVII, p.~148 (1916), Ann.\ d.\ Phys.~L,


p.~489. LI. p.~168 (1916).}

and \Name{Schwarzschild},\footnote

{\Name{K.~\Typo{Schwarzchild}{Schwarzschild}}, Ber.\ Akad.\ Berlin, 1916,


p.~548.}

independent of each other, by adapting

\Name{Sommerfeld}'s ideas to the treatment of a more extended class of non-


periodic

systems obtained a detailed explanation of the characteristic effect of an electric

field on the hydrogen spectrum discovered by \Name{Stark}. Subsequently \


Name{Sommerfeld}\footnote

{\Name{A.~Sommerfeld}, Phys.\ Zeitschr.\ XVII, p.~491 (1916).}

himself and \Name{Debye}\footnote

{\Name{P.~Debye}, Nachr.\ K. Ges.\ d.~Wiss.\ Göttingen, 1916, Phys.\


Zeitschr.\ XVII, p.~507 (1916).}

have on the same lines indicated an interpretation of the

\PageSep{4}

76
effect of a magnetic field on the hydrogen spectrum which, although no complete

explanation of the observations was obtained, undoubtedly represents an important

step towards a detailed understanding of this phenomenon.

In spite of the great progress involved in these investigations many difficulties

of fundamental nature remained unsolved, not only as regards the limited


applicability

of the methods used in calculating the frequencies of the spectrum of

a given system, but especially as regards the question of the polarisation and
intensity

of the emitted spectral lines. These difficulties are intimately connected with

the radical departure from the ordinary ideas of mechanics and electrodynamics

involved in the main principles of the quantum theory, and with the fact that it

has not been possible hitherto to replace these ideas by others forming an equally

consistent and developed structure. Also in this respect, however, great progress

has recently been obtained by the work of \Name{Einstein}\footnote

{\Name{A.~Einstein}, Verh.\ d.\ D. phys.\ Ges.\ XVIII, p.~318 (1916), Phys.\


Zeitschr.\ XVIII, p.~121 (1917).}

and \Name{Ehrenfest}.\footnote

{\Name{P.~Ehrenfest}, Proc.\ Acad.\ Amsterdam, XVI. p.~591 (1914), Phys.\


Zeitschr.\ XV. p.~657 (1914), Ann.\

d.\ Phys.\ LI. p.~327 (1916)\Add{,} Phil.\ Mag.\ XXXIII. p.~500 (1917).}

On this

state of the theory it might therefore be of interest to make an attempt to discuss

77
the different applications from a uniform point of view, and especially to consider

the underlying assumptions in their relations to ordinary mechanics and


electrodynamics.

Such an attempt has been made in the present paper, and it will be shown

that it seems possible to throw some light on the outstanding difficulties by trying

to trace the analogy between the quantum theory and the ordinary theory of

radiation as closely as possible.

The paper is divided into four parts.

\begin{description}

\item{Part I} contains a brief discussion of the general principles of the theory and

deals with the application of the general theory to periodic systems of one

degree of freedom and to the class of non-periodic systems referred to

above.

\item{Part II} contains a detailed discussion of the theory of the hydrogen


spectrum in

order to illustrate the general considerations.

\item{Part III} contains a discussion of the questions arising in connection with the
explanation

of the spectra of other elements.

78
\item{Part IV} contains a general discussion of the theory of the constitution of
atoms and

molecules based on the application of the quantum theory to the nucleus

atom.

\end{description}

{\footnotesize Copenhagen, November 1917.}

\PageSep{5}

\Part{I.}{On the general theory.}

\Section{1.}{General principles.}

The quantum theory of line-spectra rests upon the following fundamental

assumptions:

\begin{Hypothesis}{I.}

That an atomic system can, and can only, exist permanently

in a certain series of states corresponding to a discontinuous series

of values for its energy, and that consequently any change of the

energy of the system, including emission and absorption of electromagnetic

radiation, must take place by a complete transition

between two such states. These states will be denoted as the

79
``stationary states'' of the system.

\end{Hypothesis}

\begin{Hypothesis}{II.}

That the radiation absorbed or emitted during a transition

between two stationary states is ``unifrequentic'' and possesses a

frequency~$\nu$, given by the relation

\[

E' - E'' = h\nu,

\Tag{(1)}

\]

where $h$~is \Name{Planck}'s constant and where $E'$~and $E''$ are the values
of

the energy in the two states under consideration.

\end{Hypothesis}

As pointed out by the writer in the papers referred to in the introduction,

these assumptions offer an immediate interpretation of the fundamental \


Emph{principle

of combination of spectral lines} deduced from the measurements of the

frequencies of the series spectra of the elements. According to the laws discovered

by \Name{Balmer}, \Name{Rydberg} and \Name{Ritz}, the frequencies of the


lines of the series spectrum

of an element can be expressed by a formula of the type:

80
\[

\nu = f_{\tau''}(n'') - f_{\tau'}(n'),

\Tag{(2)}

\]

where $n'$ and $n''$ are whole numbers and $f_{\tau}(n)$~is one among a set of
functions

of~$n$, characteristic for the element under consideration. On the above


assumptions

this formula may obviously be \Typo{interpretated}{interpreted} by assuming that


the stationary states

of an atom of an element form a set of series, and that the energy in the \Ord{$n$}
{th}~state

of the \Ord{$\tau$}{th}~series, omitting an arbitrary constant, is given by

\[

E_{\tau}(n) = -hf_{\tau}(n).

\Tag{(3)}

\]

We thus see that the values for the energy in the stationary states of an atom

may be obtained directly from the measurements of the spectrum by means of

relation~\Eq{(1)}. In order, however, to obtain a theoretical connection between


these

\PageSep{6}

values and the experimental evidence about the constitution of the atom obtained
from

81
other sources, it is necessary to introduce further assumptions about the laws which

govern the stationary states of a given atomic system and the transitions between

these states.

Now\Pagelabel{6} on the basis of a vast amount of experimental evidence, we are


forced

to assume that an atom or molecule consists of a number of electrified particles in

motion, and, since the above fundamental assumptions imply that no emission of

radiation takes place in the stationary states, we must consequently assume that

\Emph{the ordinary laws of electrodynamics cannot be applied} to these states

without radical alterations. In many cases, however, the effect of that part of the

electrodynamical forces which is connected with the emission of radiation will

at any moment be very small in comparison with the effect of the simple
electrostatic

attractions or repulsions of the charged particles corresponding to \


Name{Coulomb}'s

law. Even if the theory of radiation must be completely altered, it is therefore a

natural assumption that it is possible in such cases to obtain a close approximation

in the description of the motion in the stationary states, by retaining only the

latter forces. In the following we shall therefore, as in all the papers mentioned

in the introduction, for the present \Emph{calculate the motions of the particles

in the stationary states as the motions of mass-points according to

ordinary mechanics} including the modifications claimed by the theory of


relativity,

82
and we shall later in the discussion of the special applications come back to

the question of the degree of approximation which may be obtained in this way.

If next we consider a transition between two stationary states, it is obvious

at once from the essential discontinuity, involved in the assumptions I~and~II, that

in general it is impossible even approximately to describe this phenomenon by

means of ordinary mechanics or to calculate the frequency of the radiation


absorbed

or emitted by such a process by means of ordinary electrodynamics. On the other

hand, from the fact that it has been possible by means of ordinary mechanics and

electrodynamics to account for the phenomenon of temperature-radiation in the

limiting region of slow vibrations, we may expect that any theory capable of
describing

this phenomenon in accordance with observations will form some sort of natural

generalisation of the ordinary theory of radiation. Now the theory of temperature-


radiation

in the form originally given by \Name{Planck} confessedly lacked internal

consistency, since, in the deduction of his radiation formula, assumptions of

similar character as I~and~II were used in connection with assumptions which


were

in obvious contrast to them. Quite recently, however, \Name{Einstein}\footnote

{\Name{A.~Einstein}, loc.\ cit.}

has succeeded,

83
on the basis of the assumptions I~and~II, to give a consistent and instructive
deduction

of \Name{Planck}'s formula by introducing certain supplementary assumptions


about

the \Emph{probability of transition of a system between two stationary states}

and about the manner in which this probability depends on the density of radiation

\PageSep{7}

of the corresponding frequency in the surrounding space, suggested from

analogy with the ordinary theory of radiation. \Name{Einstein}\Pagelabel{7}


compares the emission or

absorption of radiation of frequency~$\nu$ corresponding to a transition between


two

stationary states with the emission or absorption to be expected on ordinary

electrodynamics for a system consisting of a particle executing harmonic


vibrations

of this frequency. In analogy with the fact that on the latter theory such a system

will without external excitation emit a radiation of frequency~$\nu$, \


Name{Einstein} assumes

in the first place that on the quantum theory there will be a certain
probability~$A_{n''}^{n'}\, dt$

that the system in the stationary state of greater energy, characterised by

the letter~$n'$, in the time interval~$dt$ will start \Emph{spontaneously} to pass


to the stationary

state of smaller energy, characterised by the letter~$n''$. Moreover, on ordinary

electrodynamics the harmonic vibrator will, in addition to the above mentioned

84
independent emission, in the presence of a radiation of frequency~$\nu$ in the
surrounding

space, and dependent on the accidental phase-difference between this

radiation and the vibrator, emit or absorb radiation-energy. In analogy with this,

\Name{Einstein} assumes secondly that in the presence of a radiation in the


surrounding

space, the system will on the quantum theory, in addition to the above mentioned

probability of spontaneous transition from the state~$n'$ to the state~$n''$, possess

a certain probability, depending on this radiation, of passing in the time~$dt$

from the state~$n'$ to the state~$n''$, as well as from the state~$n''$ to the
state~$n'$. These

latter probabilities are assumed to be proportional to the intensity of the


surrounding

radiation and are denoted by $\rho_{\nu} B_{n''}^{n'}\, dt$ and $\rho_{\nu}


B_{n'}^{n''}\, dt$ respectively, where $\rho_{\nu}\, d\nu$~denotes

the amount of radiation in unit volume of the surrounding space distributed on

frequencies between $\nu$~and $\nu + d\nu$, while $B_{n''}^{n'}$~and


$B_{n'}^{n''}$ are constants which, like~$A_{n''}^{n'}$,

depend only on the stationary states under consideration. \Name{Einstein} does


not introduce

any detailed assumption as to the values of these constants, no more than to the

conditions by which the different stationary states of a given system are


determined

or to the ``a-priori probability'' of these states on which their relative occurrence

in a distribution of statistical equilibrium depends. He shows, however,

85
how it is possible from the above general assumptions, by means of \
Name{Boltzmann}'s

principle on the relation between entropy and probability and \Name{Wien}'s well
known

displacement-law, to deduce a formula for the temperature radiation which apart

from an undetermined constant factor coincides with \Name{Planck}'s, if we only


assume

that the frequency corresponding to the transition between the two states is
determined

by~\Eq{(1)}. It will therefore be seen that by reversing the line of argument,

\Name{Einstein}'s theory may be considered as a very direct support of the latter


relation.

In the following discussion of the application of the quantum theory to determine

the line-spectrum of a given system, it will, just as in the theory of temperature-


radiation,

not be necessary to introduce detailed assumptions as to the mechanism

of transition between two stationary states. We shall show, however, that the

conditions which will be used to determine the values of the energy in the
stationary

states are of such a type that the frequencies calculated by~\Eq{(1)}, in the limit

\PageSep{8}

where the motions in successive stationary states comparatively differ very little

from each other, will tend to coincide with the frequencies to be expected on

the ordinary theory of radiation from the motion of the system in the stationary

86
states. In order to obtain the necessary relation to the ordinary theory of radiation

in the limit of slow vibrations, we are therefore led directly to certain conclusions

about the probability of transition between two stationary states in this limit. This

leads again to certain general considerations about the connection between the

probability of a transition between any two stationary states and the motion of the

system in these states, which will be shown to throw light on the question of the

polarisation and intensity of the different lines of the spectrum of a given system.

In the above considerations we have by an atomic system \Typo{tacidly}{tacitly}


understood

a number of electrified particles which move in a field of force which, with the

approximation mentioned, possesses a potential depending only on the position of

the particles. This may more accurately be denoted as a system under constant

external conditions, and the question next arises about the variation in the
stationary

states which may be expected to take place during a variation of the external

conditions, \eg\ when exposing the atomic system to some variable external field

of force. Now, in general, we must obviously assume that this variation cannot

be calculated by ordinary mechanics, no more than the transition between two

different stationary states corresponding to constant external conditions. If,


however,

the variation of the external conditions is very slow, we may from the necessary

stability of the stationary states expect that the motion of the system at any

given moment during the variation will differ only very little from the motion in

87
a stationary state corresponding to the instantaneous external conditions. If now,

moreover, the variation is performed at a constant or very slowly changing rate,

the forces to which the particles of the system will be exposed will not differ at

any moment from those to which they would be exposed if we imagine that the

external forces arise from a number of slowly moving additional particles which

together with the original system form a system in a stationary state. From this

point of view it seems therefore natural to assume that, with the approximation

mentioned, the motion of an atomic system in the stationary states can be


calculated

by direct application of ordinary mechanics, not only under constant external

conditions, but in general also during a slow and uniform variation of these

conditions. This assumption, which may be denoted as the principle of the

``\Emph{mechanical transformability}'' of the stationary states, has been


introduced

in the quantum theory by \Name{Ehrenfest}\footnote

{\Name{P.~Ehrenfest}, loc.\ cit. In these papers the principle in question is called


the ``adiabatic hypothesis''

in accordance with the line of argumentation followed by \Name{Ehrenfest} in


which considerations

of thermodynamical problems play an important part. From the point of view


taken in the present

paper, however, the above notation might in a more direct way indicate the
content of the principle

and the limits of its applicability.}

and is, as it will be seen in the following

88
sections, of great importance in the discussion of the conditions to be used to fix

\PageSep{9}

the stationary states of an atomic system among the continuous multitude of

mechanically possible motions. In this connection it may be pointed out that

the principle of the mechanical transformability of the stationary states allows

us to overcome a fundamental difficulty which at first sight would seem to

be involved in the definition of the energy difference between two stationary states

which enters in relation~\Eq{(1)}. In fact we have assumed that the direct


transition

between two such states cannot be described by ordinary mechanics, while on

the other hand we possess no means of defining an energy difference between two

states if there exists no possibility for a continuous mechanical connection between

them. It is clear, however, that such a connection is just afforded by \


Name{Ehrenfest}'s

principle which allows us to transform mechanically the stationary states of a

given system into those of another, because for the latter system we may take one

in which the forces which act on the particles are very small and where we may

assume that the values of the energy in all the stationary states will tend to
coincide.

As regards the problem of the statistical distribution of the different stationary

states between a great number of atomic systems of the same kind in temperature

equilibrium, the number of systems present in the different states may be deduced

89
in the well known way from \Name{Boltzmann}'s fundamental relation between
entropy

and probability, if we know the values of the energy in these states and the

\Emph{a-priori probability} to be ascribed to each state in the calculation of the


probability

of the whole distribution. In contrast to considerations of ordinary statistical

mechanics we possess on the quantum theory no direct means of determining these

a-priori probabilities, because we have no detailed information about the


mechanism

of transition between the different stationary states. If the a-priori probabilities

are known for the states of a given atomic system, however, they may be deduced

for any other system which can be formed from this by a continuous
transformation

without passing through one of the singular systems referred to below. In fact,

in examining the necessary conditions for the explanation of the second law of
thermodynamics

\Name{Ehrenfest}\Pagelabel{9}\footnote

{\Name{P.~Ehrenfest}, Phys.\ Zeitschr.\ XV p.~660 (1914). The above


interpretation of this relation is

not stated explicitly by \Name{Ehrenfest}, but it presents itself directly if the


quantum theory is taken in

the form corresponding to the fundamental assumption~I.}

has deduced a certain general condition as regards the

variation of the a-priori probability corresponding to a small change of the external

conditions from which it follows, that the a-priori probability of a given stationary

90
state of an atomic system must remain unaltered during a continuous
transformation,

except in special cases in which the values of the energy in some of the stationary

states will tend to coincide during the transformation. In this result we possess,

as we shall see, a rational basis for the determination of the a-priori probability

of the different stationary states of a given atomic system.

%[** TN: Printer's mark]

% D.K.D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Ard., 8. Række, IV. 1.

\PageSep{10}

\Section{2.}{Systems of one degree of freedom.}

As the simplest illustration of the principles discussed in the former section

we shall begin by considering systems of a single degree of freedom, in which case

it has been possible to establish a general theory of stationary states. This is

due to the fact\Chg{,}{} that \Emph{the motion will be simply periodic}, provided
the

distance between the parts of the system will not increase infinitely with the time,

a case which for obvious reasons cannot represent a stationary state in the sense

defined above. On account of this, the discussion of the mechanical


transformability

of the stationary states can, as pointed out by \Name{Ehrenfest},\footnote

91
{\Name{P.~Ehrenfest}, loc.\ cit.\ Proc.\ Acad.\ Amsterdam, XVI, p.~591
(1914).}

for systems of

one degree of freedom be based on a mechanical theorem about periodic systems


due

to \Name{Boltzmann} and originally applied by this author in a discussion of the


bearing of

mechanics on the explanation of the laws of thermodynamics. For the sake of the

considerations in the following sections it will be convenient here to give the proof

in a form which differs slightly from that given by \Name{Ehrenfest}, and which
takes

also regard to the modifications in the ordinary laws of mechanics claimed by the

theory of relativity.

Consider for the sake of generality a conservative mechanical system of


$s$~degrees

of freedom, the motion of which is governed by \Name{Hamilton}'s equations:

\[

\frac{dp_{k}}{dt} = -\frac{\dd E}{\dd q_{k}},\quad

\frac{dq_{k}}{dt} = \frac{\dd E}{\dd p_{k}},\qquad

(k = 1, \dots, s)

\Tag{(4)}

\]

where $E$~is the total energy considered as a function of the generalised


positional

92
coordinates $q_{1}$,~\MyDots $q_{s}$ and the corresponding canonically
conjugated momenta $p_{1}$,~\MyDots $p_{s}$.\Pagelabel{10}

If the velocities are so small that the variation in the mass of the particles due to

their velocities can be neglected, the $p$'s~are defined in the usual way by

\[

p_{k} = \frac{\dd T}{\dd q_{k}},\qquad

(k = 1, \dots, s)

\]

where $T$~is the kinetic energy of the system considered as a function of the

generalised velocities $\dot{q}_{1}$,~\MyDots $\dot{q}_{s}$ $\left(\dot{q}_{k}


= \dfrac{dq_{k}}{dt}\right)$ and of $q_{1}$,~\MyDots $q_{s}$. If the relativity
modifications

are taken into account the $p$'s~are defined by a similar set of expressions

in which the kinetic energy is replaced by

% [** TN: Set in-line in the original]

\[

T' = \sum m_{0} c^{2}\bigl(1 - \sqrt{1 - v^{2}/c^{2}}\bigr),

\]

where

the summation is to be extended over all the particles of the system, and $v$~is the

velocity of one of the particles and $m_{0}$~its mass for zero velocity, while
$c$~is the

velocity of light.

93
Let us now assume that the system performs a periodic motion with the

period~$\sigma$, and let us form the expression

\[

I = \int_{0}^{\sigma} \sum_{1}^{s} p_{k} \dot{q}_{k}\, dt,

\Tag{(5)}

\]

\PageSep{11}

which is easily seen to be independent of the special choice of coordinates


$q_{1}$,~\MyDots $q_{s}$

used to \Typo{decribe}{describe} the motion of the system. In fact, if the variation


of the mass

with the velocity is neglected we get

\[

I = 2 \int_{0}^{\sigma} T\, dt,

\]

and if the relativity modifications are included, we get a quite analogous


expression

in which the kinetic energy is replaced by $T'' = \sum \frac{1}{2} m_{0} v^{2} \
sqrt{1 - v^{2}/c^{2}}$.

Consider next some new periodic motion of the system formed by a small

variation of the first motion, but which may need the presence of external forces

in order to be a mechanically possible motion. For the variation in~$I$ we get then

\[

94
\delta I = \int_{0}^{\sigma} \sum_{1}^{s} (\dot{q}_{k}\, \delta p_{k} + p_{k}\, \
delta \dot{q}_{k})\, dt

+ \sum_{1}^{s} p_{k}\, \dot{q}_{k}\, \delta t\Big|_{0}^{\sigma},

\]

where the last term refers to the variation of the limit of the integral due to the

variation in the period~$\sigma$. By partial integration of the second term in the


bracket

under the integral we get next

\[

\delta I = \int_{0}^{\sigma} \sum_{1}^{s} (\dot{q}_{k}\, \delta p_{k} - \


dot{p}_{k}\, \delta q_{k})\, dt

+ \sum_{1}^{s} p_{k}(\dot{q}_{k}\, \delta t + \delta q_{k})\Big|_{0}^{\sigma},

\]

where the last term is seen to be zero, because the term in the bracket as well
as~$p_{k}$

will be the same in both limits, since the varied motion as well as the original

motion is assumed to be periodic. By means of equations~\Eq{(4)} we get


therefore

\[

\delta I = \int_{0}^{\sigma} \sum_{1}^{s}

\left(\frac{\dd E}{\dd p_{k}}\, \delta p_{k}

+ \frac{\dd E}{\dd q_{k}}\, \delta q_{k}\right) dt

= \int_{0}^{\sigma} \delta E\, dt.

\Tag{(6)}

95
\]

Let\Pagelabel{11} us now assume that the small variation of the motion is


produced by a

small external field established at a uniform rate during a time interval~$\theta$,


long

compared with~$\sigma$, so that the comparative increase during a period is very


small.

In this case $\delta E$~is at any moment equal to the total work done by the
external

forces on the particles of the system since the beginning of the establishment of

the field. Let this moment be $t = -\theta$ and let the potential of the external field

at $t \geq 0$ be given by~$\Omega$, expressed as a function of the~$q$'s. At any


given moment

$t > 0$ we have then

\[

% [** TN: Correction to integrand numerator retrieved October 17, 2014 from

% web.ihep.su/dbserv/compas/src/bohr18/eng.pdf]

\delta E

= -\int_{-\theta}^{0} \frac{\theta \Add{ + t}}{\theta} \sum_{1}^{s} \frac{\dd \


Omega}{\dd q_{k}}\, \dot{q}_{k}\, dt

- \int_{0}^{t} \sum_{1}^{s} \frac{\dd \Omega}{\dd q_{k}}\, \dot{q}_{k}\, dt,

\]

which gives by partial integration

\PageSep{12}

96
\[

\delta E = \frac{1}{\theta} \int_{-\theta}^{0} \Omega\, dt - \Omega_{t},

\]

where the values for the $q$'s to be introduced in~$\Omega$ in the first term are
those

corresponding to the motion under the influence of the increasing external field,
and

the values to be introduced in the second term are those corresponding to the

configuration at the time~$t$. Neglecting small quantities of the same order as the

square of the external force, however, we may in this expression for~$\delta E$


instead of

the values for the $q$'s corresponding to the perturbed motion take those
corresponding

to the original motion of the system. With this approximation the first term

is equal to the mean value of the second taken over a period~$\sigma$, and we
have

consequently

\[

\int_{0}^{\sigma} \delta E\, dt = 0.

\Tag{(7)}

\]

From \Eq{(6)} and \Eq{(7)} it follows that $I$~will remain constant during the
slow

establishment of the small external field, if the motion corresponding to a constant

97
value of the field is periodic. If next the external field corresponding to~$\Omega$
is considered

as an inherent part of the system, it will be seen in the same way that $I$~will

remain unaltered during the establishment of a new small external field, and

so on. Consequently \Emph{$I$~will be invariant for any finite transformation

of the system which is sufficiently slowly performed}, provided the

motion at any moment during the process is periodic and the effect of the variation

is calculated on ordinary mechanics.

Before we proceed to the applications of this result we shall mention a simple

consequence of~\Eq{(6)} for systems for which every orbit is periodic independent
of the

initial conditions. In that case we may for the varied motion take an undisturbed

motion of the system corresponding to slightly different initial conditions. This

gives $\delta E$~constant, and from~\Eq{(6)} we get therefore

\[

\delta E = \omega\, \delta I,

\Tag{(8)}

\]

where $\omega = \dfrac{1}{\sigma}$ is the frequency of the motion. This equation


forms a simple relation

between the variations in $E$ and~$I$ for periodic systems, which will be often

used in the following.

98
Returning now to systems of one degree of freedom, we shall take our starting

point from \Name{Planck}'s original theory of a \Emph{linear harmonic vibrator}.


According to

this theory the stationary states of a system, consisting of a particle executing


linear

harmonic vibrations with a constant frequency~$\omega_{0}$ independent of the


energy, are

given by the well known relation

\[

E = nh\omega_{0},

\Tag{(9)}

\]

where $n$~is a positive entire number, $h$~\Name{Planck}'s constant, and


$E$~the total energy

which is supposed to be zero if the particle is at rest.

\PageSep{13}

From \Eq{(8)} it follows at once\Chg{,}{} that \Eq{(9)}~is equivalent to

\[

I = \int_{0}^{\sigma} p\dot{q}\, dt

= \int p\, dq

= nh,

\Tag{(10)}

\]

99
where the latter integral is to be taken over a complete oscillation of~$q$ between

its limits. On the principle of the mechanical transformability of the stationary

states we shall therefore assume, following \Name{Ehrenfest}, that \


Eq{(10)}~holds not only for a

\Name{Planck}'s vibrator but for \emph{any periodic system of one degree of


freedom} which

can be formed in a continuous manner from a linear harmonic vibrator by a gradual

variation of the field of force in which the particle moves. This condition is
immediately

seen to be fulfilled by all such systems in which the motion is of oscillating

type \ie\ where the moving particle during a period passes twice through

any point of its orbit once in each direction. If, however, we confine ourselves to

systems of one degree of freedom, it will be seen that systems in which the

motion is of rotating type, \ie\ where the particle during a period passes only once

through every point of its orbit, cannot be formed in a continuous manner from

a linear harmonic vibrator without passing through singular states in which the

period becomes infinite\Note{[sic]} long and the result becomes ambiguous. We


shall not here

enter more closely on this difficulty which has been pointed out by \
Name{Ehrenfest}, because

it disappears when we consider systems of several degrees of freedom, where we

shall see that a simple generalisation of~\Eq{(10)} holds for any system for which

every motion is periodic.

100
As regards the application of~\Eq{(9)} to statistical problems it was assumed in

\Name{Planck}'s theory that the different states of the vibrator corresponding to


different

values of~$n$ are \Emph{a-priori equally probable}, and this assumption was
strongly

supported by the agreement obtained on this basis with the measurements of the

specific heat of solids at low temperatures. Now it follows from the considerations

of \Name{Ehrenfest}, mentioned in the former section, that the a-priori probability


of a

given stationary state is not changed by a continuous transformation, and we shall

therefore expect that for any system of one degree of freedom the different states

corresponding to different entire values of~$n$ in~\Eq{(10)} are a-priori equally


probable.

As pointed out by \Name{Planck} in connection with the application of~\Eq{(9)},


it is

simply seen that statistical considerations, based on the assumption of equal


probability

for the different states given by~\Eq{(10)}, will show the necessary relation to

considerations of ordinary statistical mechanics in the limit where the latter theory

has been found to give results in agreement with experiments. Let the
configuration

and motion of a mechanical system be characterised by $s$~independent variables

$q_{1}$,~\MyDots $q_{s}$ and corresponding momenta $p_{1}$,~\MyDots


$p_{s}$ and let the state of the system be

101
represented in a $2s$-dimensional phase-space by a point with coordinates
$q_{1}$,~\MyDots $q_{s}$,

$p_{1}$,~\MyDots $p_{s}$. Then, according to ordinary statistical mechanics, the


probability for

this point to lie within a small element in the phase-space is independent of the

\PageSep{14}

position and shape of this element and simply proportional to its volume, defined

in the usual way by

\[

\delta W = \int dq_{1} \dots dq_{s}\, dp_{1} \dots dp_{s}.

\Tag{(11)}

\]

In the quantum theory, however, these considerations cannot be directly applied,

since the point representing the state of a system cannot be displaced continuously

in the $2s$-dimensional phase-space, but can lie only on certain surfaces of lower

dimensions in this space. For systems of one degree of freedom the phase-space is

a two-dimensional surface, and the points representing the states of some system

given by~\Eq{(10)} will be situated on closed curves on this surface. Now, in


general, the

motion will differ considerably for any two states corresponding to successive
entire

values of~$n$ in~\Eq{(10)}, and a simple general connection between the


quantum theory

and ordinary statistical mechanics is therefore out of question. In the limit,


however,

102
where $n$~is large, the motions in successive states will only differ very little

from each other, and it would therefore make little difference whether the points

representing the systems are distributed continuously on the phase-surface or


situated

only on the curves corresponding to~\Eq{(10)}, provided the number of systems


which

in the first case are situated between two such curves is equal to the number which

in the second case lies on one of these curves. But it will be seen that this condition

is just fulfilled in consequence of the above hypothesis of equal a-priori probability

of the different stationary states, because the element of phase-surface

limited by two successive curves corresponding to~\Eq{(10)} is equal to

\[

\BreakMath{%

\delta W = \int dp\, dq

= \left[\int p\, dq\right]_{n} - \left[\int p\, dq\right]_{n-1}

= I_{n} - I_{n-1}

= h,

}{%

\begin{aligned}

\delta W = \int dp\, dq

&= \left[\int p\, dq\right]_{n} - \left[\int p\, dq\right]_{n-1} \\

&= I_{n} - I_{n-1}

= h,

103
\end{aligned}

\Tag{(12)}

\]

so that on ordinary statistical mechanics the probabilities for the point to lie within

any two such elements is the same. We see consequently that the hypothesis of

equal probability of the different states given by~\Eq{(10)} gives the same result
as

ordinary statistical mechanics in all such applications in which the states of the

great majority of the systems correspond to large values of~$n$. Considerations of

this kind have led \Name{Debye}\footnote

{\Name{P.~Debye}, Wolfskehl-Vortrag. Göttingen 1913.}

to point out that condition~\Eq{(10)} might have a general

validity for systems of one degree of freedom, already before \Name{Ehrenfest},


on the

basis of his theory of the mechanical transformability of the stationary states,

had shown that this condition forms the only rational generalisation of \
Name{Planck}'s

condition~\Eq{(9)}.

We shall now discuss the relation between the theory of \Emph{spectra of atomic

systems of one degree of freedom}, based on \Eq{(1)}~and \Eq{(10)}, and the


ordinary

theory of radiation, and we shall see that this relation in several respects shows a

104
close analogy to the relation, just considered, between the statistical applications
of~\Eq{(10)}

and considerations based on ordinary statistical mechanics. Since the values

\PageSep{15}

for the frequency~$\omega$ in two states corresponding to different values of~$n$


in~\Eq{(10)} in

general are different, we see at once that we cannot expect a simple connection

between the frequency calculated by~\Eq{(1)} of the radiation corresponding to a

transition between two stationary states and the motions of the system in these

states, except in the limit where $n$~is very large, and where the ratio between the

frequencies of the motion in successive stationary states differs very little from
unity.

Consider now a transition between the state corresponding to $n = n'$ and the state

corresponding to $n = n''$, and let us assume that $n'$~and $n''$ are large numbers
and

that $n' - n''$ is small compared with $n'$~and~$n''$. In that case we may in~\
Eq{(8)} for

$\delta E$ put $E' - E''$ and for $\delta I$ put $I' - I''$, and we get therefore from \
Eq{(1)}~and \Eq{(10)} for

the frequency of the radiation emitted or absorbed during the transition between

the two states

\[

\nu = \frac{1}{h} (E' - E'')

= \frac{\omega}{h} (I' - I'')

= (n' - n'') \omega.

105
\Tag{(13)}

\]

Now in a stationary state of a periodic system the displacement of the particles

in any given direction may always be expressed by means of a \Name{Fourier}-


series

as a sum of harmonic vibrations:

\[

\xi = \sum C_{\tau} \cos 2\pi(\tau\omega t + c_{\tau}),

\Tag{(14)}

\]

where the $C$'s and $c$'s are constants and the summation is to be extended over
all

positive entire values of~$\tau$. On the ordinary theory of radiation we should


therefore

expect the system to emit a spectrum consisting of a series of lines of frequencies

equal to~$\tau\omega$, but, as it is seen, this is just equal to the series of


frequencies

which we obtain from~\Eq{(13)} by introducing different values for $n' - n''$. As


far as

the frequencies are concerned we see therefore that in the limit where $n$~is large

there exists a close relation between the ordinary theory of radiation and the theory

of spectra based on \Eq{(1)} and~\Eq{(10)}. It may be noticed, however, that,


while on the

106
first theory radiations of the different frequencies~$\tau\omega$ corresponding to
different

values of~$\tau$ are emitted or absorbed at the same time, these frequencies will
on

the present theory, based on the fundamental assumption I~and~II, be connected

with entirely different processes of emission or absorption, corresponding to the

transition of the system from a given state to different neighbouring stationary


states.

In order to obtain the necessary connection, mentioned in the former section,

to the ordinary theory of radiation in the limit of slow vibrations, we must further

claim that a relation, as that just proved for the frequencies, will, in the limit of

large~$n$, hold also for the intensities of the different lines in the spectrum. Since

now on ordinary electrodynamics the intensities of the radiations corresponding

to different values of~$\tau$ are directly determined from the coefficients~$C_{|


tau}$ in~\Eq{(14)}, we

must therefore expect that for large values of~$n$ these coefficients will on the
quantum

theory determine the \Emph{probability of \Typo{spontanuous}{spontaneous}


transition} from a given

stationary state for which $n = n'$ to a neighbouring state for which $n = n'' = n' - \
tau$.

Now this connection between the amplitudes of the different harmonic vibrations
into

\PageSep{16}

107
which the motion can be resolved, characterised by different values of~$\tau$, and
the

probabilities of transition from a given stationary state to the different


neighbouring

stationary states, characterised by different values of $n' - n''$, may clearly be


expected

to be of a general nature. Although, of course, we cannot without a detailed

theory of the mechanism of transition obtain an exact calculation of the latter


probabilities,

unless $n$~is large, we may expect that also for small values of~$n$ the amplitude

of the harmonic vibrations corresponding to a given value of~$\tau$ will in some

way give a measure for the probability of a transition between two states for which

$n' - n''$ is equal to~$\tau$. Thus in general there will be a certain probability of an

atomic system in a stationary state to pass spontaneously to any other state of

smaller energy, but if for all motions of a given system the coefficients~$C$ in~\
Eq{(14)}

are zero for certain values of~$\tau$, we are led to expect that no transition will be

possible, for which $n' - n''$ is equal to one of these values.

A simple illustration of these considerations is offered by the linear harmonic

vibrator mentioned above in connection with \Name{Planck}'s theory. Since in this


case

$C_{\tau}$~is equal to zero for any $\tau$ different from~$1$, we shall expect
that for this system

108
only such transitions are possible in which $n$~alters by one unit. From \
Eq{(1)}~and

\Eq{(9)} we obtain therefore the simple result that the frequency of any radiation
emitted

or absorbed by a linear harmonic vibrator is equal to the constant frequency~$\


omega_{0}$.

This result seems to be supported by observations on the absorption-spectra of

diatomic gases, showing that certain strong absorption-lines, which according to

general evidence may be ascribed to vibrations of the two atoms in the molecule

relative to each other, are not accompanied by lines of the same order of intensity

and corresponding to entire multipla of the frequency, such as it should be


expected

from~\Eq{(1)} if the system had any considerable tendency to pass between non-
successive

states. In this connection it may be noted that the fact, that in the absorption

spectra of some diatomic gases faint lines occur corresponding to the double
frequency

of the main lines,\footnote

{See \Name{E.~C. Kemble}, Phys.\ Rev., VIII, p.~701, 1916.}

obtains a natural explanation by assuming that for

finite amplitudes the vibrations are not exactly harmonic and that therefore the

molecules possess a small probability of passing also between non-successive


states.

109
\Section{3.}{Conditionally periodic systems.}

If we consider systems of several degrees of freedom the motion will be periodic

only in singular cases and the general conditions which determine the stationary
states

cannot therefore be derived by means of the same simple kind of considerations

as in the former section. As mentioned in the introduction, however, \


Name{Sommerfeld}

and others have recently succeeded, by means of a suitable generalisation of~\


Eq{(10)}, to

obtain conditions for an important class of systems of several degrees of freedom,

\PageSep{17}

which, in connection with~\Eq{(1)}, have been found to give results in convincing


agreement

with experimental results about line-spectra. Subsequently these conditions

have been proved by \Name{Ehrenfest} and especially by \Name{Burgers}\


footnote

{\Name{J.~M. Burgers}, Versl.\ Akad.\ Amsterdam, XXV, pp.~849, 918,~1055.


(1917), Ann.\ d.\ Phys.\ LII. p.~195

(1917), Phil.\ Mag.\ XXXIII, p.~514 (1917).}

to be invariant for

slow mechanical transformations.

To the generalisation under consideration we are naturally led, if we first consider

110
such systems for which the motions corresponding to the different degrees of
freedom

are dynamically independent of each other. This occurs if the expression for the

total energy~$E$ in \Name{Hamilton}\Add{'}s equations~\Eq{(4)} for a system


of $s$~degrees of freedom

can be written as a sum $E_{1} + \dots + E_{s}$, where $E_{k}$~contains


$q_{k}$~and $p_{k}$ only. An

illustration of a system of this kind is presented by a particle moving in a field

of force in which the force-components normal to three mutually perpendicular

fixed planes are functions of the distances from these planes respectively. Since in

such a case the motion corresponding to each degree of freedom in general will

be periodic, just as for a system of one degree of freedom, we may obviously


expect

that the condition~\Eq{(10)} is here replaced by a set of $s$~conditions:

\[

I_{k} = \int p_{k}\, dq_{k} = n_{k} h,\qquad

(k = 1, \MyDots s)

\Tag{(15)}

\]

where the integrals are taken over a complete period of the different $q$'s
respectively,

and where $n_{1}$,~\MyDots $n_{s}$ are entire numbers. It will be seen at once
that these conditions

are invariant for any slow transformation of the system for which the independency

of the motions corresponding to the different coordinates is maintained.

111
A more general class of systems for which a similar analogy with systems of

a single degree of freedom exists and where conditions of the same type as~\
Eq{(15)}

present themselves is obtained in the case where, although the motions


corresponding

to the different degrees of freedom are not independent of each other, it is possible

nevertheless by a suitable choice of coordinates to express each of the


momenta~$p_{k}$

as a function of $q_{k}$~only. A simple system of this kind consists of a particle


moving

in a plane orbit in a central field of force. Taking the length of the radius-vector

from the centre of the field to the particle as~$q_{1}$, and the angular distance of
this

radius-vector from a fixed line in the plane of the orbit as~$q_{2}$, we get at once

from~\Eq{(4)}, since $E$~does not contain~$q_{2}$, the well known result that
during the motion

the angular momentum~$p_{2}$ is constant and that the radial motion, given by
the

variations of $p_{1}$~and $q_{1}$ with the time, will be exactly the same as for a
system

of one degree of freedom. In his fundamental application of the quantum theory

to the spectrum of a \Emph{non-periodic system} \Name{Sommerfeld} assumed


therefore that

the stationary states of the above system are given by two conditions of the form:

\[

112
I_{1} = \int p_{1}\, dq_{1} = n_{1} h,\quad

I_{2} = \int p_{2}\, dq_{2} = n_{2} h.

\Tag{(16)}

\]

%[** TN: Printer's mark]

%D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8. Række. IV. 1.

\PageSep{18}

While the first integral obviously must be taken over a period of the radial motion,

there might at first sight seem to be a difficulty in fixing the limits of integration

of~$q_{2}$. This disappears, however, if we notice that an integral of the type


under

consideration will not be altered by a change of coordinates in which $q$~is


replaced

by some function of this variable. In fact, if instead of the angular distance of the

radius-vector we take for~$q_{2}$ some continuous periodic function of this angle


with

period~$2\pi$, every point in the plane of the orbit will correspond to one set of

coordinates only and the relation between $p$~and~$q$ will be exactly of the
same

type as for a periodic system of one degree of freedom for which the motion is of

oscillating type. It follows therefore that the integration in the second of the

conditions~\Eq{(16)} has to be taken over a complete revolution of the radius-


vector,

and that consequently this condition is equivalent with the simple condition that

113
the angular momentum of the particle round the centre of the field is equal to an

entire \Chg{multiplum}{multiple} of~$\dfrac{h}{2\pi}$. As pointed out by \


Name{Ehrenfest}, the conditions~\Eq{(16)} are

invariant for such special transformations of the system for which the central

symmetry is maintained. This follows immediately from the fact that the angular

momentum in transformations of this type remains invariant, and that the equations

of motion for the radial coordinate as long as $p_{2}$~remains constant are the

same as for a system of one degree of freedom. On\Pagelabel{18} the basis of~\
Eq{(16)}, \Name{Sommerfeld}

has, as mentioned in the introduction, obtained a brilliant explanation of the fine

structure of the lines in the hydrogen spectrum, due to the change of the mass

of the electron with its velocity.\footnote

{In this connection it may be remarked that conditions of the same type as~\
Eq{(16)} were proposed

independently by \Name{W.~Wilson} (Phil.\ Mag.\ XXIX p.~795 (1915) and


XXXI p.~156 (1916)), but by him applied

only to the simple Keplerian motion described by the electron in the hydrogen
atom if the relativity

modifications are neglected. Due to the singular position of periodic systems in


the quantum theory

of systems of several degrees of freedom this application, however, involves, as it


will appear from the

following discussion, an ambiguity which deprives the result of an immediate


physical interpretation.

Conditions analogous to~\Eq{(16)} have also been established by \


Name{Planck} in his interesting theory of the

114
``physical structure of the phase space'' of systems of several degrees of freedom
(Verh.\ d.\ D.\ Phys.\

Ges.\ XVII p.~407 and p.~438 (1915), Ann.\ d.\ Phys.\ L p.~385, (1916)\Add{)}.
This theory, which has no direct

relation to the problem of line-spectra discussed in the present paper, rests upon a
profound analysis

of the geometrical problem of dividing the multiple-dimensional phase space


corresponding to a system

of several degrees of freedom into ``cells'' in a way analogous to the division of


the phase surface of

a system of one degree of freedom by the curves given by~\Eq{(10)}.}

To this theory we shall come back in Part~II\@.

As pointed out by \Name{Epstein}\footnote

{\Name{P.~Epstein}, loc.\ cit.}

and \Name{Schwarzschild}\footnote

{\Name{K.~Schwarzschild}, loc.\ cit.}

the central systems considered

by \Name{Sommerfeld} form a special case of a more general class of systems for

which conditions of the same type as~\Eq{(15)} may be applied. These are the so\
Typo{}{ }called

\Emph{conditionally periodic systems}, to which we are led if the equations of


motion

are discussed by means of the \Name{Hamilton}-\Name{Jacobi} partial


differential equation.\footnote

115
{See f.\ inst.\ \Name{C.~V.~L. Charlier}, Die Mechanik des Himmels, Bd.~I,
Abt.~2.}

In

the expression for the total energy~$E$ as a function of the~$q$'s and the~$p$'s,
let the

latter quantities be replaced by the partial differential coefficients of some


function~$S$

\PageSep{19}

with respect to the corresponding $q$'s respectively, and consider the partial

differential equation:\Pagelabel{19}

\[

E \left(q_{1}, \MyDots q_{s}, \frac{\dd S}{\dd q_{1}}, \MyDots \frac{\dd S}{\dd


q_{s}}\right) = \alpha_{1},

\Tag{(17)}

\]

obtained by putting this expression equal to an arbitrary constant~$\alpha_{1}$. If


then

\[

S = F (q_{1}, \MyDots q_{s}, \alpha_{1}, \MyDots \alpha_{s}) + C,

\]

where $\alpha_{2}$,~\MyDots $\alpha_{s}$, and~$C$ are arbitrary constants


like~$\alpha_{1}$, is a total integral of~\Eq{(17)}, we

get, as shown by \Name{Hamilton} and \Name{Jacobi}, the general solution of the


equations of

motion~\Eq{(4)} by putting

116
\[

\frac{\dd S}{\dd \alpha_{1}} = t + \beta_{1},\quad

\frac{\dd S}{\dd \alpha_{k}} = \beta_{k},\qquad (k=2, \MyDots s)

\Tag{(18)}

\]

and

\[

\frac{\dd S}{\dd q_{k}} = p_{k},\qquad (k = 1, \MyDots s)

\Tag{(19)}

\]

where $t$~is the time and $\beta_{1}$,~\MyDots $\beta_{s}$ a new set of


arbitrary constants. By means of~\Eq{(18)}

the $q$'s are given as functions of the time~$t$ and the $2s$~constants $\
alpha_{1}$,~\MyDots $\alpha_{s}$,

$\beta_{1}$,~\MyDots $\beta_{s}$ which may be determined for instance from


the values of the $q$'s and

$\ddot{q}$'s at a given moment.

Now the class of systems, referred to, is that for which, for a suitable choice

of orthogonal coordinates, it is possible to find a total integral of~\Eq{(17)} of the


form

\[

S = \sum_{1}^{s} S_{k}(q_{k}, \alpha_{1}, \MyDots \alpha_{s}),

\Tag{(20)}

117
\]

where $S_{k}$~is a function of the $s$~constants $\alpha_{1}$,~\MyDots $\


alpha_{s}$ and of $q_{k}$~only. In this case, in

which the equation~\Eq{(17)} allows of what is called ``separation of variables'',


we get

from~\Eq{(19)} that every~$p$ is a function of the~$\alpha$'s and of the


corresponding $q$ only. If

during the motion the coordinates do not become infinite in the course of time

or converge to fixed limits, every~$q$ will, just as for systems of one degree of
freedom,

oscillate between two fixed values, different for the different~$q$'s and depending
on

the~$\alpha$'s. Like in the case of a system of one degree of freedom, $p_{k}


$~will become

zero and change its sign whenever $q_{k}$~passes through one of these limits.
Apart

from special cases, the system will during the motion never pass twice through

a configuration corresponding to the same set of values for the $q$'s and~$p$'s, but
it

will in the course of time pass within any given, however small, distance from any

configuration corresponding to a given set of values $q_{1}$,~\MyDots $q_{s}$,


representing a point

within a certain closed $s$-dimensional extension limited by $s$~pairs of $(s -


1)$-dimensional

surfaces corresponding to constant values of the~$q$'s equal to the above


mentioned

limits of oscillation. A motion of this kind is called ``conditionally periodic''. It


118
will be seen that the character of the motion will depend only on the~$\alpha$'s
and not

on the~$\beta$'s, which latter constants serve only to fix the exact configuration of
the

\PageSep{20}

system at a given moment, when the~$\alpha$'s are known. For special systems it
may

occur that the orbit will not cover the above mentioned $s$-dimensional extension

everywhere dense, but will, for all values of the~$\alpha$'s, be confined to an


extension

of less dimensions. Such a case we will refer to in the following as a case of

``degeneration''.

Since for a conditionally periodic system which allows of separation in the

variables $q_{1}$,~\MyDots $q_{s}$ the $p$'s are functions of the corresponding


$q$'s only, we may,

just as in the case of independent degrees of freedom or in the case of quasi-


periodic

motion in a central field, form a set of expressions of the type

\[

I_{k} = \int p_{k}(q_{k}, \alpha_{1}, \MyDots \alpha_{s})\, dq_{k},\qquad

(k = 1, \MyDots s)

\Tag{(21)}

\]

119
where the integration is taken over a complete oscillation of~$q_{k}$. As, in
general, the

orbit will cover everywhere dense an $s$-dimensional extension limited in the


characteristic

way mentioned above, it follows that, except in cases of degeneration, a

separation of variables will not be possible for two different sets of coordinates

$q_{1}$,~\MyDots $q_{s}$ and $q'_{1}$,~\MyDots $q'_{s}$, unless $q_{1} =


f_{1}(q'_{1})$,~\MyDots $q_{s} = f_{s}(q'_{s})$, and since a change of

coordinates of this type will not affect the values of the expressions~\Eq{(21)}, it
will be

seen that the values of the~$I$'s are completely determined for a given motion of
the

system. By putting

\[

I_{k} = n_{k} h,\qquad

(k = l, \MyDots s)

\Tag{(22)}

\]

where $n_{1}$,~\MyDots $n_{s}$ are positive entire numbers, we obtain


therefore a \Emph{set of conditions

which form a natural generalisation of condition~\Eq{(10)}} holding

for a system of one degree of freedom.

Since the~$I$'s, as given by~\Eq{(21)}, depend on the constants $\alpha_{1}$,~\


MyDots $\alpha_{s}$ only and

120
not on the~$\beta$'s, the $\alpha$'s may, in general, inversely be determined from
the values

of the~$I$'s. The character of the motion will therefore, in general, be completely

determined by the conditions~\Eq{(22)}, and especially the value for the total
energy, which

according to~\Eq{(17)} is equal to~$\alpha_{1}$, will be fixed by them. In the


cases of degeneration

referred to above, however, the conditions~\Eq{(22)} involve an ambiguity, since


in

general for such systems there will exist an infinite number of different sets of

coordinates which allow of a separation of variables, and which will lead to


different

motions in the stationary states, when these conditions are applied. As we shall

see below, this ambiguity will not influence the fixation of the total energy in the

stationary states, which is the essential factor in the theory of spectra based on~\
Eq{(1)}

and in the applications of the quantum theory to statistical problems.

A well known characteristic example of a conditionally periodic system is

afforded by a particle moving under the influence of the attractions from two

fixed centres varying as the inverse squares of the distances apart, if the relativity

modifications are neglected. As shown by \Name{Jacobi} this problem can be


solved by a

separation of variables if so called elliptical coordinates\Pagelabel{21} are used, \


ie\ if for $q_{l}$~and~$q_{2}$

121
we take two parameters characterising respectively an ellipsoid and a hyperboloid

of revolution with the centres as foci and passing through the instantaneous
position

\PageSep{21}

of the moving particle, and for~$q_{3}$ we take the angle between the plane
through

the particle and the centres and a fixed plane through the latter points, or, in

closer conformity with the above general description, some continuous periodic
function

of this angle with period~$2\pi$. A limiting case of this problem is afforded by

an electron rotating round a positive nucleus and subject to the effect of an


additional

homogeneous electric field, because this field may be considered as arising

from a second nucleus at infinite distance apart from the first. The motion in this

case will therefore be conditionally periodic and allow a separation of variables in

parabolic coordinates, if the nucleus is taken as focus for both sets of paraboloids

of revolution, and their axes are taken parallel to the direction of the electric force.

By applying the conditions~\Eq{(22)} to this motion \Name{Epstein} and \


Name{Schwarzschild} have,

as mentioned in the introduction, independent of each other, obtained an


explanation

of the effect of an external electric field on the lines of the hydrogen spectrum,

which was found to be in convincing agreement with \Name{Stark}'s


measurements. To

the results of these calculations we shall return in Part~II\@.

122
In the above way of representing the general theory we have followed the

same procedure as used by \Name{Epstein}. By introducing the so called ``angle-


variables''

well known from the astronomical theory of perturbations, \Name{Schwarzschild}


has given

the theory a very elegant form in which the analogy with systems of one degree

of freedom presents itself in a somewhat different manner. The connection


between

this treatment and that given above has been discussed in detail by \
Name{Epstein}.\footnote

{\Name{P.~Epstein}, Ann.\ d.\ Phys.\ LI, p.~168 (1916). See also Note on \
PageRef{29} of the present paper.}

As mentioned above the conditions~\Eq{(22)}, first established from analogy with

systems of one degree of freedom, have subsequently been proved generally to be

\Emph{mechanically invariant for any slow transformation for which the

system remains conditionally periodic}. The proof of this in variance has

been given quite recently by \Name{Burgers}\footnote

{\Name{J.~M. Burgers}, loc.\ cit.\ Versl.\ Akad.\ Amsterdam, XXV, p.~1055


(1917).}

by means of an interesting application of the

theory of contact-transformations based on \Name{Schwarzschild}'s introduction


of angle

123
variables. We shall not enter here on these calculations but shall only consider
some

points in connection with the problem of the mechanical transformability of the

stationary states which are of importance for the logical consistency of the general

theory and for the later applications. In~\SecRef{2} we saw that in the proof of the
mechanical

invariance of relation~\Eq{(10)} for a periodic system of one degree of freedom, it


was

essential that the comparative variation of the external conditions during the time
of one

period could be made small. This may be regarded as an immediate consequence


of the

nature of the fixation of the stationary states in the quantum theory. In fact the
answer

to the question, whether a given state of a system is stationary, will not depend

only on the motion of the particles at a given moment or on the field of force in

the immediate neighbourhood of their instantaneous positions, but cannot be given

before the particles have passed through a complete cycle of states, and so to speak

\PageSep{22}

have got to know the entire field of force of influence on the motion. If thus, in

the case of a periodic system of one degree of freedom, the field of force is varied

by a given amount, and if its comparative variation within the time of a single

period was not small, the particle would obviously have no means to get to know

the nature of the variation of the field and to adjust its stationary motion to it,

before the new field was already established. For exactly the same reasons it is a

124
necessary condition for the mechanical invariance of the stationary states of a

conditionally periodic system, that the alteration of the external conditions during

an interval in which the system has passed approximately through all possible

configurations within the above mentioned $s$-dimensional extension in the


coordinate-space

can be made as small as we like. This condition forms therefore also

an essential point in \Name{Burgers}' proof of the invariance of the conditions~\


Eq{(22)} for

mechanical transformations. Due to this we meet with a characteristic difficulty

when during the transformation of the system we pass one of the cases of

degeneration mentioned above, where, for every set of values for the~$\alpha$'s,
the orbit

will not cover the $s$-dimensional extension everywhere dense, but will be
confined

to an extension of less dimensions. It is clear that, when by a slow transformation

of a conditionally periodic system we approach a degenerate system of this

kind, the time-interval which the orbit takes to pass close to any possible
configuration

will tend to be very long and will become infinite when the degenerate

system is reached. As a consequence of this \Emph{the conditions~\Eq{(22)} will


generally

not remain mechanically invariant when we pass a degenerate

system}, what has intimate connection with the above mentioned ambiguity in the

determination of the stationary states of such systems by means of~\Eq{(22)}.

125
A typical case of a degenerate system, which may serve as an illustration of

this point, is formed by a system of several degrees of freedom for which every

motion is \Emph{simply periodic}, independent of the initial conditions. In this


case, which

is of great importance in the physical applications, we have from \Eq{(5)}~and \


Eq{(21)}, for

any set of coordinates in which a separation of variables is possible,

\[

I = \int_{0}^{\sigma} (p_{1} \dot{q}_{1} + \dots + p_{s} \dot{q}_{s})\, dt

= \kappa_{1} I_{1} + \dots + \kappa_{s} I_{s},

\Tag{(23)}

\]

where the integration is extended over one period of the motion, and where

$\kappa_{1}$,~\MyDots $\kappa_{s}$ are a set of positive entire numbers without


a common divisor. Now we

shall expect that every motion, for which it is possible to find a set of coordinates
in

which it satisfies~\Eq{(22)}, will be stationary. For any such motion we get from~\
Eq{(23)}\Pagelabel{22}

\[

I = (\kappa_{1} n_{1} + \dots + \kappa_{s} n_{s}) h = nh,

\Tag{(24)}

\]

where $n$~is a whole number which may take all positive values if, as in the
applications

126
mentioned below, at least one of the~$\kappa$'s is equal to one. Inversely, if

the system under consideration allows of separation of variables in an infinite


continuous

multitude of sets of coordinates, we must conclude that generally every

\PageSep{23}

motion which satisfies~\Eq{(24)} will be stationary, because in general it will be


possible

for any such motion to find a set of coordinates in which it satisfies also~\
Eq{(22)}.

It will thus be seen that, for a periodic system of several degrees of freedom,

condition~\Eq{(24)} forms a simple generalisation of condition~\Eq{(10)}. From


relation~\Eq{(8)},

which holds for two neighbouring motions of any periodic system, it follows

further that the energy of the system will be completely determined by the value

of~$I$, just as for systems of one degree of freedom.

Consider\Pagelabel{23} now a periodic system in some stationary state


satisfying~\Eq{(24)}, and

let us assume that an external field is slowly established at a continuous rate and

that the motion at any moment during this process allows of a separation of
variables

in a certain set of coordinates. If we would assume that the effect of the

field on the motion of the system at any moment could be calculated directly by

means of ordinary mechanics, we would find that the values of the~$I$'s with
respect to

127
the latter coordinates would remain constant during the process, but this would

involve that the values of the~$n$'s in~\Eq{(22)} would in general not be entire
numbers,

but would depend entirely on the accidental motion, satisfying~\Eq{(24)},


originally possessed

by the system. That mechanics, however, cannot generally be applied directly

to determine the motion of a periodic system under influence of an increasing


external

field, is just what we should expect according to the singular position of

degenerate systems as regards mechanical transformations. In fact, in the presence

of a small external field, the motion of a periodic system will undergo slow
variations

as regards the shape and position of the orbit, and if the perturbed motion is

conditionally periodic these variations will be of a periodic nature. Formally, we

may therefore compare a periodic system exposed to an external field with a


simple

mechanical system of one degree of freedom in which the particle performs a slow

oscillating motion. Now the frequency of a slow variation of the orbit will be seen

to be proportional to the intensity of the external field, and it is therefore obviously

impossible to establish the external field at a rate so slow that the comparative

change of its intensity during a period of this variation is small. The process which

takes place during the increase of the field will thus be analogous to that which

takes place if an oscillating particle is subject to the effect of external forces which

change considerably during a period. Just as the latter process generally will give

128
rise to emission or absorption of radiation and cannot be described by means of

ordinary mechanics, we must expect that the motion of a periodic system of several

degrees of freedom under the establishment of the external field cannot be


determined

by ordinary mechanics, but that the field will give rise to effects of the same kind

as those which occur during a transition between two stationary states


accompanied

by emission or \Typo{absorbtion}{absorption} of radiation. Consequently we shall


expect that, during

the establishment of the field, \Emph{the system will in general adjust itself

in some unmechanical way} until a stationary state is reached in which the

frequency (or frequencies) of the above mentioned slow variation of the orbit has a

relation to the additional energy of the system due to the presence of the external

\PageSep{24}

field, which is of the same kind as the relation, expressed by \Eq{(8)}~and~\


Eq{(10)}, between

the energy and frequency of a periodic system of one degree of freedom. As it will

be shown in Part~II in connection with the physical applications, this condition

is just secured if the stationary states in the presence of the field are determined

by the conditions~\Eq{(22)}, and it will be seen that these considerations offer a


means

of fixing the stationary states of a perturbed periodic system also in cases where

no separation of variables can be obtained.

129
In consequence of the singular position of the degenerate systems in the

general theory of stationary states of conditionally periodic systems, we obtain

a means of \Emph{connecting mechanically two different stationary states

of a given system} through a continuous series of stationary states without

passing through systems in which the forces are very small and the energies in all

the stationary states tend to coincide (comp.\ \PageRef{9}). In fact, if we consider


a given

conditionally periodic system which can be transformed in a continuous way into

a system for which every orbit is periodic and for which every state satisfying~\
Eq{(24)}

will also satisfy~\Eq{(22)} for a suitable choice of coordinates, it is clear in the


first

place that it is possible to pass in a mechanical way through a continuous series

of stationary states from a state corresponding to a given set of values of the~$n$'s

in~\Eq{(22)} to any other such state for which $\kappa_{1} n_{1} + \dots + \
kappa_{s} n_{s}$ possesses the same value. If,

moreover, there exists a second periodic system of the same character to which the

first periodic system can be transformed continuously, but for which the set of~$\
kappa$'s

is different, it will be possible in general by a suitable cyclic transformation to

pass in a mechanical way between any two stationary states of the given
conditionally

periodic system satisfying~\Eq{(22)}.

\begin{Remark}

130
To obtain an example of such a cyclic transformation let us take the system
consisting

of an electron which moves round a fixed positive nucleus exerting an attraction


varying as

the inverse square of the distance. If we neglect the small relativity corrections,
every orbit

will be periodic independent of the initial conditions and the system will allow of
separation

of variables in polar coordinates as well as in any set of elliptical coordinates, of


the

kind mentioned on \PageRef{21}, if the nucleus is taken as one of the foci. It is


simply seen that

any orbit which satisfies~\Eq{(24)} for a value of $n > 1$, will also satisfy~\
Eq{(22)} for a suitable choice

of elliptical coordinates. By imagining another nucleus of infinite small charge


placed at the

other focus, the orbit may further be transformed into another which satisfies~\
Eq{(24)} for the

same value of~$n$, but which may have any given value for the eccentricity.
Consider now

a state of the system satisfying~\Eq{(21)}, and let us assume that by the above
means the orbit

is originally so adjusted that in plane polar coordinates it will correspond to


$n_{1} = m$ and

$n_{2} = n - m$ in~\Eq{(16)}. Let then the system undergo a slow continuous


transformation during which

the field of force acting on the electron remains central, but by which the law of
attraction
131
is slowly varied until the force is directly proportional to the distance apart. In the
final

state, as well as in the original state, the orbit of the electron will be closed, but
during the

transformation the orbit will not be closed, and the ratio between the mean period
of

revolution and the period of the radial motion, which in the original motion was
equal to

one, will during the transformation increase continuously until in the final state it is
equal to

two. This means that, using polar coordinates, the values of $\kappa_{1}$~and~$\
kappa_{2}$ in~\Eq{(22)} which for the

first state are equal to $\kappa_{1} = \kappa_{2} = 1$, will be for the second state
$\kappa_{1} = 2$ and $\kappa_{2} = 1$. Since

during the transformation $n_{1}$~and~$n_{2}$ will keep their values, we get


therefore in the final state

$I = h\bigl(2m + (n - m)\bigr) = h(n + m)$. Now in the latter state, the system
allows a separation of

\PageSep{25}

variables not only in polar coordinates but also in any system of rectangular
Cartesian

coordinates, and by suitable choice of the direction of the axes, we can obtain that
any orbit,

satisfying~\Eq{(24)} for a value of $n > l$, will also satisfy~\Eq{(22)}. By an


infinite small change of the

force components in the directions of the axes, in such a way that the motions in
these

132
directions remain independent of each other but possess slightly different periods,
it will

further be possible to transform the elliptical orbit mechanically into one


corresponding to

any given ratio between the axes. Let us now assume that in this way the orbit of
the

electron is transformed into a circular one, so that, returning to plane polar


coordinates, we

have $n_{1} = 0$ and $n_{2} = n + m$, and let then by a slow transformation the
law of attraction

be varied until again it is that of the inverse square. It will be seen that when this
state

is reached the motion will again satisfy~\Eq{(24)}, but this time we will have $I =
h (n + m)$ instead

of $I = nh$ as in the original state. By repeating a cyclic process of this kind we


may pass

from any stationary state of the system in question which satisfies~\Eq{(24)} for a
value of $n > 1$

to any other such state without leaving at any moment the region of stationary
states.

\end{Remark}

The theory of the mechanical transformability of the stationary states gives us a

means to discuss the question of the \Emph{a-priori probability} of the different


states

of a conditionally periodic system, characterised by different sets of values for


the~$n$'s

133
in~\Eq{(22)}. In fact from the considerations, mentioned in~\SecRef{1}, it follows
that, if

the a-priori probability of the stationary states of a given system is known, it

is possible at once to deduce the probabilities for the stationary states of any

other system to which the first system can be transformed continuously without

passing through a system of degeneration. Now from the analogy with systems of

one degree of freedom it seems necessary to assume that, for a system of several

degrees of freedom for which the motions corresponding to the different


coordinates

are dynamically independent of each\Typo{}{ }other, the a-priori probability is the


same for all

the states corresponding to different sets of~$n$'s in~\Eq{(15)}. According to the


above

we shall therefore assume that the a-priori probability is the same for all states,

given by~\Eq{(22)}, of any system which can be formed in a continuous way from
a system

of this kind without passing through systems of degeneration. It will be observed

that on this assumption we obtain exactly the same relation to the ordinary theory

of statistical mechanics in the limit of large~$n$'s as obtained in the case of


systems

of one degree of freedom. Thus, for a conditionally periodic system, the volume

given by~\Eq{(11)} of the element of phase-space, including all points $q_{1}$,~\


MyDots $q_{s}$, $p_{1}$,~\MyDots $p_{s}$

which represent states for which the value of~$I_{k}$ given by~\Eq{(21)} lies
between $I_{k}$

134
and~$I_{k} + \delta I_{k}$, is seen at once to be equal to\footnote

{Comp.\ \Name{A.~Sommerfeld}, Ber.\ Akad.\ München, 1917, p.~83.}

\[

\delta W = \delta I_{1}\, \delta I_{2} \dots \delta I_{s},

\Tag{(25)}

\]

if the coordinates are so chosen that the motion corresponding to every degree of

freedom is of oscillating type. The volume of the phase-space limited by $s$~pairs

of surfaces, corresponding to successive values for the~$n$'s in the conditions~\


Eq{(22)},

will therefore be equal to~$h^{s}$ and consequently be the same for every
combination

of the~$n$'s. In the limit where the~$n$'s are large numbers and the stationary
states

%[** TN: Printer's mark]

%D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd 8. Række, IV. 1.

\PageSep{26}

corresponding to successive values for the~$n$'s differ only very little from each

other, we thus obtain the same result on the assumption of equal a-priori
probability

of all the stationary states, corresponding to different sets of values of

$n_{1}$,~\MyDots $n_{s}$ in~\Eq{(22)}, as would be obtained by application of


ordinary statistical

mechanics.

135
The\Pagelabel{26} fact that the last considerations hold for every non-degenerate
conditionally

periodic system suggests the assumption that in general \Emph{the a-priori

probability will be the same for all the states determined by~\Eq{(22)}},

even if it should not be possible to transform the given system into a system of

independent degrees of freedom without passing through degenerate systems. This

assumption will be shown to be supported by the consideration of the intensities

of the different components of the \Name{Stark}-effect of the hydrogen lines,


mentioned

in the next Part. When we consider a degenerate system, however, we cannot

assume that the different stationary states are a-priori equally probable. In such a

case the stationary states will be characterised by a number of conditions less

than the number of degrees of freedom, and the probability of a given state must

be determined from the number of different stationary states of some non-


degenerate

system which will coincide in the given state, if the latter system is continuously

transformed into the degenerate system under consideration.

In order to illustrate this, let us take the simple case of a degenerate system

formed by an electrified particle moving in a plane orbit in a central field, the

stationary states of which are given by the two conditions~\Eq{(16)}. In this case
the

plane of the orbit is undetermined, and it follows already from a comparison with

136
ordinary statistical mechanics, that the a-priori probability of the states \
Chg{characterized}{characterised}

by different combinations of $n_{1}$ and $n_{2}$ in~\Eq{(16)} cannot be the


same. Thus the

volume of the phase-space, corresponding to states for which $I_{1}$~lies


between $I_{1}$~and

$I_{1} + \delta I_{1}$ and for which $I_{2}$~lies between $I_{2}$~and $I_{2} +
\delta I_{2}$, is found by a

simple calculation\footnote

{See \Name{A.~Sommerfeld}, loc.\ cit.}

to be equal to $\delta W = 2I_{1}\, \delta I_{1}\, \delta I_{2}$, if the motion is


described

by ordinary polar coordinates. For large values of $n_{1}$~and~$n_{2}$, we must


therefore

expect that the a-priori probability of a stationary state corresponding to a

given combination $(n_{1}, n_{2})$ is proportional to~$n_{2}$. The question of


the a-priori probability

of states corresponding to small values of the~$n$'s has been discussed by

\Name{Sommerfeld} in connection with the problem of the intensities of the


different components

in the fine structure of the hydrogen lines (see Part~II). From considerations

about the volume of the extensions in the phase-space, which might be

considered as associated with the states characterised by different combinations

$(n_{1}, n_{2})$, \Name{Sommerfeld} proposes several different expressions for


the a-priori probability

of such states. Due to the necessary arbitrariness involved in the

137
choice of these extensions, however, we cannot in this way obtain a rational

determination of the a-priori probability of states corresponding to small values of

\PageSep{27}

$n_{1}$~and~$n_{2}$. On the other hand, this probability may be deduced by


regarding the

motion of the system under consideration as the degeneration of a motion


characterised

by three numbers $n_{1}$,~$n_{2}$ and~$n_{3}$, as in the general applications


of the conditions~\Eq{(22)}

to a system of three degrees of freedom. Such a motion may be obtained

for instance by imagining the system placed in a small homogeneous magnetic

field. In certain respects this case falls outside the general theory of conditionally

periodic systems discussed in this section, but, as we shall see in Part~II, it can

be simply shown that the presence of the magnetic field imposes the further
condition

on the motion in the stationary states that the angular momentum round the

axis of the field is equal to~$n'\, \dfrac{h}{2\pi}$, where $n'$~is a positive entire
number equal to or

less than~$n_{2}$, and which for the system considered in the spectral problems
must

be assumed to be different from zero. When regard is taken to the two opposite

directions in which the particle may rotate round the axis of the field, we see

therefore that for this system a state corresponding to a given combination of

$n_{1}$~and~$n_{2}$ in the presence of the field can be established in


$2n_{2}$~different ways.

138
The a-priori probability of the different states of the system may consequently for

all combinations of $n_{1}$~and~$n_{2}$ be assumed to be proportional


to~$n_{2}$.

The\Pagelabel{27} assumption just mentioned that the angular momentum round


the axis

of the field cannot be equal to zero is deduced from considerations of systems for

which the motion corresponding to special combinations of the~$n$'s in~\Eq{(22)}


would

become physically impossible due to some singularity in its character. In such


cases

we must assume that no stationary states exist corresponding to the combinations

$(n_{1}, n_{2}, \MyDots n_{s})$ under consideration, and on the above principle
of the invariance

of the a-priori probability for continuous transformations we shall accordingly

expect that the a-priori probability of any other state, which can be transformed

continuously into one of these states without passing through cases of


degeneration,

will also be equal to zero.

Let us now proceed to consider \Emph{the spectrum of a conditionally periodic

system}, calculated from the values of the energy in the stationary states by means
of

relation~\Eq{(1)}. If $E(n_{1}, \MyDots n_{s})$ is the total energy of a stationary


state determined by~\Eq{(22)}

139
and if $\nu$~is the frequency of the line corresponding to the transition between

two stationary states characterised by $n_{k} = n'_{k}$ and $n_{k} = n''_{k}$


respectively, we have

\[

\nu = \frac{1}{h}\bigl[E(n'_{1}, \MyDots n'_{s}) - E(n''_{1}, \MyDots n''_{s})\


bigr].

\Tag{(26)}

\]

In general, this spectrum will be entirely different from the spectrum to be


expected

on the ordinary theory of electrodynamics from the motion of the system.

Just as for a system of one degree of freedom we shall see, however, that in

the limit where the motions in neighbouring stationary states differ very little from

each other, there exists a close relation between the spectrum calculated on the

quantum theory and that to be expected on ordinary electrodynamics. As in~\


SecRef{2}

\PageSep{28}

we shall further see, that this connection leads to certain general considerations

about the probability of transition between any two stationary states and about

the nature of the accompanying radiation, which are found to be supported by

observations. In order to discuss this question we shall first deduce a general


expression

for the energy difference between two neighbouring states of a conditionally

periodic system, which can be simply obtained by a calculation analogous to that

140
used in~\SecRef{2} in the deduction of the relation~\Eq{(8)}.

Consider some motion of a conditionally periodic system which allows of

separation of variables in a certain set of coordinates $q_{1}$,~\MyDots $q_{s}$,


and let us assume

that at the time $t = \theta$ the configuration of the system will to a close
approximation

be the same as at the time $t = 0$. By taking $\theta$~large enough we can make

this approximation as close as we like. If next we consider some conditionally

periodic motion, obtained by a small variation of the first motion, and which
allows

of separation of variables in a set of coordinates $q'_{1}$,~\MyDots $q'_{s}$


which may differ

slightly from the set $q_{1}$,~\MyDots $q_{s}$, we get by means of \


Name{Hamilton}'s equations~\Eq{(4)}, using

the coordinates $q'_{1}$,~\MyDots $q'_{s}$,

\[

\BreakMath{%

\int_{0}^{\theta} \delta E\, dt

= \int_{0}^{\theta} \sum_{1}^{s} \left(

\frac{\dd E}{\dd p'_{k}}\, \delta p'_{k} + \frac{\dd E}{\dd q'_{k}}\, \delta


q'_{k}

\right) dt

= \int_{0}^{\theta} \sum_{1}^{s} (\dot{q}'_{k}\, \delta p'_{k} - \dot{p}'_{k}\, \


delta q'_{k})\, dt.

141
}{%

\begin{aligned}

\int_{0}^{\theta} \delta E\, dt

&= \int_{0}^{\theta} \sum_{1}^{s} \left(

\frac{\dd E}{\dd p'_{k}}\, \delta p'_{k} + \frac{\dd E}{\dd q'_{k}}\, \delta


q'_{k}

\right) dt \\

&= \int_{0}^{\theta} \sum_{1}^{s} (\dot{q}'_{k}\, \delta p'_{k} - \


dot{p}'_{k}\, \delta q'_{k})\, dt.

\end{aligned}

\]

By partial integration of the second term in the bracket this gives:

\[

\int_{0}^{\theta} \delta E\, dt

= \int_{0}^{\theta} \sum_{1}^{s} \delta(p'_{k}\dot{q}'_{k})\, dt

- \left\lvert\sum_{1}^{s} p'_{k}\, \delta q'_{k}\right\rvert_{t=0}^{t=\theta}.

\Tag{(27)}

\]

Now we have for the unvaried motion

\[

\int_{0}^{\theta} \sum_{1}^{s} p'_{k} \dot{q}'_{k}\, dt

= \int_{0}^{\theta} \sum_{1}^{s} p_{k} \dot{q}_{k}\, dt

142
= \sum_{1}^{s} N_{k} I_{k},

\]

where $I_{k}$~is defined by~\Eq{(21)} and where $N_{k}$~is the number of


oscillations performed

by~$q_{k}$ in the time interval~$\theta$. For the varied motion we have on the
other hand:

\[

\int_{0}^{\theta} \sum_{1}^{s} p'_{k} \dot{q}'_{k}\, dt

= \int_{t=0}^{t=\theta} \sum_{1}^{s} p'_{k} dq'_{k}

= \sum_{1}^{s} N_{k} I'_{k}

+ \left\lvert \sum_{1}^{s} p'_{k}\, \delta q'_{k}\right\rvert_{t=0}^{t=\theta}\!\!,

\]

where the~$I$'s correspond to the conditionally periodic motion in the coordinates

$q'_{1}$,~\MyDots $q'_{s}$, and the~$\delta q$'s which enter in the last term are
the same as those in~\Eq{(27)}.

Writing $I'_{k} - I_{k} = \delta I_{k}$, we get therefore from the latter equation

\[

\int_{0}^{\theta} \delta E\, dt

= \sum_{1}^{s} N_{k}\, \delta I_{k}.

\Tag{(28)}

\]

\PageSep{29}

In the special case where the varied motion is an undisturbed motion belonging

143
to the same system as the unvaried motion we get, since $\delta E$~will be
constant,

\[

\delta E = \sum_{1}^{s} \omega_{k}\, \delta I_{k},

\Tag{(29)}

\]

where $\omega_{k} = \dfrac{N_{k}}{\theta}$ is the mean frequency of oscillation


of~$q_{k}$ between its limits, taken

over a long time interval of the same order of magnitude as~$\theta$. This equation

forms a simple generalisation of~\Eq{(8)}, and in the general case in which a


separation

of variables will be possible only for one system of coordinates leading to a


complete

definition of the~$I$'s it might have been deduced directly from the analytical

theory of the periodicity properties of the motion of a conditionally periodic

system, based on the introduction of angle-variables.\Pagelabel{29}\footnote

{See \Name{Charlier}, Die Mechanik des Himmels, Bd.~I Abt.~2, and especially
\Name{P.~Epstein}, Ann.\ d.\

Phys.\ LI p.~178 (1916). By means of the well known theorem of \Name{Jacobi}


about the change of variables in

the canonical equations of \Name{Hamilton}, the connection between the notion


of angle-variables and the

quantities~$I$, discussed by \Name{Epstein} in the latter paper, may be briefly


exposed in the following elegant

manner which has been kindly pointed out to me by Mr.\ \Name{H.~A.


Kramers}. Consider the function

144
$S(q_{1}, \MyDots q_{s}, I_{1}, \MyDots I_{s})$ obtained from~\Eq{(20)} by
introducing for the~$\alpha$'s their expressions in terms of the~$I$'s

given by the equations~\Eq{(21)}. This function will be a many valued function


of the~$q$'s which increases

by~$I_{k}$ if $q_{k}$~describes one oscillation between its limits and comes


back to its original value

while the other~$q$'s remain constant. If we therefore introduce a new set of


variables $w_{1}, \MyDots w_{s}$

defined by

\[

w_{k} = \frac{\dd S}{\dd I_{k}},\qquad

(k = 1, \MyDots s)

\Tag[I.]{(1^{*})}

\]

it will be seen that $w_{k}$~increases by one unit while the other~$w$'s will
come back to their original

values if $q_{k}$~describes one oscillation between its limits and the other~$q$'s
remain constant. Inversely

it will therefore be seen that the~$q$'s, and also the $p$'s which were given by

\[

p_{k} = \frac{\dd S}{\dd q_{k}},\qquad

(k = i, \MyDots s)\Chg{,}{}

\Tag[I.]{(2^{*})}

\]

145
when considered as functions of the $I$'s and~$w$'s will be periodic functions of
every of the~$w$'s with

period~$1$. According to \Typo{Fourier's}{\Name{Fourier}'s} theorem any of


the~$q$'s may therefore be represented by an $s$-double

trigonometric series of the form

\[

q = \sum A_{\tau_{1}, \MyDots \tau_{s}} \cos 2\pi(\tau_{1} w_{1} + \dots \


Add{+} \tau_{s} w_{s} + \alpha_{\tau_{1}, \MyDots \tau_{s}}),

\Tag[I.]{(3^{*})}

\]

where the $A$'s and~$\alpha$'s are constants depending on the~$I$'s and where
the summation is to be extended

over all entire values of $\tau_{1}$,~\MyDots $\tau_{s}$. On account of this


property of the~$w$'s, the quantities $2\pi w_{1}$,~\MyDots

$2\pi w_{s}$ are denoted as ``angle variables''. Now from \Eq[I.]{(1^{*})}~and \


Eq[I.]{(2^{*})} it follows according to the above mentioned

theorem of \Name{Jacobi} (see for instance \Name{Jacobi}, Vorlesungen über


Dynamik §\;37) that the variations

with the time of the $I$'s and~$w$'s will be given by

\[

\frac{dI_{k}}{dt} = -\frac{\dd E}{\dd w_{k}},\quad

\frac{dw_{k}}{dt} = \frac{\dd E}{\dd I_{k}},\qquad

(k = 1, \MyDots s)

\Tag[I.]{(4^{*})}

\]

146
where the energy~$E$ is considered as a function of the $I$'s and~$w$'s.
Since~$E$, however, is determined

by the $I$'s~only we get from~\Eq[I.]{(4^{*})}, besides the evident result that


the~$I$'s are constant during the motion,

that the~$w$'s will vary linearly with the time and can be represented by

\[

w_{k} = \omega_{k} t + \delta_{k},\quad

\omega_{k} = \frac{\dd E}{\dd I_{k}},\qquad

(k = 1, \MyDots s)

\Tag[I.]{(5^{*})}

\]

where $\delta_{k}$~is a constant, and where $\omega_{k}$~is easily seen to be


equal to the mean frequency of oscillation

of~$q_{k}$. From~\Eq[I.]{(5^{*})} equation~\Eq{(28)} follows at once, and it


will further be seen that by introducing~\Eq[I.]{(5^{*})} in~\Eq[I.]{(3^{*})}

we get the result that every of the~$q$'s, and consequently also any one-valued
function of the~$q$'s,

can be represented by an expression of the type~\Eq{(31)}.

In this connection it may be mentioned that the method of \Name{Schwarzschild}


of fixing the stationary

states of a conditionally periodic system, mentioned on \PageRef{21}, consists in


seeking for a given system

a set of canonically conjugated variables $Q_{1}$,~\MyDots $Q_{s}$,


$P_{1}$,~\MyDots $P_{s}$ in such a way that the positional

147
coordinates of the system $q_{1}$,~\MyDots $q_{s}$, and their conjugated
momenta $p_{1}$,~\MyDots $p_{s}$, when considered as functions

of the $Q$'s~and~$P$'s, are periodic in every of the~$Q$'s with period~$2\pi$,


while the energy of the

system depends only on the~$P$'s. In analogy with the condition which fixes the
angular momentum in

\Name{Sommerfeld}'s theory of central systems \Name{Schwarzschild} next


puts every of the~$P$'s equal to an entire

\Chg{multiplum}{multiple} of~$\dfrac{h}{2\pi}$. In contrast to the theory of


stationary states of conditionally periodic systems based

on the possibility of separation of variables and the fixation of the~$I$'s by~\


Eq{(22)}, this method does not

lead to an absolute fixation of the stationary states, because, as pointed out by \


Name{Schwarzschild} himself,

the above definition of the~$P$'s leaves an arbitrary constant undetermined in


every of these quantities.

In many cases, however, these constants may be simply determined from


considerations of mechanical

transformability of the stationary states, and as pointed out by \Name{Burgers}


(loc.\ cit.\ Versl.\ Akad.\ Amsterdam

XXV p.~1055 (1917)\Typo{}{).} \Name{Schwarzschild}'s method possesses on


the other hand the essential advantage of

being applicable to certain classes of systems in which the displacements of the


particles may be

represented by trigonometric series of the type~\Eq{(31)}, but for which the


equations of motion cannot be

solved by separation of variables in any fixed set of coordinates. An interesting


application of this to
148
the spectrum of rotating molecules, given by \Name{Burgers}, will be mentioned
in Part~IV\@.}

From~\Eq{(29)} it follows moreover

\PageSep{30}

that, if the system allows of a separation of variables in an infinite continuous

multitude of sets of coordinates, the total energy will be the same for all motions

corresponding to the same values of the~$I$'s, independent of the special set of

coordinates used to calculate these quantities. As mentioned above and as we have

already shown in the case of purely periodic systems by means of~\Eq{(8)}, the
total

energy is therefore also in cases of degeneration completely determined by the

conditions~\Eq{(22)}.

Consider now a transition between two stationary states determined by~\Eq{(22)}

by putting $n_{k} = n'_{k}$ and $n_{k} = n''_{k}$ respectively, and let us assume
that $n'_{1}$,~\MyDots $n'_{s}$,

$n''_{1}$,~\MyDots $n''_{s}$ are large numbers, and that the differences $n'_{k} -
n''_{k}$ are small compared

with these numbers. Since the motions of the system in these states will differ

relatively very little from each other we may calculate the difference of the energy

by means of~\Eq{(29)}, and we get therefore, by means of~\Eq{(1)}, for the


frequency of the

radiation corresponding to the transition between the two states

\[

149
\nu = \frac{1}{h} (E' - E'')

= \frac{1}{h} \sum_{1}^{s} \omega_{k} (I'_{k} - I''_{k})

= \sum_{1}^{s} \omega_{k} (n'_{k} - n''_{k}),

\Tag{(30)}

\]

which is seen to be a direct generalisation of the expression~\Eq{(13)} in~\


SecRef{2}.

Now, in complete analogy to what is the case for periodic systems of one degree

of freedom, it is proved in the analytical theory of the motion of conditionally

periodic systems mentioned above that for the latter systems the coordinates

$q_{1}$,~\MyDots $q_{s}$, and consequently also the displacements of the


particles in any given

direction, may be expressed as a function of the time by an $s$-double infinite

\Name{Fourier} series of the form:

\PageSep{31}

\[

\xi = \sum C_{\tau_{1}, \MyDots \tau_{s}}

\cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \dots \Add{+} \tau_{s} \omega_{s})t +


c_{\tau_{1}, \MyDots \tau_{s}}\bigr\},

\Tag{(31)}

\]

where the summation is to be extended over all positive and negative entire values

150
of the~$\tau$'s, and where the~$\omega$'s are the above mentioned mean
frequencies of oscillation

for the different~$q$'s. The constants $C_{\tau_{1}, \MyDots \tau_{s}}$ depend


only on the~$\alpha$'s in the equations~\Eq{(18)}

or, what is the same, on the~$I$'s, while the constants $c_{\tau_{1}, \MyDots \
tau_{s}}$ depend on

the~$\alpha$'s as well as on the~$\beta$'s. In general the quantities $\tau_{1} \


omega_{1} + \dots \Add{+} \tau_{s} \omega_{s}$ will be

different for any two different sets of values for the~$\tau$'s, and in the course of
time

the orbit will cover everywhere dense a certain $s$-dimensional extension. In a

case of degeneration, however, where the orbit will be confined to an extension of

less dimensions, there will exist for all values of the~$\alpha$'s one or more
relations of the

type $m_{1} \omega_{1} + \dots \Add{+} m_{s} \omega_{s} = 0$ where


the~$m$'s are entire numbers and by the introduction

of which the expression~\Eq{(31)} can be reduced to a \Name{Fourier} series


which is

less than $s$-double infinite. Thus in the special case of a system of which every

orbit is periodic we have $\dfrac{\omega_{1}}{\kappa_{1}} = \dots = \dfrac{\


omega_{s}}{\kappa_{s}} = \omega$, where the~$\kappa$'s are the numbers

which enter in equation~\Eq{(23)}, and the \Name{Fourier} series for the


displacements in

the different directions will in this case consist only of terms of the simple form

$C_{\tau} \cos 2\pi \{\tau \omega t + c_{\tau}\}$, just as for a system of one
degree of freedom.

151
On the ordinary theory of radiation, we should expect from~\Eq{(31)} that the

spectrum emitted by the system in a given state would consist of an $s$-double


infinite

series of lines of frequencies equal to $\tau_{1} \omega_{1} + \dots + \tau_{s} \


omega_{s}$. In general, this

spectrum would be completely different from that given by~\Eq{(26)}. This


follows already

from the fact that the~$\omega$'s will depend on the values for the constants $\
alpha_{1}$,~\MyDots $\alpha_{s}$

and will vary in a continuous way for the continuous multitude of mechanically

possible states corresponding to different sets of values for these constants. Thus

in general the~$\omega$'s will be quite different for two different stationary states

corresponding to different sets of~$n$'s in~\Eq{(22)}, and we cannot expect any


close relation

between the spectrum calculated on the quantum theory and that to be expected

on the ordinary theory of mechanics and electrodynamics. In the limit,

however, where the~$n$'s in~\Eq{(22)} are large numbers, the ratio between
the~$\omega$'s for

two stationary states, corresponding to $n_{k} = n'_{k}$ and $n_{k} = n''_{k}$


respectively, will tend

to unity if the differences $n'_{k} - n''_{k}$ are small compared with the~$n$'s,
and as seen

from~\Eq{(30)} the spectrum calculated by \Eq{(1)}~and \Eq{(22)} will in this


limit just tend to

coincide with that to be expected on the ordinary theory of radiation from the

motion of the system.

152
As far as the frequencies are concerned, we thus see that for conditionally

periodic systems there exists a connection between the quantum theory and the

ordinary theory of radiation of exactly the same character as that shown in~\
SecRef{2}

to exist in the simple case of periodic systems of one degree of freedom. Now on

ordinary electrodynamics the coefficients $C_{\tau_{1}\Add{,} \dots, \tau_{s}}$


in the expression~\Eq{(31)} for the

displacements of the particles in the different directions would in the well known

way determine the intensity and polarisation of the emitted radiation of the

corresponding frequency $\tau_{1} \omega_{1} + \dots \Add{+} \tau_{s} \


omega_{s}$. As for systems of one degree of freedom

\PageSep{32}

we must therefore conclude that, in the limit of large values for the~$n$'s, the
probability

of spontaneous transition between two stationary states of a conditionally

periodic system, as well as the polarisation of the accompanying radiation, can be

determined directly from the values of the coefficient $C_{\tau_{1}, \MyDots \


tau_{s}}$ in~\Eq{(31)} corresponding

to a set of~$\tau$'s given by $\tau_{k} = n'_{k} - n''_{k}$, if $n'_{1}$,~\MyDots


$n'_{s}$ and $n''_{1}$,~\MyDots $n''_{s}$ are the numbers which

characterise the two stationary states.

Without\Pagelabel{32} a detailed theory of the mechanism of transition between


the stationary

153
states we cannot, of course, in general obtain an exact determination of the \
Emph{probability

of spontaneous transition} between two such states, unless the~$n$'s are

large numbers. Just as in the case of systems of one degree of freedom, however,

we are naturally led from the above considerations to assume that, also for

values of the~$n$'s which are not large, there must exist an intimate connection

between the probability of a given transition and the values of the corresponding

\Name{Fourier} coefficient in the expressions for the displacements of the


particles in the

two stationary states. This allows us at once to draw certain important conclusions.

Thus, from the fact that in general negative as well as positive values for the~$\
tau$'s

appear in~\Eq{(31)}, it follows that we must expect that in general not only such

transitions will be possible in which all the~$n$'s decrease, but that also transitions

will be possible for which some of the~$n$'s increase while others decrease. This

conclusion, which is supported by observations on the fine structure of the


hydrogen

lines as well as on the \Name{Stark} effect, is contrary to the suggestion, put


forward

by \Name{Sommerfeld} with reference to the essential positive character of


the~$I$'s, that

every of the~$n$'s must remain constant or decrease under a transition. Another

direct consequence of the above considerations is obtained if we consider a system

for which, for all values of the constants $\alpha_{1}$,~\MyDots $\alpha_{s}$, the
coefficient $C_{\tau_{1}, \MyDots \tau_{s}}$ corresponding

154
to a certain set $\tau_{1}^{0}$,~\MyDots $\tau_{s}^{0}$ of values for the~$\
tau$'s is equal to zero in the expressions

for the displacements of the particles in every direction. In this case we

shall naturally expect that no transition will be possible for which the relation

$n'_{k} - n''_{k} = \tau_{k}^{0}$ is satisfied for every~$k$. In the case where


$C_{\tau_{1}^{0}, \MyDots \tau_{s}^{0}}$ is equal to zero

in the expressions for the displacement in a certain direction only, we shall expect

that all transitions, for which $n'_{k} - n''_{k} = \tau_{k}^{0}$ for every~$k$,
will be accompanied by

a radiation which is polarised in a plane perpendicular to this direction.

A simple illustration of the last considerations is afforded by the system mentioned

in the beginning of this section, and which consists of a particle executing

motions in three perpendicular directions which are independent of each other. In

that case all the \Name{Fourier} coefficients in the expressions for the
displacements

in any direction will disappear if more than one of the~$\tau$'s are different from

zero. Consequently we must assume that only such transitions are possible for

which only one of the~$n$'s varies at the same time, and that the radiation
corresponding

to such a transition will be linearly polarised in the direction of the displacement

of the corresponding coordinate. In the special case where the motions

in the three directions are simply harmonic, we shall moreover conclude that none

\PageSep{33}

155
of the~$n$'s can vary by more than a single unit, in analogy with the
considerations

in the former section about a linear harmonic vibrator.

Another\Pagelabel{33} example which has more direct physical importance, since


it includes

all the special applications of the quantum theory to spectral problems mentioned

in the introduction, is formed by a conditionally periodic system possessing an axis

of symmetry. In all these applications a separation of variables is obtained in a

set of three coordinates $q_{1}$,~$q_{2}$ and~$q_{3}$, of which the first two


serve to fix the position

of the particle in a plane through the axis of the system, while the last is

equal to the angular distance between this plane and a fixed plane through the

same axis. Due to the symmetry, the expression for the total energy in \
Name{Hamilton}'s

equations will not contain the angular distance~$q_{3}$ but only the angular
momentum~$p_{3}$

round the axis. The latter quantity will consequently remain constant during

the motion, and the variations of $q_{1}$~and~$q_{2}$ will be exactly the same
as in a conditionally

periodic system of two degrees of freedom only. If the position of the

particle is described in a set of cylindrical coordinates $z$,~$\rho$,~$\theta$,


where $z$~is the displacement

in the direction of the axis, $\rho$~the distance of the particle from this axis

and $\theta$~is equal to the angular distance~$q_{3}$, we have therefore

156
\[

\begin{aligned}

z &= \sum C_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2}) t + c_{\tau_{1}, \tau_{2}}\bigr\} \\

\llap{\text{and}\quad}

\rho &= \sum C'_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2}) t + c'_{\tau_{1}, \tau_{2}}\bigr\},

\end{aligned}

\Tag{(32)}

\]

where the summation is to be extended over all positive and negative entire values

of $\tau_{1}$~and~$\tau_{2}$, and where $\omega_{1}$~and~$\omega_{2}$ are


the mean frequencies of oscillation of the

coordinates $q_{1}$~and~$q_{2}$. For the rate of variation of~$\theta$ with the


time we have further

\[

\BreakMath{%

\frac{d\theta}{dt} = \dot{q}_{3}

= \frac{\dd E}{\dd p_{3}}

= f(q_{1}, q_{2}, p_{1}, p_{2}, p_{3})

= ± \sum C''_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2}) t + c''_{\tau_{1}, \tau_{2}}\bigr\},

}{%

\begin{aligned}

157
\frac{d\theta}{dt} = \dot{q}_{3}

&= \frac{\dd E}{\dd p_{3}}

= f(q_{1}, q_{2}, p_{1}, p_{2}, p_{3}) \\

&= ± \sum C''_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2}) t + c''_{\tau_{1}, \tau_{2}}\bigr\},

\end{aligned}

\]

where the two signs correspond to a rotation of the particle in the direction of
increasing

and decreasing~$q_{3}$ respectively, and are introduced to separate the two

types of symmetrical motions corresponding to these directions. This gives

\[

± \theta = 2\pi \omega_{3} t + \sum C'''_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\


tau_{1} \omega_{1} + \tau_{2} \omega_{2}) t + c'''_{\tau_{1}, \tau_{2}}\bigr\},

\Tag{(33)}

\]

where the positive constant $\omega_{3} = \dfrac{1}{2\pi} C''_{0,0}$ is the mean


frequency of rotation round the

axis of symmetry of the system. Considering now the displacement of the particle

in rectangular coordinates $x$,~$y$ and~$z$, and taking as above the axis of


symmetry

as $z$-axis, we get from \Eq{(32)}~and~\Eq{(33)} after a simple contraction of


terms

\[

158
\BreakMath{%

\begin{alignedat}{3}

x &= \rho \cos \theta

&&= &&\sum D_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2} + \omega_{3}) t + d_{\tau_{1}, \tau_{2}}\bigr\} \\

\llap{\text{and}}\

y &= \rho \sin \theta

&&= ± &&\sum D_{\tau_{1}, \tau_{2}} \sin 2\pi \bigl\{(\tau_{1} \omega_{1}


+ \tau_{2} \omega_{2} + \omega_{3}) t + d_{\tau_{1}, \tau_{2}}\bigr\},

\end{alignedat}

}{%

\begin{aligned}

x &= \rho \cos \theta \\

&= \sum D_{\tau_{1}, \tau_{2}} \cos 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2} + \omega_{3}) t + d_{\tau_{1}, \tau_{2}}\bigr\} \\

\llap{\text{and}}\

y &= \rho \sin \theta \\

&= ± \sum D_{\tau_{1}, \tau_{2}} \sin 2\pi \bigl\{(\tau_{1} \omega_{1} + \


tau_{2} \omega_{2} + \omega_{3}) t \rlap{$+ d_{\tau_{1}, \tau_{2}}\bigr\},$}

\end{aligned}

\Tag{(34)}

\]

159
where the $D$'s and~$d$'s are new constants, and the summation is again to be
extended

over all positive and negative values of $\tau_{1}$~and~$\tau_{2}$.

%[** TN: Printer's mark]

%D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8. Række, IV. 1. 5

\PageSep{34}

From \Eq{(32)} and~\Eq{(34)} we see that the motion in the present case may be
considered

as composed of a number of linear harmonic vibrations parallel to the axis

of symmetry and of frequencies equal to the absolute values of $(\tau_{1} \


omega_{1} + \tau_{2} \omega_{2})$,

together with a number of circular harmonic motions round this axis of frequencies

equal to the absolute values of $(\tau_{1} \omega_{1} + \tau_{2} \omega_{2} + \


omega_{3})$, and possessing the same direction

of rotation as that of the moving particle or the opposite if the latter expression

is positive or negative respectively. According to ordinary electrodynamics

the radiation from the system would therefore consist of a number of components

of frequency $\tau_{1} \omega_{1} + \tau_{2} \omega_{2}$ polarised parallel to


the axis of symmetry, and a number

of components of frequencies $\tau_{1} \omega_{1} + \tau_{2} \omega_{2} + \


omega_{3}$ and of circular polarisation round

this axis (when viewed in the direction of the axis). On the present theory we

shall consequently expect that in this case only two kinds of transitions between

160
the stationary states given by~\Eq{(22)} will be possible. In both of these
$n_{1}$~and~$n_{2}$

may vary by an arbitrary number of units, but in the first kind of transition, which

will give rise to a radiation polarised parallel to the axis of the system,
$n_{3}$~will

remain unchanged, while in the second kind of transition $n_{3}$~will decrease or


increase

by one unit and the emitted radiation will be circularly polarised round

the axis in the same direction as or the opposite of that of the rotation of the
particle

respectively.

In the next Part we shall see that these conclusions are supported in an

instructive manner by the experiments on the effects of electric and magnetic fields

on the hydrogen spectrum. In connection with the discussion of the general theory,

however, it may be of interest to show that the formal analogy between the
ordinary

theory of radiation and the theory based on \Eq{(1)}~and \Eq{(22)}, in case of


systems possessing

an axis of symmetry, can be traced not only with respect to frequency

relations but also by considerations of \Emph{conservation of angular


momentum}.\Pagelabel{34}

For a conditionally periodic system possessing an axis of symmetry the angular

momentum round this axis is, with the above choice of coordinates, according to~\
Eq{(22)}

161
equal to $\dfrac{I_{3}}{2\pi} = n_{3}\, \dfrac{h}{2\pi}$. If therefore, as assumed
above for a transition corresponding

to an emission of linearly polarised light, $n_{3}$~is unaltered, it means that

the angular momentum of the system remains unchanged, while if $n_{3}$~alters


by

one unit, as assumed for a transition corresponding to an emission of circularly

polarised light, the angular momentum will be altered by~$\dfrac{h}{2\pi}$. Now


it is easily

seen that the ratio between this amount of angular momentum and the amount of

energy~$h\nu$ emitted during the transition is just equal to the ratio between the
amount

of angular momentum and energy possessed by the radiation which according to

ordinary electrodynamics would be emitted by an electron rotating in a circular

orbit in a central field of force. In fact, if $a$~is the radius of the orbit, $\nu$~the
frequency

of revolution and $F$~the force of reaction due to the electromagnetic field

of the radiation, the amount of energy and of angular momentum round an

axis through the centre of the field perpendicular to the plane of the orbit, lost

\PageSep{35}

by the electron in unit of time as a consequence of the radiation, would be equal to

$2\pi\nu aF$ and $aF$~respectively. Due to the principles of conservation of


energy and of

angular momentum holding in ordinary electrodynamics, we should therefore


expect

162
that the ratio between the energy and the angular momentum of the emitted
radiation

would be~$2\pi\nu$,\footnote

{Comp.\ \Name{K.~Schaposchnikow}, Phys.\ Zeitschr.\ XV, p.~454 (1914).}

but this is seen to be equal to the ratio between the energy~$h\nu$ and

the angular momentum~$\dfrac{h}{2\pi}$ lost by the system considered above


during a transition

for which we have assumed that the radiation is circularly polarised. This
agreement

would seem not only to support the validity of the above considerations but

also to offer a direct support, independent of the equations~\Eq{(22)}, of the


assumption

that, \Emph{for an atomic system possessing an axis of symmetry, the total

angular momentum round this axis is equal to an entire multiple of~$\dfrac{h}{2\


pi}$}.

A further illustration of the above considerations of the relation between

the quantum theory and the ordinary theory of radiation is obtained if we consider

a conditionally periodic system subject to the \Emph{influence of a small


perturbing

field of force}. Let us assume that the original system allows of separation of

variables in a certain set of coordinates $q_{1}$,~\MyDots $q_{s}$, so that the


stationary states are

determined by~\Eq{(22)}. From the necessary stability of the stationary states we


must

163
conclude that the perturbed system will possess a set of stationary states which

only differ slightly from those of the original system. In general, however, it will

not be possible for the perturbed system to obtain a separation of variables in any

set of coordinates, but if the perturbing force is sufficiently small the perturbed

motion will again be of conditionally periodic type and may be regarded as a


superposition

of a number of harmonic vibrations just as the original motion. The displacements

of the particles in the stationary states of the perturbed system will

therefore be given by an expression of the same type as~\Eq{(31)} where the


fundamental

frequencies~$\omega_{k}$ and the amplitudes $C_{\tau_{1}, \MyDots \tau_{s}}$


may differ from those corresponding to

the stationary states of the original system by small quantities proportional to the

intensity of the perturbing forces. If now for the original motion the coefficients

$C_{\tau_{1}, \MyDots \tau_{s}}$ corresponding to certain combinations of


the~$\tau$'s are equal to zero for all

values of the constants $\alpha_{1}$,~\MyDots $\alpha_{s}$, these coefficients


will therefore for the perturbed

motion, in general, possess small values proportional to the perturbing forces.


From

the above considerations we shall therefore expect that, in addition to the main

probabilities of such transitions between stationary states which are possible for the

original system, there will for the perturbed system exist small probabilities of new

transitions corresponding to the above mentioned combinations of the~$\tau$'s.


Consequently

164
we shall expect that the effect of the perturbing field on the spectrum

of the system will consist partly in a small displacement of the original lines\
Add{,}

partly in the appearance of new lines of small intensity.

A simple example of this is afforded by a system consisting of a particle moving

in a plane and executing harmonic vibrations in two perpendicular directions with


frequencies

\PageSep{36}

$\omega_{1}$~and~$\omega_{2}$. If the system is undisturbed all coefficients


$C_{\tau_{1}, \tau_{2}}$ will be zero,

except $C_{1,0}$ and~$C_{0, 1}$. When, however, the system is perturbed, for
instance by an

arbitrary small central force, there will in the \Name{Fourier} expressions for the
displacements

of the particle, in addition to the main terms corresponding to the

fundamental frequencies $\omega_{1}$~and~$\omega_{2}$, appear a number of


small terms corresponding

to frequencies given by $\tau_{1} \omega_{1} + \tau_{2} \omega_{2}$ where $\


tau_{1}$~and~$\tau_{2}$ are entire numbers which may

be positive as well as negative. On the present theory we shall therefore expect

that in the presence of the perturbing force there will appear small probabilities

for new transitions which will give rise to radiations analogous to the so\Typo{}
{ }called

harmonics and combination tones in acoustics, just as it should be expected on the

165
ordinary theory of radiation where a direct connection between the emitted
radiation

and the motion of the system is assumed. Another example of more direct

physical application is afforded by the effect of an external homogeneous electric

field in producing new spectral lines. In this case the potential of the perturbing

force is a linear function of the coordinates of the particles and, whatever is the

nature of the original system, it follows directly from the general theory of
perturbations

that the frequency of any additional term in the expression for the perturbed

motion, which is of the same order of magnitude as the external force, must
correspond

to the sum or difference of two frequencies of the harmonic vibrations into

which the original motion can be resolved. With applications of these


considerations

we will meet in Part~II in connection with the discussion of \Name{Sommerfeld}'s


theory

of the fine structure of the hydrogen lines and in Part~III in connection with the

problem of the appearance of new series in the spectra of other elements under the

influence of intense external electric fields.

As mentioned we cannot without a more detailed theory of the mechanism of


transition

between stationary states obtain quantitative information as regards the general

question of the intensities of the different lines of the spectrum of a conditionally

166
periodic system given by~\Eq{(26)}, except in the limit where the~$n$'s are large
numbers,

or in such special cases where for all values of the constants $\alpha_{1}$,~\
MyDots $\alpha_{s}$ certain

coefficients $C_{\tau_{1}, \MyDots \tau_{s}}$ in~\Eq{(31)} are equal to zero.


From considerations of analogy, however,

we must expect that it will be possible also in the general case to obtain an

\Emph{estimate of the intensities} of the different lines in the spectrum by


comparing

the intensity of a given line, corresponding to a transition between two stationary

states characterised by the numbers $n'_{1}$,~\MyDots $n'_{s}$ and $n''_{1}$,~\


MyDots $n''_{s}$ respectively, with the

intensities of the radiations of frequencies $\omega_{1} (n'_{1} - n''_{1}) + \dots


+ \omega_{s} (n'_{s} - n''_{s})$ to be expected

on ordinary electrodynamics from the motions in these states; although of

course this estimate becomes more uncertain the smaller the values for the~$n$'s
are.

As it will be seen from the applications mentioned in the following Parts this is

supported in a general way by comparison with the observations.

\vfill

\begin{center}

\footnotesize

Færdig fra Trykkeriet d.~27, April 1918.

\end{center}

167
% [** TN: Publisher's titlepage with series header and price]

\iffalse%****

\begin{center}

Mémoires de l'Académie Royale des Sciences et des Lettres de Danemark.


Copenhague. \\

Section des Sciences, 8\ieme~série, t.~IV, \numero~1, fasc.~2.

\Large ON \\

\LARGE THE QUANTUM THEORY \\

OF LINE-SPECTRA \\

\footnotesize

BY \\

\large

N. BOHR \\

\vfill

\textsf{PART II} \\

\vfill

\footnotesize

\textsc{D. Kgl.\ Danske Vidensk.\ Selsk.\ Skripter, naturvidensk.\ og mathem.\


Afd.,~8. Række,~IV. 1, 2.}

\vfill\vfill

168
\large

KØBENHAVN \\

\textsc{hovedkommissionær}: ANDR. FRED. HØST \&~SON, \textsc{kgl., hof-


boghandel} \\

BIANCO LUNOS BOGTRYKKERI \\

1918

Pris: 4~Kr.

\end{center}

%\cleardoublepage

\fi % [** TN: End of publisher's titlepage]

\Part{II.}{On the hydrogen spectrum.}

\Section{1.}{The simple theory of the series~spectrum~of~hydrogen.}

As well known, the frequencies of the lines of the series spectrum of hydrogen

may, if we look apart from the fine structure of the single lines revealed by
instruments

of high dispersive power, be represented by the formula

\[

\nu = K\left(\frac{1}{n''^{2}} - \frac{1}{n'^{2}}\right),

169
\Tag{(35)}

\]

where $K$~is a constant, and $n'$~and $n''$ a set of two entire numbers, different
for the

different lines of the spectrum. According to the general principles of the quantum

theory of line spectra discussed in the first section of Part~I, we shall therefore

expect that this spectrum is emitted by a system which possesses a series of


stationary

states in which the numerical value of the energy in the \Ord{$n$}{th}~state,


omitting

an arbitrary constant, with a high degree of approximation is given by

\[

E_{n} = \frac{Kh}{n^{2}},

\Tag{(36)}

\]

where $h$~is \Name{Planck}'s constant which enters in the fundamental relation~\


EqI{(1)}.

Now according to \Name{Rutherford}'s theory of atomic structure, a neutral


hydrogen

atom must be expected to consist of an electron and a positive nucleus of a mass

very large compared with that of the electron, which move under the influence of

a mutual attraction inversely proportional to the square of the distance apart.


Assuming

that the motion in the stationary states may be determined by ordinary

170
mechanics, and neglecting for the moment the small modifications claimed by the

theory of relativity, we find that each of the particles will describe an elliptical

orbit with their common centre of gravity at one of the foci, and from the well

known laws for a Keplerian motion we have that the frequency of revolution~$\
omega$

and the major axis~$2\alpha$ of the relative orbit of the particles, quite
independent of

the degree of eccentricity of this orbit, are given by

\[

\omega = \sqrt{\frac{2W^{3} (M + m)}{\pi^{2} N^{2} e^{4} Mm}},\quad

2\alpha = \frac{Ne^{2}}{W},

\Tag{(37)}

\]

where $W$~is the work necessary to remove the electron to infinite distance from
the

nucleus, while $Ne$~and $M$ are the charge and the mass of the nucleus, and $-
e$~and

$m$ the charge and the mass of the electron.

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8. Raekke, IV. 1.


6

\PageSep{38}

As explained in Part~I, there will in general be no simple connection between

171
the motion of a system in the stationary states and the spectrum emitted during

transitions between these states; such a connection, however, must be expected to

exist in the limit where the motions in successive stationary states differ
comparatively

little from each other. In the present case this connection claims in the first

place that the frequency of revolution tends to zero for increasing~$n$. According
to

\Eq{(36)}~and \Eq{(37)} we may therefore put the value of~$W$ in the \


Ord{$n$}{th}~stationary state equal to

\[

W_{n} = \frac{Kh}{n^{2}}.

\Tag{(38)}

\]

Moreover, since \Eq{(35)}~can be written in the form

\[

\nu = (n' - n'') K\, \frac{n' + n''}{n'^{2} n''^{2}},

\]

it is seen to be a necessary condition that the frequency of revolution for large

values of~$n$ is asymptotically given by

\[

\omega_{n} \sim \frac{2K}{n^{3}},

\Tag{(39)}

\]

if we wish that the frequency of the radiation emitted during a transition between

172
two stationary states, for which the numbers $n'$~and $n''$ are large compared
with

their difference $n' - n''$, shall tend to coincide with one of the frequencies of the

spectrum which on ordinary electrodynamics would be emitted from the system in

these states. But from \Eq{(37)}~and \Eq{(38)} it will be seen that \


Eq{(39)}~claims the fulfilment

of the relation

\[

K = \frac{2\pi^{2} N^{2} e^{4} Mm}{h^{3} (M + m)}

= \frac{2\pi^{2} N^{2} e^{4} m}{h^{3} (1 + m/M)}.

\Tag{(40)}

\]

As shown in previous papers, this relation is actually found to be fulfilled

within the limit of experimental errors if we put $N = 1$ and for $e$,~$m$,


and~$h$ introduce

the values deduced from measurements on other phenomena; a result

which may be considered as affording a strong support for the validity of the

general principles discussed in Part~I, as well as for the reality of the atomic model

under consideration. Further it was found that, if in formula~\Eq{(35)} for the


hydrogen

spectrum the constant~$K$ is replaced by a constant which is four times larger,


this

formula represents to a high degree of approximation the frequencies of the lines of


a

173
spectrum emitted by helium, when this gas is subject to a condensed discharge.
This

was to be expected on \Name{Rutherford}'s theory, according to which a neutral


helium

atom contains two electrons and a nucleus of a charge twice that of the nucleus

of the hydrogen atom. A helium atom from which one electron is removed will

thus form a dynamical system perfectly similar to a neutral hydrogen atom, and

may therefore be expected to emit a spectrum represented by~\Eq{(35)} if in~\


Eq{(40)} we put

$N = 2$. Moreover a closer comparison of the helium spectrum under


consideration

\PageSep{39}

with the hydrogen spectrum has shown that the value of the constant~$K$ in the

former spectrum was not exactly four times as large as that in the latter, but that

the ratio between these constants within the limit of experimental errors agreed

with the value to be expected from~\Eq{(40)}, when regard is taken to the


different masses

of the nuclei of the atoms of hydrogen and helium corresponding to the different

atomic weights of these elements.\footnote

{For the \Typo{litterature}{literature} on this subject the reader is referred to the


papers cited in the introduction.}

Introducing the expression for~$K$ given by~\Eq{(40)} in the formulæ \Eq{(37)}


and \Eq{(38)},

we find for the values of $W$,~$\omega$ and~$2\alpha$ in the stationary states

174
\[

\BreakMath{%

W_{n} = \frac{1}{n^{2}}\, \frac{2\pi^{2} N^{2} e^{4} Mm}{h^{2} (M + m)},\


quad

\omega_{n} = \frac{1}{n^{3}}\, \frac{4\pi^{2} N^{2} e^{4} Mm}{h^{3} (M +


m)},\quad

2\alpha_{n} = n^{2}\, \frac{h^{2} (M + m)}{2\pi^{2} N e^{2} Mm}.

}{%

\begin{gathered}

W_{n} = \frac{1}{n^{2}}\, \frac{2\pi^{2} N^{2} e^{4} Mm}{h^{2} (M + m)}, \\

\omega_{n} = \frac{1}{n^{3}}\, \frac{4\pi^{2} N^{2} e^{4} Mm}{h^{3} (M +


m)},\quad

2\alpha_{n} = n^{2}\, \frac{h^{2} (M + m)}{2\pi^{2} N e^{2} Mm}.

\end{gathered}

\Tag{(41)}

\]

Now for a mechanical system as that under consideration, for which every motion

is periodic independent of the initial conditions, we have that the value of the total

energy will be completely determined by the value of the quantity~$I$, defined

by equation~\EqI{(5)} in Part~I\@. As mentioned this follows directly from


relation~\EqI{(8)},

which shows at the same time that for a system for which every motion is periodic

the frequency will be completely determined by~$I$ or by the energy only. For the

175
value of~$I$ in the stationary states of the hydrogen atom we get by means of~\
EqI{(8)}

from \Eq{(37)}~and \Eq{(41)}, since in this case $I$~will obviously become zero
when $W$~becomes

infinite,

\[

\BreakMath{%

I = \int_{W_{n}}^{\infty} \frac{dW}{\omega}

= \sqrt{\frac{\pi^{2} N^{2} e^{4} Mm}{2(M + m)}} \int_{W_{n}}^{\infty}


W^{-3/2}\, dW

= \sqrt{\frac{2\pi^{2} N^{2} e^{4} Mm}{W_{n}(M + m)}}

= nh.

}{%

\begin{aligned}

I = \int_{W_{n}}^{\infty} \frac{dW}{\omega}

&= \sqrt{\frac{\pi^{2} N^{2} e^{4} Mm}{2(M + m)}} \int_{W_{n}}^{\infty}


W^{-3/2}\, dW \\

&= \sqrt{\frac{2\pi^{2} N^{2} e^{4} Mm}{W_{n}(M + m)}}

= nh.

\end{aligned}

\]

This result will be seen to be consistent with condition~\EqI{(24)} which, as


mentioned

176
in Part~I, presents itself as a direct generalisation to periodic systems of several

degrees of freedom of condition~\EqI{(10)} which determines the stationary states


of a

system of one degree of freedom, and which again on \Name{Ehrenfest}'s


principle of

the mechanical transformability of the stationary states forms a rational


generalisation

of \Name{Planck}'s fundamental formula~\EqI{(9)} for the possible values of the


energy of a

linear harmonic vibrator.

In this connection it will be observed, that the relation discussed above between

the hydrogen spectrum and the motion of the atom in the limit of small frequencies

is completely analogous to the general relation, discussed in \SecRef[I]{2} in


Part~I,

between the spectrum which on the quantum theory would be emitted by a system

of one degree of freedom, the stationary states of which are determined by~\
EqI{(10)}, and

the motion of the system in these states. It will at the same time be noted that, in

case of hydrogen, this relation implies that the motion of the particles in the

stationary states of the atom will not in general be simply harmonic, or in other

words that the orbit of the electron will not in general be circular. In fact if the

motion of the particles were simply harmonic, as the motion of a \Name{Planck}'s


vibrator,

\PageSep{40}

177
we should expect on the considerations in Part~I that no transition between two

stationary states of the atom would be possible for which $n'$~and~$n''$ differ by
more

than one unit; but this would obviously be inconsistent with the observations,

since for instance the lines of the ordinary Balmer series, according to the theory,

correspond to transitions for which $n'' = 2$ while $n'$~takes the values $3$,~$4$,
$5$,~\dots\Add{.}

In connection with this consideration it may be remarked that, adopting a


terminology

well known from acoustics, we may from the point of view of the quantum

theory regard the higher members of the Balmer series ($n' = 4$, $5$,~\dots) as the

``harmonics'' of the first member ($n' = 3$), although of course the frequencies of

the former lines are by no means entire multipla of the frequency of the latter line.

While in the above way it was possible to obtain a simple interpretation of

certain main features of the hydrogen spectrum, it was not found possible in this
way

to account in detail for such phenomena in which the deviation of the motion of

the particles from a simple Keplerian motion plays an essential part. This is the

case in the problem of the fine structure of the hydrogen lines, which is due to

the effect of the small variation of the mass of the electron with its velocity, as

well as in the problems of the characteristic effects of external electric and


magnetic

fields on the hydrogen lines. As mentioned in the introduction, a progress of

178
fundamental importance in the treatment of such problems was made by \
Name{Sommerfeld},

who obtained a convincing explanation of the fine structure of the hydrogen

lines by means of his theory of the stationary states of central systems, in which

the single condition $I = nh$ was replaced by the two conditions~\EqI{(16)}; and
the

theory was further developed by \Name{Epstein} and \Name{Schwarzschild},


who on this line

established the general theory, based on the conditions~\EqI{(22)}, of the


stationary states

of a conditionally periodic system for which the equations of motion may be

solved by means of separation of variables in the \Typo{Hamilton-Jacobi}{\


Name{Hamilton}-\Name{Jacobi}} partial differential

equation. If the hydrogen atom is exposed to a homogeneous electric or to a

homogeneous magnetic field, the atom forms a system of this class, and, as shown

by \Name{Epstein} and \Name{Schwarzschild} as regards the \Name{Stark}


effect and by \Name{Sommerfeld} and

\Name{Debye} as regards the \Name{Zeeman} effect, the theory under


consideration leads to values

for the total energy of the atom in the stationary states, which together with
relation~\EqI{(1)}

lead again to values for the frequencies of the radiations emitted during

the transitions between these states, which are in agreement with the measured
frequencies

of the components into which the hydrogen lines are split up in the presence

of the fields. As pointed out in Part~I, it is possible moreover to throw light

179
on the question of the intensities and polarisations of these components on the

basis of the necessary formal relation between the quantum theory of line spectra

and the ordinary theory of radiation in the limit where the motions in successive

stationary states differ very little from each other. In the following sections the

mentioned problems will be discussed in detail. As regards the fixation of the

stationary states we shall not, however, follow the same procedure as used by the

authors just mentioned, which rests upon the immediate application of the
conditions~\EqI{(22)},

\PageSep{41}

but it will be shown how the conditions which fix the stationary states of

the perturbed atom may be obtained by a direct examination of the small deviations

of the motion of the electron from a simple Keplerian motion. In this way it

seems possible to obtain a more direct illustration of the principles discussed in

Part~I; and we shall see moreover that the treatment in question may be used also

in cases where the method of separation of variables cannot be applied.

In Part~III the problem of the series spectra of other elements will be treated

from a similar point of view. As pointed out by the writer in an earlier paper, a

simple explanation of the pronounced analogy between these spectra and the
hydrogen

spectrum is offered by the fact, that the atomic systems, involved in the

emission of the spectra under consideration, in a certain sense may be regarded as

a perturbed hydrogen atom. On the other hand, a clue to the interpretation of the

180
characteristic difference between the hydrogen spectrum and the spectra of other

elements was first obtained by \Name{Sommerfeld}'s theory of the stationary


states of central

systems referred to above. As shown by \Name{Sommerfeld}, it is possible on this

theory to account in general outlines for the well known laws governing the
frequencies

of the series spectra of the elements; and, as it will be shown in Part~III, it

is also possible, on the basis of the formal relation between the quantum theory and

the ordinary theory of radiation, in this way to obtain a simple interpretation of the

laws governing the remarkable differences in the intensities with which the various

series of lines appear, which on the combination principle would constitute the

complete spectra under consideration. As regards the detailed discussion of these

spectra, however, it is necessary to bear in mind that the part played by the inner

electrons in the atoms of the elements in question forms a far more intricate
problem

than the perturbing effect of a fixed external field on the hydrogen atom.

For the treatment of this problem the theory of conditionally periodic systems

based on the conditions~\EqI{(22)} does not seem to suffice, while, as it will be


shown

in Part~III, it appears that the method of perturbations exposed in the following

lends itself naturally also to this case.

\Section{2.}{The stationary states of a perturbed~periodic~system.}

181
In Part~I it was shown that the problem of the fixation of the stationary

states of a periodic system of several degrees of freedom, which is subject to the

perturbing influence of a small external field, cannot be treated directly on the


basis

of the general principle of the mechanical transformability of the stationary states

by considering the influence, which on ordinary mechanics a slow establishment

of the external field would exert on the motion of some arbitrarily chosen
stationary

state of the undisturbed system (see Part~I, \PageRefI{23}). This is an immediate

consequence of the fact, mentioned in the former section, that the stationary states

\PageSep{42}

of the perturbed system are characterised by a greater number of extra-mechanical

conditions than the stationary states of the undisturbed system. On the other hand,

we were led to assume from the general formal relation between the quantum
theory

of line spectra and the ordinary theory of radiation, that it is possible to obtain

information about the stationary states of the perturbed system from a direct

consideration of the slow variations which the periodic orbit undergoes as a

consequence of the mechanical effect of the external field on the motion. Thus, if

these variations are of periodic or conditionally periodic type, we may expect that,

in the presence of the external field, the values for the additional energy of the

system in the stationary states are related to the small frequency or frequencies of

182
the perturbations, in a manner analogous to the relation between energy and
frequency

in the stationary states of an ordinary periodic or conditionally periodic system.

If\Pagelabel{42} the equations of motion for the perturbed system can be solved
by means

of separation of variables, it is easily seen that the relation in question is fulfilled

if the stationary states are determined by the conditions~\EqI{(22)}. Consider thus


a system

for which every orbit is periodic, and let us assume that in the presence of a given

small external field a separation of variables is possible in a certain set of


coordinates

$q_{1}$,~\MyDots $q_{s}$. For the undisturbed system we have then, according


to equation~\EqI{(23)},

that the quantity~$I$, defined by~\EqI{(5)}, is equal to $\kappa_{1} I_{1} + \dots


+ \kappa_{s} I_{s}$, where

$I_{1}$,~\MyDots $I_{s}$ are defined by~\EqI{(21)} and calculated with respect


to the set of coordinates

just mentioned, and where the~$\kappa$'s are a set of entire positive numbers
without a

common divisor. For simplicity let us assume that at least one of the~$\kappa$'s,
say~$\kappa_{s}$,

is equal to one, and that consequently, as mentioned on \PageRefI[page]{22}, the


number~$n$ in~\EqI{(24)},

which characterises the stationary states of the undisturbed system, may take

all positive values. This condition will be fulfilled in case of all the applications to

183
spectral problems discussed below; it will be seen, however, that the extension to

problems where this condition is not fulfilled will only necessitate small
modifications

in the following considerations. By use of~\EqI{(29)} we get now for the


difference

in the total energy of two slightly different states of the perturbed system

\[

\delta E = \sum_{1}^{s} \omega_{k}\, \delta I_{k}

= \omega_{s} \sum_{1}^{s} \kappa_{k}\, \delta I_{k}

+ \sum_{1}^{s-1} (\omega_{k} - \kappa_{k} \omega_{s})\, \delta I_{k}.

\Tag{(42)}

\]

Since for the undisturbed system $\omega_{k} = \kappa_{k} \omega_{s}$, the


differences $\omega_{k} - \kappa_{k} \omega_{s}$ appearing

in the last term will, for the perturbed system, be small quantities which will just

represent the frequencies of the slow variations which the orbit undergoes in the

presence of the external field. These quantities will in the following be denoted
by~$\frakv_{k}$.

Consider now the multitude of states of the perturbed system for which $\
sum_{1}^{s} \kappa_{k} I_{k}$

is equal to~$nh$, where $n$~is a given entire positive number. This multitude will
be

seen to include all possible stationary states of the perturbed system, which
satisfy~\EqI{(22)},

and the motion of which differs at any moment only slightly from some

184
stationary motion of the undisturbed system, satisfying~\EqI{(24)} for the given
value of~$n$.

\PageSep{43}

Denoting the value of the energy of the undisturbed system in such a state
by~$E_{n}$,

and the value of the energy of the perturbed system in a state belonging to

the multitude under consideration by $E_{n} + \frakE$, we get from~\Eq{(42)}

\[

\delta \frakE = \sum_{1}^{s-1} \frakv_{k}\, \delta I_{k}

\Tag{(43)}

\]

for the energy difference between two neighbouring states of this multitude. Since

this relation has the same form as~\EqI{(29)}, we see consequently that by putting
$I_{1}$,~\MyDots $I_{s-1}$

equal to entire multipla of~$h$, as claimed by the conditions~\EqI{(22)}, we


obtain exactly

the same relation between the additional energy~$\frakE$ and the small
frequencies~$\frakv_{k}$,

impressed on the system by the external field, as that which holds between the

total energy and the fundamental frequencies in the stationary states of a


conditionally

periodic system of $s - 1$~degrees of freedom.

\begin{Remark}

185
As\Pagelabel{43} a simple illustration of these calculations let us consider the
system consisting

of a particle moving in a plane and subject to an attraction from a fixed point,


which varies

proportional to the distance apart. If undisturbed, the motion of this system will be
periodic

independent of the initial conditions, and the particle will describe an elliptical
orbit with its

centre at the fixed point. Moreover the equations of motion of the undisturbed
system may

be solved by means of separation of variables in polar coordinates, as well as in


any set of

rectangular coordinates. In the first case we have, taking for~$q_{1}$ the length of
the radius vector

from the fixed point to the particle and for~$q_{2}$ the angular distance of this
radius vector

from a fixed direction, $\kappa_{1} = 2$ and $\kappa_{2} = 1$, while in the


second case we have $\kappa_{1} = \kappa_{2} = 1$. In

the presence of an external field the orbit will in general not remain periodic, but
will in the

course of time cover a continuous extension of the plane. If the external field is
sufficiently

small, however, the orbit will at any moment only differ little from a closed
elliptical orbit,

but in the course of time the lengths and directions of the principal axes of this
ellipse will

undergo slow variations. In general the perturbed system will not allow of
separation of
186
variables, but two cases obviously present themselves in which such a separation is
still

possible; in the first case the external field is central with the fixed point as centre,
and a

separation is possible in polar coordinates; in the second case the external field of
force is

perpendicular to a given line and varies as some function of the distance from this
line, and

separation is possible in a set of rectangular coordinates with the axes parallel and
perpendicular

to the given line. In the first case the perturbations will not affect the lengths of

the principal axes of the elliptical orbit and will only produce a slow uniform
rotation of

the directions of these axes, while in the second case the lengths of the principal
axes as well

as their directions will perform slow oscillations. It will consequently be seen that,
by fixing

the stationary states of the perturbed system by means of the conditions~\


EqI{(22)}, the cycles of

shapes and positions which the orbit of the particle will pass through in the
stationary states

will be entirely different in the two cases. In both cases, however, it will be seen
that the

frequency $\frakv = \omega_{1} - \kappa_{1} \omega_{2}$ will be equal to the


frequency with which the orbit at regular

intervals re-assumes its shape and position. By fixing the stationary states by~\
EqI{(22)} we obtain

187
therefore, as seen from~\Eq{(43)}, in both cases that the relation between this
frequency and the

additional energy of the system due to the presence of the field will be the same as
the

relation between energy and frequency in the stationary states of a system of one
degree of

freedom; and it will be seen that the above considerations offer a dynamical
interpretation

of the characteristic discontinuity involved in the application of the method of


separation of

%[** TN: Footnote mark before punctuation in the original]

variables to the fixation of the stationary states of perturbed periodic systems.\


footnote

{In this connection it may be of interest to note that the possibility of a rational
interpretation of

the discontinuity in question would seem to be essentially connected with the


form of the principles

of the quantum theory adopted in this paper. If for instance the quantum theory is
taken in the

form proposed by \Name{Planck} in his second theory of temperature radiation,


the consequent development

to periodic systems of several degrees of freedom would seem to involve a


serious difficulty as regards

the question of the necessary stability of the temperature equilibrium among a


great number of systems

for small variations of the external conditions. In fact, in connection with the
development of his theory

188
of the ``physical structure of the phase space'', mentioned in Part~I on \
PageRefI[page]{18}, in which conditions

of the same type as~\EqI{(22)} are established, \Name{Planck} has deduced


expressions for the total energy of a

great number of systems in temperature equilibrium, which, if applied to systems


of the same kind

as those considered in the above example, show a dependency of this energy on


the temperature which

is different, according to whether polar coordinates or rectangular coordinates are


used as basis for the

structure of the phase space.}

\end{Remark}

\PageSep{44}

In general it will not be possible to solve the equations of motion of the perturbed

system by means of separation of variables in a fixed set of positional coordinates,

but we shall see that the problem of the fixation of the stationary states

of the perturbed system may be attacked by a direct examination of the additional

energy of the system and its relation to the slow variations of the orbit, on the

basis of the usual theory of perturbations well known from celestial mechanics.

Consider a system for which every orbit, if undisturbed, is periodic independent of

the initial conditions, and let us assume that the equations of motion for some set

of coordinates $q_{1}$,~$q_{2}$,~\MyDots $q_{s}$ are solved by means of the \


Typo{Hamilton-Jacobi}{\Name{Hamilton}-\Name{Jacobi}} partial

189
differential equation, given by formula~\EqI{(17)} in Part~I\@. The motion of the
system is

then determined by the equations~\EqI{(18)}, and the orbit is characterised by


means of

the constants $\alpha_{1}$,~\MyDots $\alpha_{s}$, $\beta_{1}$,~\MyDots $\


beta_{s}$. If now the system is subject to some small

external field of force, the motion will no more be periodic, but, defining in the

usual way the osculating orbit at a given moment as the periodic orbit which

would result if the external forces vanished suddenly at this moment, we find that

the constants $\alpha_{1}$,~\MyDots $\alpha_{s}$, $\beta_{1}$,~\MyDots $\


beta_{s}$, characterising the osculating orbit, will vary

slowly with the time. Assuming for the present that the external forces possess a

constant potential~$\Omega$ given as a function of the~$q$'s, we have according


to the theory

of perturbations that the rates of variation of the orbital constants of the osculating

orbit will be given by\footnote

{See f.\ inst.\ \Name{C.~V.~L. Charlier}, Die Mechanik des Himmels, Bd.~I,
Abt.~1, §~10.}

\[

\frac{d\alpha_{k}}{dt} = -\frac{d\Omega}{d\beta_{k}},\quad

\frac{d\beta_{k}}{dt} = \frac{d\Omega}{d\alpha_{k}},\qquad

(k = 1, \MyDots s)

\Tag{(44)}

\]

190
where $\Omega$~is considered as a function of $\alpha_{1}$,~\MyDots $\
alpha_{s}$, $\beta_{1}$,~\MyDots $\beta_{s}$ and~$t$, obtained by

introducing for the~$q$'s their expressions as functions of these quantities obtained

by solving~\EqI{(18)}. The equations~\Eq{(44)} allow to follow completely the


perturbing

effect of the external field on the motion of the system. For the problem under

consideration, however, a detailed examination of the perturbations is not


necessary.

In fact, we shall not be concerned with the small deformation of the orbit

characterised by the small oscillations of the orbital constants within a time


interval

of the same order of magnitude as the period of the osculating orbit, but only

\PageSep{45}

with the so\Typo{}{ }called ``secular perturbations'' of the orbit, characterised by


the total

variation of these constants taken over a time interval long compared with the

period of the osculating orbit. As we shall see below, these variations may, with

an approximation sufficient for our purpose, be obtained directly by taking mean

values on both sides of the equations~\Eq{(44)}. Before entering on these


calculations,

however, it may be observed that the part played by the constants $\alpha_{1}$
and $\beta_{1}$ differs

essentially from that played by the other orbital constants $\alpha_{2}$,~\MyDots


$\alpha_{s}$, $\beta_{2}$,~\MyDots $\beta_{s}$.

Thus from the formulæ \EqI{(17)} and \EqI{(18)} on \PageRefI[page]{19}, it


follows that $\alpha_{1}$~is the total

191
energy corresponding to the osculating orbit, while $\beta_{1}$~will represent the
moment

in which the system would pass some distinguished point in this orbit. If for
instance

we consider the perturbations of a Keplerian motion, we may for~$\beta_{1}$ take

the so called time of perihelium\Note{[sic]} passage. When discussing the secular


perturbations

of the shape and position of the orbit, we see therefore in the first place that

the variations of~$\beta_{1}$ may be left out of consideration. Further, it follows


from the

principle of conservation of energy, that $\alpha_{1} + \Omega$ will remain


constant during the

motion, and that consequently during the perturbations $\alpha_{1}$~will change


only by

small quantities of the same order as~$\lambda\alpha_{1}$, where $\


lambda$~denotes a small constant of the

same order of magnitude as the ratio between the external forces and the internal

forces of the system. Moreover, since the period~$\sigma$ of the undisturbed


motion depends

on $\alpha_{1}$~only, it follows that the period of the osculating orbit will remain

constant during the perturbations, with neglect of small quantities of the same

order as~$\lambda\sigma$. On the other hand it follows from~\Eq{(44)} that, in a


time interval of the

same order as~$\sigma/\lambda$, the constants $\alpha_{2}$,~\MyDots $\


alpha_{s}$, $\beta_{2}$,~\MyDots $\beta_{s}$ will in general undergo variations

of the same order of magnitude as the values of these constants themselves.

192
As mentioned above, the total variations of the constants $\alpha_{2}$,~\MyDots
$\alpha_{s}$, $\beta_{2}$,~\MyDots $\beta_{s}$,

which characterise the \Emph{secular perturbations of the shape and position of

the orbit}, may be obtained by taking mean values on both sides of the equations~\
Eq{(44)}.

Introducing a function~$\psi$ of the~$\alpha$'s and~$\beta$'s, equal to the mean


value of the

potential~$\Omega$ taken over a period~$\sigma$ of the motion of the


undisturbed system and

defined by the formula

\[

\psi = \frac{1}{\sigma} \int_{t}^{t+\sigma} \Omega\, dt,

\Tag{(45)}

\]

it is easily seen, since $\sigma$~depends only on~$\alpha_{1}$, that the mean


values of the partial

differential coefficients of~$\Omega$ with respect to $\alpha_{2}$,~\MyDots $\


alpha_{s}$, $\beta_{2}$,~\MyDots $\beta_{s}$, taken over an approximate

period of the perturbed motion, may, if we look apart from small

quantities proportional to~$\lambda^{2}$, be replaced by the values of the


corresponding partial

differential coefficients of~$\psi$ at some moment within this period. With the
approximation

mentioned we get therefore

193
\[

\frac{D\alpha_{k}}{Dt} = -\frac{\dd \psi}{\dd \beta_{k}},\quad

\frac{D\beta_{k}}{Dt} = \frac{\dd \psi}{\dd \alpha_{k}},\qquad

(k = 2, \MyDots s)

\Tag{(46)}

\]

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8.~Række,


IV.~1. 7

\PageSep{46}

where the differential symbols on the left sides are written to indicate mean values

of the rates of variation of the orbital constants during an approximate period of

the perturbed motion. From the definition of~$\psi$ it follows that this quantity in

general will depend on~$\alpha_{1}$ as well as on $\alpha_{2}$,~\MyDots $\


alpha_{s}$, $\beta_{2}$,~\MyDots $\beta_{s}$, but that it will not

depend upon~$\beta_{1}$. From the above considerations it follows further that,


with the

approximation in question, $\alpha_{1}$~may be considered as a constant in the


expressions

on the right sides of~\Eq{(46)}, while for $\alpha_{2}$,~\MyDots $\alpha_{s}$,


$\beta_{2}$,~\MyDots $\beta_{s}$ we may take a set of values

corresponding to some moment within the period to which the mean values on the

left sides refer.

194
It will be seen that the equations~\Eq{(46)} allow to follow the secular
perturbations

during a time interval sufficiently long for the external forces to produce a
considerable

change in the shape and position of the original orbit, if in the total

variations of the orbital constants $\alpha_{2}$,~\MyDots $\alpha_{s}$, $\


beta_{2}$,~\MyDots $\beta_{s}$ we look apart from small

quantities of the same order as the small oscillations of these constants within a

single period. As a consequence of the secular variations, the orbit will pass
through

a cycle of shapes and positions, which will depend on its original shape and
position

and on the character of the perturbing field, but not on the intensity of this

field. In fact, as seen from~\Eq{(46)}, the variations in the shape and position of
the orbit

will remain the same if $\psi$~is multiplied by a constant factor, which will only
influence

the rate at which these variations are performed. It will further be observed

that the problem of determining the secular perturbations by means of~\Eq{(46)}


consists

in solving a set of equations of the same type as the Hamiltonian equations

of motion for a system of $s - 1$~degrees of freedom. In these equations the


quantity~$\psi$

plays formally the same part as the total energy in the usual mechanical problem,

and in analogy with the principle of conservation of energy it follows directly


from~\Eq{(46)}

195
that, with neglect of small quantities proportional to~$\lambda^{2}$, \Emph{the
value of~$\psi$ will

remain constant during the perturbations}, even if the external forces act

through a time interval of the same order as~$\sigma/\lambda$. In fact, with


neglect of small

quantities proportional to~$\lambda^{2}$, we have

\[

\BreakMath{%

\frac{D\psi}{Dt}

= \sum_{2}^{s} \left(\frac{\dd \psi}{\dd \alpha_{k}}\, \frac{D \alpha_{k}}{Dt}

+ \frac{\dd \psi}{\dd \beta_{k}}\, \frac{D \beta_{k}}{Dt}\right)

= \sum_{2}^{s} \left(-\frac{\dd \psi}{\dd \alpha_{k}}\, \frac{\dd \psi}{\dd \


beta_{k}}

+ \frac{\dd \psi}{\dd \beta_{k}}\, \frac{\dd \psi}{\dd \alpha_{k}}\


right)

= 0.

}{%

\begin{aligned}

\frac{D\psi}{Dt}

&= \sum_{2}^{s} \left(\frac{\dd \psi}{\dd \alpha_{k}}\, \frac{D \alpha_{k}}


{Dt}

+ \frac{\dd \psi}{\dd \beta_{k}}\, \frac{D \beta_{k}}{Dt}\right) \\

&= \sum_{2}^{s} \left(-\frac{\dd \psi}{\dd \alpha_{k}}\, \frac{\dd \psi}{\dd \


beta_{k}}

196
+ \frac{\dd \psi}{\dd \beta_{k}}\, \frac{\dd \psi}{\dd \alpha_{k}}\
right)

= 0.

\end{aligned}

\]

Since at any moment $\psi$~will differ only by small quantities proportional to~$\
lambda^{2}$

from the mean value of the potential of the external forces taken over an
approximate

period of the perturbed motion, it follows from the above that, with neglect

of small quantities of this order, also the mean value of the inner energy~$\
alpha_{1}$ of the

perturbed system, taken over an approximate period, will remain constant during

the perturbations, even if the perturbing forces act through a time interval long

enough to produce a considerable change in the shape and position of the orbit.

In the special case, where the perturbed system allows of separation of variables,

this last result may be shown to follow directly from formula~\EqI{(28)} in Part~I\
@.

Taking for the time interval~$\theta$ in this formula the period~$\sigma$ of the
undisturbed

\PageSep{47}

motion, we get $N_{k} = \kappa_{k}$, where $\kappa_{1}$,~\MyDots $\


kappa_{s}$ are the numbers entering in formula~\EqI{(23)}.

197
Comparing a given perturbed motion of the system with some undisturbed

motion of which it may be regarded as a small variation, we get therefore from~\


EqI{(28)},

with neglect of small quantities proportional to the square of the intensity

of the external forces,

\[

\int_{0}^{\sigma} \delta E\, dt

= \sum_{1}^{s} \kappa_{k}\, \delta I_{k},

\Tag{(47)}

\]

where the $I$'s are calculated with respect to a set of coordinates in which a
separation

can be obtained for the perturbed motion, and where $\delta E$~is the difference

between the total energy of the undisturbed motion and the energy which the

system would possess in its perturbed state, if the external forces vanished
suddenly

at the moment under consideration, and which in the above calculations was

denoted by~$\alpha_{1}$. Now the energy~$E$ of the undisturbed motion is


determined completely

by the value of $I = \sum \kappa_{k} I_{k}$. If therefore the perturbed motion is


all the

time compared with a neighbouring undisturbed motion of given constant energy,

it follows directly from~\Eq{(47)}, that, with neglect of small quantities of the


same order

as the square of the external forces, the integral on the left side, taken over an

198
approximate period of the perturbed motion, will remain unaltered during the
perturbations

through any time interval, however long.

Before\Pagelabel{47} proceeding with the applications of the equations~\Eq{(46)}


which apply to

the case of a constant perturbing field, it will be necessary to consider \Emph{the


effect

of a slow and uniform establishment of the external field}. Let us

assume that, within the interval $0 < t < \theta$ where $\theta$~denotes a quantity
of the same

order as~$\sigma/\lambda$, the intensity of the external field increases uniformly


from zero to the

value corresponding to the potential~$\Omega$. Since the variation in the


perturbing field

during a single period will only be a small quantity of the same order as~$\
lambda^{2}$, we

see in the first place that the secular variations of the constants $\alpha_{2}$,~\
MyDots $\alpha_{s}$, $\beta_{2}$,~\MyDots $\beta_{s}$,

with the same approximation as for a constant field, will be given by a set of
equations

of the same form as~\Eq{(46)}, with the only difference that $\psi$~is replaced
by~$\dfrac{t}{\theta} \psi$.

Moreover it may be shown that in these equations the quantity~$\alpha_{1}$ may


be considered

as constant, just as in the equations which hold for a constant perturbing

199
field. In fact the total variation in~$\alpha_{1}$ at any moment~$t$ will be equal
to the total

work performed by the external forces since the beginning of the establishment

of the perturbing field, and will therefore be given by

\[

\Delta_{t} \alpha_{1}

= -\int_{0}^{t} \frac{t}{\theta} \sum_{1}^{s} \frac{\dd \Omega}{\dd q_{k}}\, \


dot{q}_{k}\, dt

= \frac{1}{\theta} \int_{0}^{t} \Omega\, dt - \frac{t}{\theta} \Omega_{t},

\Tag{(48)}

\]

where the expression on the right side is obtained by partial integration; but, since
both

terms in this expression are of the same order of magnitude as~$\lambda \


alpha_{1}$, we see that the

\PageSep{48}

total variation in~$\alpha_{1}$ within the interval in question will, just as in case
of a constant

perturbing field, be only a small quantity of this order. We get therefore

the result, that, for the same shape and position of the original orbit, the cycle of

shapes and positions passed through by the orbit during the increase of the external

field will be the same as that which would appear for a constant perturbing field,

and that, with neglect of small quantities proportional to~$\lambda^{2}$, the value
of the

200
function~$\psi$ will consequently remain constant during the establishment of the
field.

With this approximation we get therefore from~\Eq{(48)}, putting $t = \theta$,

\[

\Delta_{\theta} \alpha_{1} + \Omega_{\theta}

= \frac{1}{\theta} \int_{0}^{\theta} \Omega\, dt

= \psi,

\]

which shows that the change in the total energy of the system, due to the slow and

uniform establishment of the external field, is just equal to the value of the
function~$\psi$,

and consequently equal to the mean value of the potential of the external

forces taken over an approximate period of the perturbed motion. This result may

also be expressed by stating, that, with neglect of small quantities proportional to


the

square of the external forces, the mean value of the inner energy taken over an

approximate period of the perturbed motion will be equal to the energy possessed

by the system before the establishment of the perturbing field.

Returning now to the problem of the fixation of the stationary states of a

periodic system subject to the influence of a small external field of constant


potential,

we shall base our considerations on the fundamental assumption that these states

are distinguished between the continuous multitude of mechanically possible states

201
by a relation between the additional energy of the system due to the presence of

the external field and the frequencies of the slow variations of the orbit produced

by this field, which is analogous to the relation discussed on \PageRef[page]{42}


in the special case

in which the perturbed system allows of separation of variables in a fixed set of

coordinates. On this assumption we shall expect in the first place that, apart from

small quantities proportional to~$\lambda$, the cycles of shapes and positions of


the orbit

belonging to the stationary states of the perturbed system will depend only on the

character of the external field, but not on its intensity. Since now, as shown above,

such a cycle will remain unaltered during a slow and uniform increase of the
intensity

of the external field if the effect of the external forces is calculated by means

of ordinary mechanics, we are therefore, with reference to the principle of the

mechanical transformability of the stationary states, led to the conclusion that it is

possible by direct application of ordinary mechanics, not only to follow the secular

perturbations of the orbit in the stationary states corresponding to a constant


external

field, but also to calculate the variation in the energy of the system in the

stationary states which results from a slow and uniform change in the intensity of

this field. If we denote the energy in the stationary states of the perturbed system

by $E_{n} + \frakE$, where $E_{n}$~is the value of the energy in the stationary
state of the undisturbed

\PageSep{49}

202
system characterised by a given entire value of~$n$ in the condition $I = nh$,

we may therefore conclude from the above that \Emph{the additional energy~$\
frakE$ in

the stationary states of the perturbed system will be equal to the

value in these states of the function~$\psi$ defined by~\Eq{(45)}, if we look

apart from small quantities proportional to the square of the intensity

of the external forces}. It will be seen that this result is equivalent

to the statement, that the mean value of the inner energy taken over an approximate

period of the perturbed motion will be equal to the value~$E_{n}$ of the energy in
the

corresponding stationary state of the undisturbed system. In case of the perturbed

system allowing of separation of variables in a fixed set of coordinates, this result

may be simply shown to be a direct consequence of the fixation of the stationary

states by means of the conditions~\EqI{(22)}. In fact, if we assume that the


undisturbed

motion, considered in~\Eq{(47)}, corresponds to some stationary state,


satisfying~\EqI{(24)}

for a given value of~$n$, and that the perturbed motion is also stationary and
satisfies~\EqI{(22)},

we see that the right side of~\Eq{(47)} will be zero, and we get the result that

the mean value of the inner energy in the stationary states of the system, with the

approximation mentioned, will not be altered in the presence of the external field.

Due to the above result that the additional energy~$\frakE$ in the stationary states

203
of the perturbed system, with neglect of small quantities proportional to~$\
lambda^{2}$, may

be taken equal to the value in these states of the function~$\psi$ entering in the

equations~\Eq{(46)} which determine the secular perturbations of the orbits, we


are

now able to draw further conclusions from the fact, mentioned above, that these

equations are of the same type as the Hamiltonian equations of motion for a

mechanical system of $s - 1$~degrees of freedom. In fact, we see that \Emph{the


fixation

of the stationary states of the perturbed system is reduced to a problem

which is formally analogous to the fixation of these states for

a mechanical system of less degrees of freedom}. As it will appear from

the following applications this problem may, quite independent of the possibility of

separation of variables for the perturbed system, be treated directly on the basis of

the fundamental relation between energy and frequency in the stationary states of

periodic or conditionally periodic systems, discussed in Part~I, if only the solution

of the equations~\Eq{(46)} is of a periodic or conditionally periodic character. In


this connection

it may once more be emphasised that these equations, according to the

manner in which they were deduced, allow to follow the secular perturbations only

through a time interval of the same order of magnitude as that sufficient for the

external forces to produce a finite alteration in the shape and position of the orbit.

With reference to the necessary stability of the stationary states of an atomic


system,

204
it seems justified, however, to conclude that any possible small discrepancy
between

the motion to be expected from a rigorous application of ordinary mechanics and


that

determined by a calculation of the secular perturbations, based on the equations~\


Eq{(46)}, cannot cause a material change in the character of the stationary states as

fixed by a consideration of the periodicity properties of these perturbations. On the

\PageSep{50}

other hand, from the point of view of the general formal relation between the

quantum theory and the ordinary theory of radiation, we must be prepared to find

that the motion and the energy in the stationary states of a perturbed periodic

system, for which we only know that the secular perturbations as determined by~\
Eq{(46)}

are of conditionally periodic type, will not be as sharply defined as the motion

and the energy in the stationary states of a conditionally periodic system for which

the equations of motion allow of a rigorous solution by means of the method of

separation of variables. Thus, if we consider a large number of similar atomic

systems of the type in question, we may be prepared to find that the values of the

additional energy in a given stationary state will for the different systems deviate

from each other by small quantities; but it must be expected that the values of

the additional energy for the large majority of systems will differ from the value

of~$\psi$, as determined by the method indicated above, only by small quantities


proportional

to~$\lambda^{2}$, and that only for a small fraction (at most of the same order
as~$\lambda^{2}$)

205
of the systems the values of the additional energy will show deviations from

this value of~$\psi$, which are of the same order as~$\lambda$.

\medskip

As\Pagelabel{50} to the application of the preceding considerations to special


problems, it will

be seen in the first place that in case of a \Emph{perturbed periodic system


possessing

two degrees of freedom}, as for instance that considered in the example

on \PageRef[page]{43}, the problem of the fixation of the stationary states of the


perturbed

system in the presence of a small external field allows of a general solution on

the basis of the method developed above, because in this case the secular
perturbations

will in general be \Emph{simply periodic}. In fact, in this case the shape and

position of the orbit are characterised by two constants $\alpha_{2}$ and~$\


beta_{2}$, and from the

equations~\Eq{(46)}, which will be analogous to the equations of motion of a


system of

one degree of freedom, it follows directly that during the perturbations $\


alpha_{2}$~will be

a function of~$\beta_{2}$ and that in general these quantities will be periodic


functions of

the time with a period~$\fraks$ which, besides on~$\alpha_{1}$, will depend on


the value of $\psi$~only.

Considering two slightly different states of the perturbed system for which the

206
corresponding states of the undisturbed system (\ie\ the states which would appear

if the external forces vanished at a slow and uniform rate) possess the same energy

and consequently the same value for the quantity~$I$ defined by~\EqI{(5)}, we
get therefore

by a calculation completely analogous to that leading to relation~\EqI{(8)} in


Part~I, which

was deduced directly from the Hamiltonian equations, for the difference in the

values of the function~$\psi$ for these two states

\[

\delta \psi = \frakv\, \delta \frakI,

\Tag{(49)}

\]

where $\frakv = \dfrac{1}{\fraks}$ is the frequency of the secular perturbations,


and where the quantity~$\frakI$

is defined by

\[

\frakI = \int_{0}^{\fraks} \alpha_{2}\, \frac{D\beta_{2}}{Dt}\, dt

= \int \alpha_{2}\, D\beta_{2},

\Tag{(50)}

\]

\PageSep{51}

where the latter integral is taken over a complete oscillation of~$\beta_{2}$. In


order to fix

the stationary states, it will now be seen in the first place that, among the multitude

207
of states of the perturbed system for which the value of~$I$ in the corresponding

states of the undisturbed system is equal to~$nh$ where $n$~is a given positive
integer,

the state for which $\frakI = 0$ must beforehand be expected to be a stationary


state. In

fact, for this value of~$\frakI$, the shape and position of the orbit will not undergo
secular perturbations

but will remain unaltered for a constant external field as well as during a slow

and uniform establishment of this field. In contrast to what in general will take
place

during a slow establishment of the external field, we may therefore expect that, for

this special shape and position of the orbit, a direct application of ordinary
mechanics

will be legitimate in calculating the effect of the establishment of the field, since

there will in this case obviously be nothing to cause the coming into play of some

non-mechanical process, connected with the mechanism of a transition between

two stationary states accompanied by the emission or absorption of a radiation of

small frequency. With reference to relation~\Eq{(49)} we see therefore that, by


fixing the

stationary states of the perturbed system by means of the condition

\[

\frakI = \frakn h,

\Tag{(51)}

\]

208
where $\frakn$~is an entire number, we obtain a relation between the additional
energy

$\frakE = \psi$ of the system in the presence of the field and the frequency~$\
frakv$ of the secular

perturbations, which is exactly of the same type as that which holds between the
energy

and frequency in the stationary states of a system of one degree of freedom, and

which is expressed by \EqI{(8)}~and~\EqI{(10)}. By means of~\Eq{(51)} it is


possible, with neglect of

small quantities proportional to the square of the perturbing forces, directly to

determine the value of the additional energy in the stationary states of a periodic

system of two degrees of freedom subject to an arbitrarily given small external


field

of force, and consequently with this approximation, by use of the fundamental


relation~\EqI{(1)},

to determine the effect of this field on the frequencies of the spectrum of the

undisturbed periodic system. In general this effect will consist in a splitting up of

each of the spectral lines into a number of components which are displaced from

the original position of the line by small quantities proportional to the intensity of

the external forces.

When we pass to \Emph{perturbed periodic systems of more than two degrees

of freedom}, the general problem is more complex. For a given external field,

however, it may be possible to choose a set of orbital constants $\alpha_{2}$,~\


MyDots $\alpha_{s}$,

209
$\beta_{2}$,~\MyDots $\beta_{s}$ in such a way, that during the motion every of
the~$\beta$'s will depend on

the corresponding $\beta$ only, while every of the~$\beta$'s will oscillate between
two fixed

limits. From analogy with the theory of ordinary conditionally periodic systems

which allow of separation of variables, the perturbations may in such a case be

said to be \Emph{conditionally periodic}, and, from a calculation quite analogous


to

that leading to equation~\EqI{(29)} in Part~I which is based entirely on the use of


the

Hamiltonian equations, we get for the difference in~$\psi$ for two slightly
different states

\PageSep{52}

of the perturbed system, for which the value of~$I$ in the corresponding states of
the

undisturbed system is the same,

\[

\delta \psi = \sum_{1}^{s-1} \frakv_{k}\, \delta \frakI_{k},

\Tag{(52)}

\]

where $\frakv_{k}$~is the mean frequency of oscillation of~$\beta_{k+1}$


between its limits, and where

the quantities~$\frakI_{k}$ are defined by\Pagelabel{52}

\[

\frakI_{k} = \int \alpha_{k+1}\, D\beta_{k+1},\qquad

210
(k = 1, \MyDots s-1)

\Tag{(53)}

\]

where the integral is taken over a complete oscillation of~$\beta_{k+1}$. In


analogy with

the expression~\EqI{(31)} for the displacements of the particles of an ordinary


conditionally

periodic system which allows of separation of variables, we get further in the

present case that every of the $\alpha$'s and~$\beta$'s may be expressed as a


function of the

time by a sum of harmonic vibrations of small frequencies

\[

\BreakMath{%

\left.

\begin{gathered}

\alpha \\

\beta

\end{gathered}

\right\}

= \sum \frakC_{\frakt_{1}, \MyDots \frakt_{s-1}} \cos 2\pi \bigl\{(\frakt_{1} \


frakv_{1} + \dots \Add{+} \frakt_{s-1} \frakv_{s-1})t + \frakc_{\frakt_{1}, \
MyDots \frakt_{s-1}}\bigr\},

}{%

\begin{split}

211
\MathBack

\left.

\begin{gathered}

\alpha \\

\beta

\end{gathered}

\right\}

= \sum \frakC_{\frakt_{1}, \MyDots \frakt_{s-1}} \cos 2\pi \bigl\{(\frakt_{1} \


frakv_{1} + \dots \Add{+} \frakt_{s-1} \frakv_{s-1})t \\

+ \frakc_{\frakt_{1}, \MyDots \frakt_{s-1}}\bigr\},

\end{split}

\Tag{(54)}

\]

where the $\frakC$'s~and $\frakc$'s are constants, the former of which, besides
on~$I$, depend on

the $\frakI$'s~only, and where the summation is to be extended over all positive
and

negative entire values of the~$\frakt$'s. If therefore the secular perturbations are


conditionally

periodic, we may conclude that the stationary states of the perturbed

system, corresponding to a given stationary state of the undisturbed system, will be

characterised by the $s - 1$~conditions

\[

212
\frakI_{k} = \frakn_{k} h,\qquad

(k = 1, \MyDots s - l)

\Tag{(55)}

\]

where $\frakn_{1}$,~\MyDots $\frakn_{s-1}$ form a set of entire numbers. In


fact, as seen from~\Eq{(52)}, we obtain

in this way a relation between the additional energy and the frequencies of the

secular perturbations of exactly the same type as that holding for the energy and

frequencies of ordinary conditionally periodic systems and expressed by \


EqI{(22)}~and

\EqI{(29)}; moreover we may conclude beforehand that the state in which every of
the

quantities~$\frakI_{k}$, defined by~\Eq{(53)}, is equal to zero must belong to the


stationary states of the

perturbed system, because in this case the orbit will not undergo secular
perturbations

for a constant external field, nor during a slow and uniform establishment of this

field. Since the conditions~\Eq{(55)}, with neglect of small quantities proportional


to the

square of the intensities of the external forces, allow to determine the additional

energy of the system due to the presence of the external field, we see therefore that

the effect of this field on the spectrum of the undisturbed system, if the secular

perturbations are conditionally periodic, will consist in a splitting up of each


spectral

line in a number of components, in analogy with the effect of a perturbing

213
field on the spectrum of a periodic system of two degrees of freedom. In general,

however, the perturbations, which a periodic system of more than two degrees of

\PageSep{53}

freedom undergoes in the presence of a given external field, cannot be expected

to be conditionally periodic and to exhibit periodicity properties of the type


expressed

by formula~\Eq{(54)}. In such cases it seems impossible to define stationary

states in a way which leads to a complete fixation of the total energy in these

states, and we are therefore led to the conclusion, that the effect of the external

field on the spectrum will not consist in the splitting up of the spectral lines of the

original system into a number of sharp components, but in a diffusion of these

lines over spectral intervals of a width proportional to the intensity of the external

forces.

In special cases in which the secular perturbations of a perturbed periodic

system of more than two degrees of freedom are of conditionally periodic type, it

may occur that these perturbations are characterised by a number of fundamental

frequencies, which is less than~$s - 1$. In such cases, in which the perturbed
periodic

system from analogy with the terminology used in Part~I may be said to be

\Emph{degenerate}, the necessary relation between the additional energy and the
frequencies

of the secular perturbations is secured by a number of conditions less than that

214
given by~\Eq{(55)}, and the stationary states are consequently characterised by a
number

of conditions less than~$s$. With a typical example of such systems we meet if, for

a perturbed periodic system of more than two degrees of freedom, the secular
perturbations

are \Emph{simply periodic} independent of the initial shape and position of

the orbit. In direct analogy to what holds for perturbed periodic systems of two

degrees of freedom, the difference between the values of~$\psi$ in two slightly
different

states of the perturbed system, corresponding to the same value of~$I$, will in the

present case be given by

\[

\delta \psi = \frakv\, \delta \frakI,

\Tag{(56)}

\]

where $\frakv$~is the frequency of the secular perturbations, and where $\


frakI$~is defined by

\[

\frakI = \int_{0}^{\frakv} \sum_{2}^{s} \alpha_{k} \frac{D\beta_{k}}{Dt}\, dt,

\Tag{(57)}

\]

where $\fraks = 1/\frakv$ is the period of the perturbations. We may therefore


conclude that

the stationary states of the perturbed system, corresponding to a given stationary

215
state of the undisturbed system, will be characterised by the single condition

\[

\frakI = \frakn h,

\Tag{(58)}

\]

in which $\frakn$~is an entire number, and which will be seen to be completely


analogous

to the condition which fixes the stationary states of ordinary periodic systems of

several degrees of freedom.

\medskip

In the following sections we shall apply the preceding considerations to the

problem of the fixation of \Emph{the stationary states of the hydrogen atom},

when the relativity modifications are taken into account, and when the atom is
exposed

to small external fields. In this discussion we shall for the sake of simplicity

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8. Række, IV. 1.


8

\PageSep{54}

consider the mass of the nucleus as infinite in the calculations of the perturbations

of the orbit of the electron.\Pagelabel{54} This involves, in the expression for the
additional

energy of the system, the neglect of small terms of the same order as the product

216
of the intensity of the external forces with the ratio between the mass of the
electron

and the mass of the nucleus, but due to the smallness of the latter ratio

the error introduced by this simplification will be of no importance in the


comparison

of the results with the measurements. Since in the case under consideration

the system possesses three degrees of freedom, the equations which determine the

secular perturbations of the orbit of the electron will correspond to the equations

of motion of a system of two degrees of freedom, and it will therefore not be


possible

to give a general treatment of the problem of the stationary states. Thus, for

any given external field, we meet with the question whether the perturbations are

conditionally periodic and, if so, in what set of orbital constants this periodicity

may be conveniently expressed. Now, in many spectral problems, the external field

possesses \Emph{axial symmetry round an axis through the nucleus}, and in this

case it is easily shown that the problem of the fixation of the stationary states

allows of a general solution. A choice of orbital constants which is suitable for the

discussion of this problem, and which is well known from the astronomical theory

of planetary perturbations, is obtained by choosing for~$\alpha_{2}$ the total


angular momentum

of the electron round the nucleus and for~$\alpha_{3}$ the component of this
angular momentum

round the axis of the field. For the set of~$\beta$'s, corresponding to this set of~$\
alpha$'s,

217
we may take $\beta_{2}$~equal to the angle, which the major axis makes with the
line

in which the plane of the orbit cuts the plane through the nucleus perpendicular

to the axis of the field, and $\beta_{3}$~equal to the angle between this line and a
fixed

direction in the latter plane. For the problem under consideration it will be seen

that, with this choice of constants, the mean value~$\psi$ of the potential of the
perturbing

field will, besides on~$\alpha_{1}$, generally depend on $\alpha_{2}$~and~$\


beta_{2}$ as well as on~$\alpha_{3}$,

but due to the symmetry round the axis it will obviously not depend on~$\
beta_{3}$. In

consequence of this, the equations~\Eq{(46)}, which determine the secular


perturbations,

will possess the same form as the Hamiltonian equations of motion for a particle

moving in a plane and subject to a central field of force. Thus corresponding to the

conservation of angular momentum for central systems, we get in the first place

from~\Eq{(46)} that $\alpha_{3}$~will remain unaltered during the perturbations.


Next corresponding

to the simple periodicity of the radial motion in central systems, we see from~\
Eq{(46)},

if $\alpha_{3}$ as well as~$\alpha_{1}$ is considered as a constant, that during


the perturbations $\alpha_{2}$~will

be a function of~$\beta_{2}$ and vary in a simple periodic way with the time. The
perturbations

of the orbit of the electron produced by an external field which possesses

218
axial symmetry will therefore always be of \Emph{conditionally periodic} type,

quite independent of the possibility of separation of variables for the perturbed

system. As regards the form of the conditions which fix the stationary states, it
may

be noted, however, that with the choice of orbital constants under consideration

the~$\beta$'s will not, as it was assumed for the sake of simplicity in the general
discussion

\PageSep{55}

on \PageRef[page]{52}, oscillate between fixed limits, but it will be seen that $\


beta_{2}$~during

the perturbations may either oscillate between two such limits or increase (or
decrease)

continuously, while $\beta_{3}$~will always vary in the latter manner. This


constitutes,

however, only a formal difficulty of the same kind as that mentioned in

Part~I in connection with the discussion of the conditions~\EqI{(16)}, which fix


the

stationary states of a system consisting of a particle moving in a central field of

force. Thus from a simple consideration it will be seen that, in complete analogy

to the relations \Eq{(52)} and~\Eq{(53)}, we get in the present case for the
difference between

the energy of two slightly different states of the perturbed system, which
correspond

to the same value of~$I$,

\[

219
\delta \psi = \frakv_{1}\, \delta \frakI_{1} + \frakv_{2}\, \delta I_{2},

\Tag{(59)}

\]

where $\frakv_{1}$~is the frequency with which the shape of the orbit and its
position relative

to the axis of the field repeats itself at regular intervals and which is characterised

by the variation of $\alpha_{2}$~and~$\beta_{2}$, while $\frakv_{2}$~is the


mean frequency of rotation of

the plane of the orbit round this axis characterised by the variation of~$\beta_{3}$,
and

where $\frakI_{1}$~and~$\frakI_{2}$ are defined by the equations

\[

\frakI_{1} = \int \alpha_{2}\, D\beta_{2},\quad

\frakI_{2} = \int_{0}^{2\pi} \alpha_{3}\, D\beta_{3} = 2\pi \alpha_{3}.

\Tag{(60)}

\]

In case $\beta_{2}$~varies in an oscillating manner with the time, the first integral
must be

taken over a complete oscillation of this orbital constant, while, if $\


beta_{2}$~during the

perturbations increases or decreases continuously, the integral in the expression


for~$\frakI_{1}$

must be taken over an interval of~$2\pi$, just as the integral in the expression

for~$\frakI_{2}$. By fixing the stationary states of the perturbed system by means


of the two\Pagelabel{55}

220
conditions\footnote

{Quite apart from the problem of perturbed periodic systems, the second of these
conditions

would also follow directly from certain interesting considerations of \


Name{Epstein} (Ber.\ d.\ D. Phys.\ Ges.~XIX,

p.~116 (1917)) about the stationary states of systems which allow of what may be
called ``partial separation

of variables''. In this case it is possible to choose a set of positional coordinates


$q_{1}$,~\MyDots $q_{s}$ in such

a way that, for some of the coordinates, the conjugated momenta may be
considered as functions of

the corresponding $q$'s~only, so that, for these coordinates, quantities~$I$ may


be defined by~\EqI{(21)} in the

same way as for systems for which a complete separation of variables can be
obtained. From analogy

with the theory of the stationary states of the latter systems, \Name{Epstein}
proposes therefore the assumption,

that some of the conditions to be fulfilled in the stationary states of the systems in
question may

be obtained by putting the $I$'s~thus defined equal to entire multipla of~$h$. It


will be seen that, in case

of systems possessing an axis of symmetry, this leads to the second of the


conditions~\Eq{(61)}, which expresses

the condition that in the stationary states the total angular momentum round the
axis must

be equal to an entire multiple of~$h/2\pi$. As pointed out in Part~I on \


PageRefI[page]{34}, this condition would also

221
seem to obtain an independent support from considerations of conservation of
angular momentum

during a transition between two stationary states.}

\[

\frakI_{1} = \frakn_{1} h,\quad

\frakI_{2} = \frakn_{2} h,

\Tag{(61)}

\]

where $\frakn_{1}$~and~$\frakn_{2}$ are entire numbers, it will therefore be


seen that we obtain the right

relation between the additional energy $\frakE = \psi$ of the perturbed atom and
the frequencies

\PageSep{56}

of the secular perturbations of the orbit of the electron. It will moreover be

seen that a state in which the electron moves in a circular orbit perpendicular to

the axis of the field, and which beforehand must be expected to belong to the

stationary states of the perturbed atom since this orbit will not undergo secular

perturbations during a uniform establishment of the external field, will be included

among the states determined by~\Eq{(61)}. In fact, if $n$~is the number which
characterises

the corresponding stationary state of the undisturbed system, this state of the

perturbed system will correspond to $\frakn_{1} = 0$, $\frakn_{2} = n$ or to $\


frakn_{1} = n$, $\frakn_{2} = n$, according

to whether $\beta_{2}$~during the perturbations oscillates between fixed limits, or


increases

222
(or decreases) continuously. As regards the application of the conditions~\
Eq{(61)} it is

of importance to point out that, from considerations of the invariance of the a-


priori

probability of the stationary states of an atomic system during continuous


transformations

of the external conditions (see Part~I, \PageRefI[page]{9} and~\PageRefI[]{27}),


it seems necessary

to conclude that no stationary state exists corresponding to $\frakn_{2} = 0$. For


this value

of~$\frakn_{2}$ the motion of the electron would take place in a plane through the
axis,

but for certain external fields such motions cannot be regarded as physically

realisable stationary states of the atom, since in the course of the perturbations the

electron would collide with the nucleus (compare \PageRef[page]{68}).

A special case of an external field possessing axial symmetry, in which the

secular perturbations are very simple, presents itself if the external forces form

\Emph{a central field with the nucleus at the centre}. In this case the solution

of the problem of the fixation of the stationary states is given by \


Name{Sommerfeld}'s

general theory of central systems, discussed \Typo{i}{in} Part~I, which rests upon
the fact that

these systems allow of separation of variables in polar coordinates. In connection

with the above considerations it may be of interest, however, to consider the


problem

223
in question directly from the point of view of perturbed periodic systems, because

it presents a characteristic example of a \Emph{degenerate} perturbed system. In


the present

case $\psi$~will, besides on~$\alpha_{1}$, depend on $\alpha_{2}$~only, and


from the equations~\Eq{(46)} we get

therefore the well known result, that the angular momentum of the electron and the

plane of its orbit will not vary during the perturbations, and that the only secular

effect of the perturbing field will consist in a slow uniform rotation of the direction

of the major axis. For the frequency of this rotation we get from~\Eq{(46)}

\[

\frakv = \frac{1}{2\pi}\, \frac{D\beta_{2}}{Dt}

= \frac{1}{2\pi}\, \frac{\delta\psi}{\delta\alpha_{2}},

\Tag{(62)}

\]

from which we get directly for the difference between the values of~$\psi$ for two
neighbouring

states of the perturbed system, for which the corresponding value of~$I$ is the

same,

\[

\delta\psi = 2\pi \frakv\, \delta\alpha_{2}.

\Tag{(63)}

\]

This relation, which corresponds to~\Eq{(56)}, is seen to coincide with~\Eq{(59)},


since in the

224
present case $\frakv_{2} = 0$ and $\frakI_{1} = 2\pi\alpha_{2}$. From~\Eq{(63)}
it follows that the necessary relation

between the additional energy of the atom and the frequency of the perturbations

\PageSep{57}

is secured if the stationary states in the presence of a small external central

field are characterised by the condition\Pagelabel{57}

\[

\frakI = 2\pi \alpha_{2} = \frakn h,

\Tag{(64)}

\]

where $\frakn$~is an entire number. This condition, which is equivalent with the
second

of \Name{Sommerfeld}'s conditions~\EqI{(16)}, corresponds to~\Eq{(58)} and is


seen to coincide

with the first of the conditions~\Eq{(61)}, while the second of the latter conditions
in the

special case under consideration \Typo{looses}{loses} its validity corresponding


to the fact that

the orientation of the plane of the orbit in space is obviously arbitrary. Since, for

a Keplerian motion, the major axis of the orbit depends on the total energy only

while the minor axis is proportional to the angular momentum, it will be seen
from~\Eq{(64)}

that the presence of a small external field imposes the restriction on the motion

of the atom in the stationary states, that the minor axis of the orbit of the electron

225
must be equal to an entire multiple of the \Ord{$n$}{th}~part of the major axis,
which was

given by~$2\alpha_{n}$ in~\Eq{(41)}. This result has been pointed out by \


Name{Sommerfeld} as a consequence

of the application of the conditions~\EqI{(16)}.

\medskip

In the preceding it has been shown how it is possible to attack the problem

of the stationary states of a perturbed periodic system by an examination of the

secular perturbations of the shape and position of the orbit, and to fix these states

if the perturbations are of periodic or conditionally periodic type. While these


considerations

allow to determine the possible values for the total energy of the perturbed

system and thereby the \Emph{frequencies} of the components into which the

lines of the spectrum of the undisturbed system are split up in the presence of the

external field, it is necessary, however, for the discussion of the \Emph{intensities}


and

\Emph{polarisations} of these components to consider more closely the motion of


the

particles in the perturbed system and the relation of the total energy of this system

to the fundamental frequencies which characterise the motion. In the first place it

will be seen that, if the secular perturbations as determined by the equations~\


Eq{(46)}

are of conditionally periodic type, the displacements of the particles of the system

in any given direction may, with neglect of small quantities proportional to the

226
intensity of the external forces, be represented, within a time interval sufficiently

large for these forces to produce a considerable change in the shape and position

of the orbit, as a sum of harmonic vibrations by expressions of the type:

\[

\BreakMath{%

\xi = \sum C_{\tau, \frakt_{1}, \MyDots \frakt_{s-1}} \cos 2\pi \bigl\{(\tau\


omega_{P} + \frakt_{1} \frakv_{1} + \dots \Add{+} \frakt_{s-1} \frakv_{s-1})t

+ c_{\tau, \frakt_{1}, \MyDots \frakt_{s-1}}\bigr\},

}{%

\MathBack

\begin{split}

\xi = \sum C_{\tau, \frakt_{1}, \MyDots \frakt_{s-1}} \cos 2\pi \bigl\{(\tau\


omega_{P} + \frakt_{1} \frakv_{1} + \cdots \\

{} \Add{+} \frakt_{s-1} \frakv_{s-1})t

+ c_{\tau, \frakt_{1}, \MyDots \frakt_{s-1}}\bigr\},

\end{split}

\Tag{(65)}

\]

where the summation is to be extended over all positive and negative entire values

of $\tau$,~$\frakt_{1}$,~\MyDots $\frakt_{s-1}$, and where the $C$'s~and~$c$'s


are two sets of constants, the former of

which depend only on the values of the quantities $\frakI_{1}$,~\MyDots $\


frakI_{s-1}$ defined by~\Eq{(53)} and

227
on the value of the quantity~$I$, which characterises the corresponding state of the

undisturbed system which would appear if the external field vanished at a slow and

uniform rate. While the quantities $\frakv_{1}$,~\MyDots $\frakv_{s-1}$ are the


same as those which appear

\PageSep{58}

in the formula~\Eq{(54)}, and represent the small frequencies of the secular


perturbations

of the shape and position of the orbit, the quantity~$\omega_{P}$ may be


considered as

representing the mean frequency of revolution of the particles in their


approximately

periodic orbit. As regards the total energy of the perturbed system, it may next be

proved that, looking apart from small quantities proportional to the square of the

intensity of the external forces, the difference in the total energy in two slightly

different states of the perturbed system, for which the values of $I$, $\
frakI_{1}$,~\MyDots $I_{s-1}$ differ

by $\delta I$, $\delta \frakI_{1}$,~\MyDots $\delta \frakI_{s-1}$ respectively, is


given by the relation\Typo{.}{}\Pagelabel{58}\footnotemark

\[

\delta E = \omega_{P}\, \delta I

+ \sum_{1}^{s-1} \frakv_{k}\, \delta \frakI_{k},

\Tag{(66)}

\]

\footnotetext{From a comparison with formula~\EqI{(8)}, holding for the energy


difference between two neighbouring

228
states of the undisturbed system, and with formula~\Eq{(52)}, it will be seen
that \Eq{(66)}~implies the condition

$\omega_{P} = \omega + \dd\psi/\dd I$, where $\omega$~is the frequency of


revolution in the corresponding state of the undisturbed

system characterised by the given value of~$I$, and where, in the partial
differential coefficient, $\psi$~is considered

as a function of~$I$ and $\frakI_{1}$,~\MyDots $\frakI_{s-1}$. This relation can


be verified by means of a consideration

based on the perturbation equations~\Eq{(44)}, which takes into account the


simple relation between $\alpha_{1}$~and~$I$

for the undisturbed system, as well as the relation between the mean rate of
variation of~$\beta_{1}$ with

the time and the difference between $\omega_{P}$~and~$\omega$. We shall not


enter, however, on the details of the

rather intricate calculations involved in such a consideration, since the problems


in question allow of

a more elegant treatment by means of another analytical method. Thus it will be


shown by Mr.~\Name{H.~A.

Kramers}, in the paper mentioned in the end of~\SecRef{4}, that, quite


independent of the possibility of separation

of variables for the perturbed system in a fixed set of positional coordinates, the
theory of secular

perturbations exposed in this section offers---if these perturbations as determined


by~\Eq{(46)} are of conditionally

periodic type---a means of disclosing a set of \Emph{angle variables}, which may


be used to

describe the motion of the perturbed system with the same degree of
approximation as that involved
229
in the preceding calculations. According to the definition of angle variables,
mentioned in the Note on

\PageRefI[page]{29} in Part~I, this means that it is possible, in stead of the


positional coordinates $q_{1}$,~\MyDots $q_{s}$ of the

perturbed system and their conjugated momenta $p_{1}$,~\MyDots $p_{s}$ to


introduce a new set of $s$~variables in

such a way, that the $q$'s~and~$p$'s are periodic in every of the new variables
with period~$1$, when they are

considered as functions of these variables and of their canonically conjugated


momenta. These momenta

will just coincide with the quantities denoted above by $I$, $\frakI_{1}$,~\
MyDots $\frakI_{s-1}$, and the corresponding angle

variables may conveniently be denoted by $w$, $\frakw_{1}$,~\MyDots $\


frakw_{s-1}$ respectively. Introducing the new variables,

the total energy of the perturbed system will be a function of~$I$, $\frakI_{1}$,~\
MyDots $\frakI_{s-1}$ only, if we look apart

from small quantities proportional to~$\lambda^{2}$. With this approximation


we get consequently by a calculation,

analogous to that given in the Note referred to, that the angle variables $w$, $\
frakw_{1}$,~\MyDots $\frakw_{s-1}$ may be represented

as linear functions of the time within an interval of the same order as~$\sigma/\
lambda$. Denoting the rates

of variation of $w$, $\frakw_{1}$,~\MyDots $\frakw_{s-1}$ by $\omega$, $\


frakv_{1}$,~\MyDots $\frakv_{s-1}$ respectively, the \Chg{formulae}
{formulæ} \Eq{(65)} and \Eq{(66)} are therefore

directly obtained, just as the corresponding \Chg{formulae}{formulæ} \


EqI{(31)} and \EqI{(29)} in Part~I\@. In this connection it

230
will be observed that, due to the possibility of introduction of angle variables, the
conditions~\Eq{(67)} appear

in the same form as that in which the conditions, which fix the stationary states of
ordinary conditionally

periodic systems which allow of separation of variables, have been formulated


by \Name{Schwarzschild},

and which, as mentioned in the Note in Part~I, has already been applied by \
Name{Burgers} to certain systems

for which such a separation cannot be obtained.}

which coincides with~\Eq{(52)} if $\delta \frakI = 0$, and which will be seen to be
completely

analogous with formula~\EqI{(29)} in Part~I, holding for an ordinary


conditionally periodic

system which allows of separation of variables in a fixed set of positional


coordinates;

just as \Eq{(65)}~is analogous to formula~\EqI{(31)} representing the


displacements of the particles

\PageSep{59}

for such a system. Since moreover, in complete analogy to the conditions~\


EqI{(22)},

the stationary states of the perturbed system are characterised by\Pagelabel{59}

\[

I = nh,\quad

\frakI_{k} = \frakn_{k} h,\qquad

(k = 1, \MyDots s - 1)

\Tag{(67)}

231
\]

we see consequently that, for sufficiently small intensity of the external forces, we

obtain in the region of large values of~$n$ and of the~$\frakn$'s a connection


between the

frequencies of the components of the spectral lines, determined on the quantum


theory

by means of relation~\EqI{(1)}, and those to be expected on ordinary


electrodynamics,

which is of exactly the same type as the analogous connection, discussed in Part~I,

in case of ordinary conditionally periodic systems which allow of separation of

variables. In perfect analogy with the general considerations in Part~I, we are

therefore led directly to certain simple conclusions as regards the intensities and

polarisations of the components into which the lines of the spectrum of the
undisturbed

periodic system are split up in the presence of the external field. Thus we

shall expect that there will exist an intimate connection between the probability of

spontaneous transition between two stationary states of the perturbed system, for

which $n = n'$, $\frakn_{k} = \frakn_{k}'$ and $n = n''$, $\frakn_{k} = \


frakn_{k}''$ respectively, and the values in these

states of the coefficient $C_{\tau, \frakt_{1}, \MyDots \frakt_{s-1}}$ in the


expressions for the displacements of the

particles, for which $\tau = n' - n''$ and $\frakt_{k} = \frakn_{k}' - \frakn_{k}''$. If
for instance, for a certain set

of values of~$\tau$ and $\frakt_{1}$,~\MyDots $\frakt_{s-1}$, the coefficient


$C_{\tau, \frakt_{1}, \MyDots \frakt_{s-1}}$ in the expressions for the

232
displacements in every direction will be equal to zero for all motions of the
perturbed

system, we shall expect that the corresponding transitions between two

stationary states will be impossible in the presence of the given external field; and

if this coefficient is zero for the displacements of the particles in a certain direction

only, we shall expect that the corresponding transitions will give rise to the

emission of a radiation which is polarised in a plane perpendicular to this direction.

With a characteristic example of these considerations we meet in the case of

the spectrum of a hydrogen atom exposed to an external field of force which


possesses

axial symmetry round an axis through the nucleus. In analogy with the

resolution of the motion of an ordinary conditionally periodic system which


possesses

an axis of symmetry in its constituent harmonic vibrations, discussed in

Part~I on \PageRefI[page]{33}, it follows from the discussion of the general


character of the

secular perturbations on \PageRef[page]{54} that the motion of the electron in the


perturbed atom

in this case can be resolved in a number of linear harmonic vibrations parallel to

the axis with frequencies $\tau\omega_{P} + \frakt_{1} \frakv_{1}$ and in a


number of circular harmonic rotations

perpendicular to the axis with frequencies $\tau\omega_{P} + \frakt_{1}\


frakv_{1} + \frakv_{2}$. In complete analogy with

the considerations in Part~I, we are therefore led to conclude that in the present

233
case only two types of transitions between the stationary states of the perturbed

atom are possible. In the transitions of the first type $\frakn_{2}$~will remain
unaltered and

the emitted radiation will give rise to components of the hydrogen lines which will

show linear polarisation parallel to the axis. In the transitions of the second type

$\frakn_{2}$~will change by one unit and the emitted radiation will show circular
polarisation

when viewed in the direction of the axis. Remembering that, according to

\PageSep{60}

the conditions~\Eq{(61)}, the angular momentum of the system round the axis in
the

stationary states is equal to~$\frakn_{2}\, \dfrac{h}{2\pi}$, it will be seen


moreover that, also in the present

case, these conclusions obtain an independent support from a consideration of


conservation

%[** TN: Footnote mark before punctuation in the original.]

of angular momentum during the transitions (Compare Part~I \PageRefI[page]


{34}).\footnote

{\Emph{Note added during the proof.} In an interesting paper by \


Name{A.~Rubinowicz} (Phys.\ Zeitschr.\

XIX, p.~441 and p.~465 (1918)) which has just been published, a similar
consideration of conservation

of angular momentum has been used to draw conclusions, as regards the


possibility of transitions

between the stationary states of a conditionally periodic system possessing an axis


of symmetry, and

234
as regards the character of the polarisation of the radiation accompanying these
transitions. In this way

\Name{Rubinowicz} has arrived at several of the results discussed in the present


paper; in this connection,

however, it may be remarked that, from a consideration of conservation of


angular momentum, it is

not possible, even for systems possessing axial symmetry, to obtain as complete
information, as regards

the number and polarisation of the possible components, as from a consideration


based on the resolution

of the motion of the electron in harmonic vibrations.}

In the following we will meet with applications of these considerations when


discussing

the effect of electric and magnetic fields on the hydrogen lines. In the latter

case, however, the preceding considerations need some modifications due to the

fact, that the external forces acting on the electron cannot be derived from a

potential expressed as a function of its positional coordinates; to this point we shall

come back in~\SecRef{5}.

Before\Pagelabel{60} leaving the general theory of perturbed periodic systems we


shall still

consider the problem of the effect on the spectrum of a periodic system,


undergoing

secular perturbations of conditionally periodic type under the influence of a

given small external field, if this system is further subject to \Emph{the influence
of

235
a second external field which is small compared with the first field},

but the perturbing effect of which is yet large compared with the small effects

on the motion, proportional to the square of the intensity of the first perturbing

field, which were neglected in the preceding calculations. This problem is

closely analogous to the problem, briefly discussed in Part~I, of the effect of a

small perturbing field on the spectrum of an ordinary conditionally periodic

system which allows of separation of variables. As mentioned on \PageRefI[page]


{34}, we have

in this case, quite independent of the possibility of separation of variables for the

perturbed system, that in general the motion under the influence of the external

field may still be represented as a sum of harmonic vibrations by a formula of

the type~\EqI{(31)}, if we look apart from small terms proportional to the square
of the

perturbing forces. Corresponding to this we have in the case under consideration

that, independent of the nature of the second external field, the resultant secular

perturbations may in general be expressed as a sum of harmonic vibrations of small

frequencies of the type~\Eq{(54)}, if we look apart from small terms of the same
order

as the product of the secular perturbations produced by the first external field with

the square of the ratio between the intensities of the forces due to the first and

those due to the second external field. Let us denote this ratio by~$\mu$ and let, as

above, $\lambda$~represent a small constant of the same order as the ratio


between the external

\PageSep{61}

236
forces due to the first field and the internal forces of the system. On the

basis of the general relation between energy and frequency in the stationary states,

we may then expect that it is possible to fix the motion in these states for the
perturbed

periodic system in the presence of both external fields with neglect of small

terms of the same order as the largest of the quantities $\mu^{2}$~and~$\lambda$,


and to fix the

corresponding values for the energy with neglect of small terms of the same order

as the largest of the quantities $\lambda\mu^{2}$ and~$\lambda^{2}$.\


Pagelabel{61}\footnote

{In analogy with the considerations on \PageRef[page]{50} it may be expected,


however, that these limits

for the definition of the energy in the stationary states will hold only for the great
majority among

a large number of atomic systems. Thus in the present case we must be prepared
to find that for a small

fraction of the systems of the same order as~$\mu^{2}$ (if $\mu^{2} > \
lambda$) the energy will differ from that fixed by

the method under consideration by small quantities of the same order as~$\mu\
lambda$.}

In general, however, the effect on the

spectrum of the perturbed system, produced by the second external field, may be

calculated without considering the perturbing effect of this field in detail. In fact, it

is in general possible, by means of the principle of the mechanical transformability

of the stationary states, with the approximation mentioned to determine the


alteration

237
of the energy of the system, due to the presence of the second external

field, directly from the character of the secular perturbations produced by the first

external field only. Thus let us assume that the second field is slowly established

at a uniform rate within a time interval of the same order of magnitude as

that in which the system will pass approximately through any state belonging

to the cycle of shapes and positions, which the orbit passes through in the

stationary states in the presence of the first external field only. Denoting a time

interval of this order by~$\theta$ and the potential of the first perturbing field
by~$\Omega$ and

that of the second by~$\Delta\Omega$, we get then, by a calculation quite


analogous to that

given in Part~I on \PageRefI[page]{11} for the alteration in the mean value of the
energy of a

periodic system during a slow establishment of a small external field, that the

alteration in the mean value of $\alpha_{1} + \Omega$ taken over a time interval
of the same

order as~$\theta$, due to the establishment of the second external field, will be a

small quantity of the same order of magnitude as~$\theta(\Delta\Omega)^{2}$; but


with the notation

used above this means, in general, a small quantity of the same order as~$\lambda\
mu^{2}$. It

follows consequently that, with this approximation, the alteration in the energy in

a given stationary state, due to the presence of the second perturbing field, is equal

to the mean value of the potential of this field taken over the cycle of shapes and

positions, which the orbit would pass through in the corresponding stationary state

238
of the perturbed system under the influence of the first external field only. In

general, the effect on the spectrum will therefore consist in a small displacement of

the original components proportional to the intensity of the forces due to the
second

perturbing field; and as regards the degree of approximation with which these
displacements

are defined, it will be seen from the above that, if $\mu$~is smaller than~$\sqrt{\
lambda}$,

the fixation of the energy in the stationary states in the presence of the second
external

field, and therefore also the determination of the frequencies of the spectral

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem, Afd. 8. Række, IV. 1. 9

\PageSep{62}

lines by means of~\EqI{(1)}, allow of the same degree of approximation as the


fixation

of the energy in the stationary states of the original perturbed periodic system.

If $\mu$~is larger than~$\sqrt{\lambda}$, however, the stationary states will in


general not be as well

defined as for the original system, and from relation~\EqI{(1)} we may therefore
expect

that the components will be diffuse, although, as long as $\mu$~remains small


compared

with unity, the width of the components will remain small compared with the

displacements from their positions in the presence of the first external field alone.

239
Only when $\mu$~becomes of the same order as unity, the simultaneous effect of
both

perturbing fields may be expected to consist in a diffusion of the lines of the


undisturbed

periodic system;\Pagelabel{62} unless of course the secular perturbations due to


the

simultaneous presence of both fields are still of conditionally periodic type, as it

may happen in special problems. In certain cases the second external field will not

only give rise to small displacements of the original components but also to the

appearance of new components of small intensities proportional to~$\mu^{2}$.


This occurs

if for the original perturbed periodic system, due to some \Typo{pecularity}


{peculiarity} of the motion,

some of the coefficients $C_{\tau,\frakt_{1}, \MyDots \frakt_{s-1}}$ in the


expressions~\Eq{(65)} for the displacements

of the particles as a sum of harmonic vibrations, corresponding to certain


combinations

of the numbers $\tau$, $\frakt_{1}$,~\MyDots $\frakt_{s-1}$, are equal to zero,


while in the presence

of the second external field these coefficients are small quantities proportional
to~$\mu$

(\Typo{Compare}{compare} Part~I, \PageRefI[page]{34}).\footnote

{As regards the degree of definition with which the positions of the new
components will be

determined, we must be prepared to find that the frequencies of these components


are only defined

240
with neglect of small quantities proportional to~$\lambda\mu$. Compare the
detailed discussion of the example

in~\SecRef{5} on \PageRef[page]{97}.}

In the preceding considerations it has been assumed

that the perturbed system in the presence of the first external field is non-
degenerate.

In case, however, this system is \Emph{degenerate}, it is obviously impossible,

by a direct application of the principle of the mechanical transformability of the

stationary states, to determine the alteration in the energy in the stationary states

of the system, which will be due to the presence of a second external field small

compared with the first field; because, as mentioned, the stationary states of the

system, in the presence of this field only, will be determined by a number of


conditions

which is less than the number~$s$ of degrees of freedom, and that consequently

the cycles of shapes and positions, which the orbit will pass through in these states,

will not be completely determined. For the calculation of the energy in the \
Typo{stationay}{stationary}

states it will therefore be necessary to consider the secular perturbing effect of the

second external field on these cycles. In the special case where the secular
perturbations

due to the first field are simply periodic, it wall in this way be seen that the

problem of the fixation of the stationary states in the presence of the second
external

field, by means of the method exposed in this section, may be reduced to

241
the problem of the fixation of the stationary states of a system of $s - 2$~degrees
of

freedom. If, as in the applications considered below, $s$~is equal to~$3$, this
problem

allows of a general solution, and we must therefore expect that in this case the

\PageSep{63}

effect on the spectrum of the perturbed system produced by an arbitrary second

external field, which is small compared with the first, will consist in the splitting

up of every component into a number of separate components, just as the effect of

an arbitrary small external field on the lines of the spectrum of a simple periodic

system of two degrees of freedom. We will meet with applications of the above

considerations when considering the effect on the hydrogen spectrum of the


combined

action of different external fields and when considering the effect of an external

field on the spectra of other elements, which latter problem will be discussed

in Part~III\@.

\Section{3.}{The fine structure of the~hydrogen~lines.}

An instructive application of the calculations in the last section may be made

in connection with the fine structure of the hydrogen lines, which, according to

\Name{Sommerfeld}'s theory mentioned in Part~I on \PageRefI[page]{18}, may


be explained by taking into

242
account the small variation of the mass of the electron with its velocity, claimed by

the theory of relativity. In this connection it must first of all be remarked that all

the general considerations in the preceding sections, as regards relations between

energy and frequency and as regards the mechanical transformability of the


stationary

states, hold unaltered if the relativity modifications are taken into account.

This follows from the fact that the Hamiltonian equations~\EqI{(4)}, which are
taken

as a basis for all the previous calculations, may be used to describe the motion

also in this case. If, when the relativity modifications are taken into account, the

motion of the system is simply periodic independent of the initial conditions, we

shall consequently expect that the stationary states are characterised by the
condition

$I = nh$ only, and that the energy and frequency are the same for all states

corresponding to a given value of~$n$ in this equation. Further the stationary


states

will also in the relativity case be fixed by~\EqI{(22)}, if the system is


conditionally

periodic and allows of separation of variables; while the stationary states of a

perturbed periodic system, also in the relativity case, will be characterised by the

conditions~\Eq{(67)}, if the secular perturbations are of conditionally periodic


type.

Now, when the relativity modifications are taken into account, the motion of

243
the particles in the hydrogen atom will not, as assumed in~\SecRef{1}, be exactly
periodic, but

the orbit of the electron will be of the same type as that, which would appear on

ordinary Newtonian mechanics, if the law of attraction between the particles


differed

slightly from that of the inverse square. If, for the moment, we consider the mass

of the nucleus as infinite, the system will allow of a separation of variables in polar

coordinates, and the stationary states may consequently be fixed by the


conditions~\EqI{(16)}.

In this way \Name{Sommerfeld} obtained an expression for the total energy in the

stationary states, which, with neglect of small quantities of higher order than the

\PageSep{64}

square of the ratio of the velocity of the electron and the velocity of light~$c$, is

given by\footnote

{\Name{A.~Sommerfeld}, Ann.\ d.\ Phys.\ LI, p.~53 (1916). Compare also \


Name{P. Debye}, Phys.\ Zeitschr., XVII, p.~512

(1916). In the special case of circular orbits ($n_{1} = 0$), this expression
coincides with an expression

previously deduced by the writer (Phil.\ Mag.\ XXIX p.~332 (1915)), by a direct
application of the condition

$I = nh$ to these periodic motions.}

\[

\llap{$E$} = -\frac{2\pi^{2} N^{2} e^{4}m}{h^{2} (n_{1} + n_{2})^{2}}

\left[1 + \frac{\pi^{2} N^{2} e^{4}}{c^{2} h^{2} (n_{1} + n_{2})^{2}}

\left(1 + 4\frac{n_{1}}{n_{2}}\right)\right]\!,

244
\Tag{(68)}

\]

where, as in the calculations in~\SecRef{1}, the charge and the mass of the
electron are

denoted by $-e$ and~$m$, and for sake of generality the charge of the nucleus
by~$Ne$.

Further $n_{1}$~and~$n_{2}$ are the integers appearing on the right side of the
conditions~\EqI{(16)}

as factors to \Name{Planck}'s constant. While $n_{1}$~may take the values


$0$,~$1$, $2$,~\dots, it

will be seen that $n_{2}$~can only take the values $1$,~$2$,~\dots, because in the
present case

there will obviously not correspond any stationary state to $n_{2} = 0$, since in
such

a state the electron would collide with the nucleus. Introducing the experimental

values for $e$,~$h$ and~$c$, it is found that $e^{2}/hc$ is a small quantity of the
same order as~$10^{-3}$;

and, unless $N$~is large number, the second term within the bracket on the

right side of~\Eq{(68)} will consequently be very small compared with unity.
Putting

$n_{1} + n_{2} = n$, it will further be seen that the factor outside the bracket will
coincide

with the expression for~$W_{n}$ given by~\Eq{(41)} in~\SecRef{1}, if we look


apart from the

small correction due to the finite mass of the nucleus. Due to the presence of the

second term within the bracket, we thus see that, for any value of~$n$, formula~\
Eq{(68)} gives
245
a set of values for~$E$ which differ slightly from each other and from~$-W_{n}
$. \Name{Sommerfeld}'s

theory leads therefore to a direct explanation of the fact, that the hydrogen

lines, when observed by instruments of high dispersive power, are split up in a

number of components situated closely to each other; and, by means of formula~\


Eq{(68)}

in connection with relation~\EqI{(1)}, it was actually found possible, within the


limits

of experimental errors, to account for the frequencies of the components of this

\Typo{socalled}{so called} \Emph{fine structure} of the hydrogen lines.


Moreover the theory was supported

in the most striking way by \Name{Paschen}'s\footnote

{\Name{F.~Paschen}, Ann.\ d.\ Phys.\ L, p.~901 (1916). See also \Name{E.~J.


Evans} and \Name{C.~Croxson}, Nature, XCVII,

p.~56 (1916).}

recent investigation of the fine structure

of the lines of the analogous helium spectrum, the frequencies of which are

represented approximately by formula~\Eq{(35)}, if in the expression for~$K$,


given by~\Eq{(40)},

we put $N = 2$. As it should be expected from~\Eq{(68)}, the components of


these lines

were found to show frequency differences several times larger than those of the

hydrogen lines, and from his measurements \Name{Paschen} concluded, that it


was possible

on \Name{Sommerfeld}'s theory to account completely for the frequencies of all


the components

246
observed.

We shall not enter here on the details of the calculation leading to~\Eq{(68)}, but

shall only show how this formula may be simply interpreted from the point of

\PageSep{65}

view of perturbed periodic systems. Thus, by a simple application of relativistic

mechanics, it is found that, if the equation of a Keplerian ellipse in polar


coordinates

is given by $r = f(\theta)$, the equation of the orbit of the electron in the

case under consideration will be given by $r = f(\gamma\theta)$ where $\


gamma$~is a constant given

by $\gamma^{2} = 1 - \left(\dfrac{Ne^{2}}{pc}\right)^{2}$, in which


expression~$p$ denotes the angular momentum of the

electron round the nucleus.\footnote

{See f.\ inst.\ \Name{A.~Sommerfeld}, loc.\ cit.\ p.~47.}

Now in the stationary states the quantity in the

bracket, which is of the same order of magnitude as the ratio between the velocity

of the electron and the velocity of light, will be very small, unless $N$~is a large

number, and it will therefore be seen that the orbit of the electron can be described

as a periodic orbit on which a slow uniform rotation is superposed. Denoting the

frequency of revolution in the periodic orbit by~$\omega$ and the frequency of the
superposed

rotation by~$\frakv_{R}$, we have, with neglect of small quantities of higher


order

247
than the square of the ratio between the velocity of the electron and the velocity

of light,

\[

\frakv_{R} = \omega(1 - \gamma) = \frac{1}{2} \omega \left(\frac{Ne^{2}}{pc}\


right)^{2}.

\Tag{(69)}

\]

Comparing this formula with equation~\Eq{(62)} and remembering that, with the
approximation

in question, $p$~may be replaced by the quantity denoted in~\SecRef{2} by~$\


alpha_{2}$, we

see that the frequency of the secular rotation of the orbit will be the same as

that which would appear, if the variation of the mass of the electron was neglected,

but if the atom was subject to a small external central force the mean value of

the potential of which, taken over a revolution of the electron, was equal to

\[

\psi = -\omega \frac{\pi N^{2}e^{4}}{c^{2} \alpha_{2}}.

\Tag{(70)}

\]

This is simply shown, however, to be equal to the expression for~$\psi$


corresponding

to a small attractive force varying as the inverse cube of the distance. In fact, let

the potential of such a force be given by $\Omega = C/r^{2}$, where $C$~is a


constant and $r$~the

248
length of the radius vector from the nucleus to the electron. By means of

the relation $\alpha_{2} = mr^{2} \dot{\theta}$, where $\theta$~is the angular


distance of the radius vector from

a fixed line in the plane of the orbit, we get then

\[

\psi = \frac{1}{\sigma} \int_{0}^{\sigma} \frac{C}{r^{2}}\, dt

= \frac{\omega mC}{\alpha_{2}} \int_{0}^{2\pi} d\theta

= \frac{2\pi \omega mC}{\alpha_{2}},

\]

which expression is seen to coincide with~\Eq{(70)}, if $C = -\dfrac{N^{2}


e^{4}}{2c^{2}m}$.

If the relativity modifications are taken into account, and if for a moment

we would imagine that the nucleus, in addition to its usual attraction, exerted

\PageSep{66}

a small repulsion on the electron, proportional to the inverse cube of the distance

and equal and opposite to the attraction just mentioned, we would therefore obtain

a system for which, with neglect of small quantities of higher order than

the square of the ratio between the velocity of the electron and the velocity of

light, every orbit would be periodic independent of the initial conditions, and for

which consequently the stationary states would be fixed by the single condition

$I = nh$.\Pagelabel{66} Now the actual hydrogen atom may obviously be


considered as a \Emph{perturbed

system}, formed by this periodic system, when it is exposed to a small

249
central field for which the value of~$\psi$ is given by~\Eq{(70)}. With the
approximation

mentioned, we get therefore for the total energy in the stationary states of the atom

\[

E = E_{n}' - \frac{8\pi^{4} N^{4} e^{8}m}{h^{4} c^{2}}\, \frac{1}{n^{3} \


frakn},

\Tag{(71)}

\]

where $E_{n}'$~is the energy in the stationary states of the periodic system just
mentioned,

and where the last term is obtained by introducing in~\Eq{(70)} the value of~$\
alpha_{2}$ given

by~\Eq{(64)} and the value of~$\omega_{n}$ given by~\Eq{(41)}, neglecting the


small correction due to

the finite mass of the nucleus. Remembering that in our notation $n_{1} + n_{2} =
n$

and $n_{2} = \frakn$, it will be seen that, as regards the small differences in the
energy of

the different stationary states corresponding to the same value of~$n$, formula~\
Eq{(71)}

gives the same result as \Name{Sommerfeld}'s formula~\Eq{(68)}. In fact,


comparing \Eq{(68)}~and~\Eq{(71)},

we get

\[

E_{n}' = -\frac{2\pi^{2} N^{2} e^{4} m}{h^{2} n^{2}}

\left(1 - \frac{3\pi^{2} N^{2} e^{4}}{c^{2} h^{2} n^{2}}\right),

250
\Tag{(72)}

\]

which is seen to be a function of $n$~only. This expression might also have been
deduced

directly from the condition $I = nh$ by considering, for instance, a circular orbit,

in which case the calculation can be very simply performed.

In connection with the above calculations, it may be remembered that the fixation

of the stationary states, leading to the formulæ \Eq{(68)} or~\Eq{(71)}, is based on


the assumption,

that the motion of the electron can be determined as that of a mass point

which moves in a conservative field of force, according to the laws of ordinary

relativistic mechanics, and that we have looked apart from all such forces which,

according to the ordinary theory of electrodynamics, would act on an accellerated\


Note{[sic]}

charged particle, and which constitute the reaction from the radiation which on this

theory would accompany the motion of the electron. Some procedure of this kind,

which means a radical departure from the ordinary theory of electrodynamics, is

obviously necessary in the quantum theory in order to avoid dissipation of energy

in the stationary states. Since we are entirely ignorant as regards the mechanism

of radiation, we must be prepared, however, to find that the above treatment will

allow to determine the motion in the stationary states, only with an approximation

which looks apart from small quantities of the same order as the ratio between

251
the radiation forces in ordinary electrodynamics and the main forces on the
electron

\PageSep{67}

due to the attraction from the nucleus.\footnote

{Compare Part~I, \PageRefI[p.]{6}. It may in this connection be noted that the


degree of approximation, involved

in the determination of the frequencies of an atomic system by means of


relation~\EqI{(1)} if in the fixation

of the stationary states we look apart from small forces of the same order of
magnitude as the radiation

forces in ordinary electrodynamics, would appear to be intimately connected with


\Emph{the limit of

sharpness of the spectral lines}, which depends on the total number of waves
contained in the

radiation emitted during the transition between two stationary states. In fact, from
a consideration

based on the general connection between the quantum theory and the ordinary
theory of radiation, it

seems natural to assume that the rate, at which radiation is emitted during a
transition between two

stationary states, is of the same order of magnitude as the rate, at which radiation
would be emitted

from the system in these states according to ordinary electrodynamics. But this
will be seen to imply

that the total number of waves in question will just be of the same order as the
ratio between the

252
main forces acting on the particles of the system and the reaction from the
radiation in ordinary

electrodynamics.}

Now it is easily shown that this ratio

will be a small quantity of the same order of magnitude as $N^{2}\left(\


dfrac{e^{2}}{pc}\right)^{3}$, and it would

therefore beforehand seem justified in the expression for the total energy in the

stationary states to retain small terms of the same order as the second term in~\
Eq{(71)},

while at the same time it might appear highly questionable, whether, in the

complete expression for the total energy in the stationary states deduced by \
Name{Sommerfeld}

and \Name{Debye} on the basis of the conditions~\EqI{(16)}, it has a physical


meaning to

retain terms of higher order than those retained in formula~\Eq{(68)}; unless


$N$~is a large

number, as in the theory of the Röntgenspectra to be discussed in Part~III\@.

While the preceding considerations, which deal with the determination of the

energy in the stationary states of the hydrogen atom, allow to determine \Emph{the
frequency

of the radiation} which would be emitted during a transition between

two such states, they leave quite untouched the problem of the actual occurrence

of these transitions in the luminous gas, and therefore give no direct information

about the \Emph{number and relative intensities of the components} into which

253
the hydrogen lines may be expected to split up as a consequence of the relativity

modifications. This problem has recently been discussed by \Name{Sommerfeld},\


footnote

{\Name{A.~Sommerfeld}, Ber.\ Akad.\ München, 1917, p.~83.}

who in

this connection emphasises the importance of the different a-priori probabilities of

the stationary states, characterised by different sets of values of the~$n$'s in the


conditions~\EqI{(16)}.

Thus \Name{Sommerfeld} attempts to obtain a measure for the relative intensities

of the components of the fine structure of a given line, by comparing the

intensities observed with the products of the values of the a-priori probabilities of

the two states, involved in the emission of the components under consideration;
and he

tries in this connection to test different expressions for these a-priori probabilities

(See Part~I, \PageRefI[\Chg{pag.}{page}]{26}). In this way, however, it was not


found possible to account

in a satisfactory manner for the observations; and the difficulty in obtaining an


explanation

of the intensities on this basis was also strikingly brought out by the

fact, that the number and relative intensities of the components observed varied in

a remarkable way with the experimental conditions under which the lines were

\PageSep{68}

excited. Thus \Name{Paschen} found a greater number of components in the fine


structure

of the helium lines, mentioned above, when the gas was subject to a condensed

254
interrupted discharge, than when a continuous voltage was applied. It would seem,
however,

that all the facts observed obtain a simple interpretation on the basis of the

general considerations about the relation between the quantum theory of line
spectra

and the ordinary theory of radiation discussed in Part~I\@. According to this


relation,

we shall assume that the probability, for a transition between two given stationary

states to take place, will depend not only on the a-priori probability of these states,

which is determining for their occurrence in a distribution of statistical


equilibrium,

but will also depend essentially on the motion of the particles in these states,

characterised by the harmonic vibrations in which this motion can be resolved.

Now, in the absence of external forces, the motion of the electron in the hydrogen

atom forms a special simple case of the motion of a conditionally periodic system

possessing an axis of symmetry, and may therefore be represented by


trigonometric

series of the type deduced for such motions in Part~I\@. Taking a line through the

nucleus perpendicular to the plane of the orbit as $z$-axis, we get from the
calculations

on \PageRefI[page]{32}

\[

z = \text{const.}

\]

and

255
\[

\BreakMath{

x = \sum C_{\tau} \cos 2\pi \bigl\{(\tau\omega_{1} + \omega_{2})t + c_{\tau}\


bigr\},\quad

±y = \sum C_{\tau} \sin 2\pi \bigl\{(\tau\omega_{1} + \omega_{2})t + c_{\tau}\


bigr\},

}{

\begin{aligned}

x &= \sum C_{\tau} \cos 2\pi \bigl\{(\tau\omega_{1} + \omega_{2})t + c_{\tau}\


bigr\}, \\

±y &= \sum C_{\tau} \sin 2\pi \bigl\{(\tau\omega_{1} + \omega_{2})t + c_{\tau}\


bigr\},

\end{aligned}

\Tag{(73)}

\]

where $\omega_{1}$~is the frequency of the radial motion and $\omega_{2}$~is


the mean frequency of

revolution, and where the summation is to be extended over all positive and
negative

entire values of~$\tau$. It will thus be seen that the motion may be considered as a

superposition of a number of circular harmonic vibrations, for which the direction

of rotation is the same as, or the opposite of, that of the revolution of the electron

round the nucleus, according as the expression $\tau \omega_{1} + \omega_{2}$ is


positive or negative

256
respectively. From\Pagelabel{68} the relation just mentioned between the quantum
theory of line

spectra and the ordinary theory of radiation, we shall therefore in the present case

expect that, if the atom is not disturbed by external forces, only such transitions

between stationary states will be possible, in which the plane of the orbit remains

unaltered, and in which the number~$n_{2}$ in the conditions~\EqI{(16)}


decreases or increases

by one unit; \ie\ where the angular momentum of the electron round the nucleus

decreases or increases by~$h/2\pi$. From the relation under consideration, we shall

further expect that there will be an intimate connection between the probability of

a spontaneous transition of this type between two stationary states, for which
$n_{1}$~is

equal to $n_{1}'$~and $n_{1}''$ respectively, and the intensity of the radiation of


frequency

$(n_{1}' - n_{1}'') \omega_{1} ± \omega_{2}$, which on ordinary


electrodynamics would be emitted by the

atom in these states, and which would depend on the value~$C_{\tau}$ of the
amplitude

of the harmonic rotation, corresponding to $\tau = ±(n_{1}' - n_{1}'')$, which


appears in the

motion of the electron. Without entering upon a closer examination of the


numerical

values of these amplitudes, it will directly be seen that the amplitudes of the

harmonic rotations, which have the same direction as the revolution of the electron,

\PageSep{69}

257
in general, are considerably larger than the amplitudes of the rotations in the
opposite

direction, and we shall accordingly expect that the probability of spontaneous

transition will in general be much larger for transitions, in which the angular

momentum decreases, than for transitions in which it increases. This expectation

is verified by \Name{Paschen}'s observations of the fine structure of the helium


lines,

which show that, for a given line, the components corresponding to the transitions

of the former kind are by far the strongest. On \Name{Paschen}'s photographs,


however,

especially in the case of the application of a condensed discharge to the

vacuum tube containing the gas, there appear, in addition to the main components

corresponding to transitions for which the angular momentum changes by~$h/2\


pi$,

a number of weaker components, corresponding to transitions for which the

angular momentum remains unchanged or changes by higher multipla of~$h/2\pi$.


This

fact obtains a simple interpretation on the considerations in Part~I on \


PageRefI[page]{34}

about the influence of small external forces on the spectrum of a conditionally

periodic system. Thus, in the presence of small perturbing forces, the motion will

generally not remain in a plane, and in the trigonometric series representing

the displacement of the electron in space, there will occur small terms
corresponding

258
to frequencies $(\tau_{1} \omega_{1} + \tau_{2} \omega_{2})$, where $\
tau_{2}$~may be different from one. In

the presence of such forces, we shall therefore expect that, in addition to the
regular

probabilities of the above mentioned main transitions, there will appear small

probabilities for other transitions.\footnote

{\Emph{Note added during the proof.} As remarked in Part~I, this consideration


obtains a striking

confirmation by the observation of the appearance of new series of lines in the


ordinary series spectra

of helium and other elements, when the atoms are exposed to an intense external
electric field. As it

will be discussed more closely in Part~III, it is possible in this way to account in


detail for the manifold

results, regarding the appearance of such series in the helium spectrum, which
have been published

quite recently by \Name{J.~Stark} (Ann.\ d.\ Phys.\ LVI, p.~577 (1918)) and by \
Name{G.~Liebert} (ibid.\ LVI, p.~589 and

p.~610 (1918)).}

A detailed discussion of these problems will be

given in a later paper by Mr.\ \Name{H.~A. Kramers},\Pagelabel{69} who on my


proposal has kindly

undertaken to examine the resolution of the motion of the electron in its constituent

harmonic vibrations more closely, and who has deduced \Typo{explicite}


{explicit} expressions for

the amplitudes of these vibrations, not only for the motion of the electron in the

259
undisturbed atom, but also for the perturbed motion in the presence of a small

external homogeneous electric field. As it will be shown by \Name{Kramers},


these calculations

allow to account in particulars\Note{[sic]} for the observations of the relative


intensities

of the components of the fine structure of the hydrogen lines and the analogous

helium lines, as well as for the characteristic way in which this phenomenon is

influenced by the variation of the experimental conditions.

\Section{4.}{The effect of an external electric~field on~the~hydrogen~lines.}

As mentioned in the introduction, a detailed theory of the characteristic effect

of an external homogeneous electric field on the hydrogen spectrum, discovered by

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8. Række, IV. 1.


10

\PageSep{70}

\Name{Stark}, has been given by \Name{Epstein} and \Name{Schwarzschild} on


the basis of the general

theory of conditionally periodic systems which allow of separation of variables.

Before we enter on the discussion of the results of the calculations of these authors,

we shall first, however, show how the problem may be treated in a simple way

260
by means of the considerations about perturbed periodic systems, developed in~\
SecRef{2}.

Consider an electron of mass~$m$ and charge~$-e$, rotating round a positive

nucleus of infinite mass and of charge~$Ne$, and subject to a homogeneous


electric field

of intensity~$F$, and let us for the present neglect the small effect of the relativity

modifications. Using rectangular coordinates, and taking the nucleus as origin and

the $z$-axis parallel to the external field, we get for the potential of the system

relative to the external field, omitting an arbitrary constant,

\[

\Omega = eFz.

\]

Calculating now the mean value of~$\Omega$ over a period~$\sigma$ of the


undisturbed

motion, we see at once, from considerations of symmetry, that this mean value~$\
psi$

will depend only on the component of the external electric force in the direction

of the major axis of the orbit. We have therefore

\[

\psi = eF \cos\phi\, \frac{1}{\sigma} \int_{0}^{\sigma} r \cos\theta\, dt,

\]

261
where $\phi$~is the angle between the $z$-axis and the major axis, taken in the
direction

from the nucleus to the aphelium, and where $r$~is the length of the radius-vector

from the nucleus to the electron, and $\theta$~the angle between this radius-vector
and

the major axis. By means of the well known equations for a Keplerian motion

\[

r \cos\theta = \alpha(\cos u + \eps),\quad

\frac{dt}{\sigma} = (1 + \eps \cos u)\frac{du}{2\pi},

\]

where $2\alpha$~is the major axis, $\eps$~the eccentricity and $u$~the \


Typo{socalled}{so called} eccentric anomaly,

this gives

\[

\BreakMath{

\psi = eF \cos\phi\, \frac{1}{2\pi} \int_{0}^{2\pi} \alpha(\cos u + \eps)(1 + \eps \


cos u)\, du

= \frac{3}{2} \eps\alpha eF \cos\phi.

}{%

\begin{aligned}

\psi &= eF \cos\phi\, \frac{1}{2\pi} \int_{0}^{2\pi} \alpha(\cos u + \eps)(1 + \eps \


cos u)\, du \\

&= \frac{3}{2} \eps\alpha eF \cos\phi.

\end{aligned}

262
}

\Tag{(74)}

\]

We see thus that $\psi$~is equal to the potential energy relative to the external

field, which the system would possess, if the electron was placed at a point,
situated

on the major axis of the ellipse and dividing the distance~$2\eps a$ between the
foci in

the ratio~$3:1$. This point may be denoted as the ``electrical centre'' of the orbit.

From the approximate constancy of~$\psi$ during the motion, proved in~\
SecRef{2}, it follows

therefore in the first place that, with neglect of small quantities of the same order

of magnitude as the ratio between the external force and the attraction from the

nucleus, \Emph{the electrical centre will during the perturbations of the

orbit remain in a fixed plane perpendicular to the direction of the

external force}. From the considerations in~\SecRef{2} it follows further, that the
total

\PageSep{71}

energy in the stationary states of the system in the presence of the field, with
neglect

of small quantities proportional to~$F^{2}$, will be equal to $E_{n} + \psi$,


where $E_{n}$~is the energy

of the hydrogen atom in its undisturbed stationary state. Since both $\eps$~and $\
cos\phi$ are

263
numerically smaller than one, we obtain therefore at once from~\Eq{(74)} a lower
and

an upper limit for the possible variations of the energy in the stationary states, due

to the field. Introducing from~\Eq{(41)} the values of $E_{n}$~and~$\alpha_{n}


$, and neglecting, here

as well as in the following calculations in this section, the small correction due

to the finite mass of the nucleus---not only in the expression for the additional

energy but, for the sake of brevity, also in the main term---we get for these limits

\[

E = -\frac{2\pi^{2} N^{2} e^{4} m}{h^{2} n^{2}}

± \frac{3h^{2} n^{2}}{8\pi^{2} Nem}\, F,

\Tag{(75)}

\]

which formula coincides with the expression previously deduced by the writer by

applying the condition $I = nh$ to the two (physically not realisable) limiting
cases,

corresponding to $z = 1$ and $\cos\phi = ±1$, in which the orbit remains periodic


in

the presence of the field.\footnote

{See \Name{N.~Bohr}, Phil.\ Mag.\ XXVII, p.~506 (1914) and XXX, p.~394
(1915). Compare also \Name{E.~Warburg},

Verh.\ d.\ D.\ Phys.\ Ges.\ XV, p.~1259 (1913), where it was pointed out, for the
first time, that the effect of

an electric field on the hydrogen lines to be expected on the quantum theory was
of the same order

264
of magnitude as the effect observed by \Name{Stark}.}

In order to obtain further information as to the values of the energy in the

stationary states in the presence of the field, it is necessary to consider more


closely

the variation of the orbit during the perturbations. Since the external forces possess

axial symmetry, the problem of the stationary states might be treated by means of

the procedure indicated in~\SecRef{2} on \PageRef[page]{55}. In the present


special case, however,

the stationary states of the atom may be very simply determined, due to the fact

that the secular perturbations are simply periodic independent of the initial shape

and position of the orbit, so that we are concerned with a \Emph{degenerate} case
of a

perturbed periodic system. This property of the perturbations follows already from

some calculations given by \Name{Schwarzschild}\footnote

{\Name{K.~Schwarzschild}, Verh.\ d.\ D.\ Phys.\ Ges.\ XVI, p.~20 (1914).}

in a previous attempt to explain the

\Name{Stark} effect of the hydrogen lines, without the help of the quantum theory,
by

means of a direct consideration of the harmonic vibrations into which the motion

may be resolved, according to the analytical theory of conditionally periodic

systems. Starting from the above result, that the electrical centre moves in a plane

perpendicular to the direction of the external field, the periodicity of the


perturbations

265
may also be proved in the following way, by means of a simple consideration

of the variation of the angular momentum of the electron round the nucleus, due

to the effect of the external electric force.

\begin{Remark}

Using again rectangular coordinates with the nucleus at the origin and the $z$-axis
parallel

to the direction of the electric force, and calling the coordinates of the electrical
centre $\xi$,~$\eta$,~$\zeta$,

we have according to formula~\Eq{(74)}

\PageSep{72}

\[

\xi^{2} + \eta^{2} + \zeta^{2}

= \left(\frac{3}{2} \eps\alpha\right)^{2},\qquad

\zeta = \text{const.}

\Tag[II.]{(1^{*})}

\]

Denoting the components parallel to the $x$~$y$ and $z$-axis of the angular
momentum of the

electron round the nucleus, considered as a vector, by $P_{x}$,~$P_{y}$


and~$P_{z}$, we have next

\[

P_{x}^{2} + P_{y}^{2} + P_{z}^{2} = (1 - \eps^{2})(2\pi m\alpha^{2}\


omega)^{2},\qquad

266
P_{z} = \text{const.}

\Tag[II.]{(2^{*})}

\]

Since the angular momentum is perpendicular to the plane of the orbit, we have
further

\[

\xi P_{x} + \eta P_{y} + \zeta P_{z} = 0.

\Tag[II.]{(3^{*})}

\]

Now we have for the mean values of the rates of variation of $P_{x}$~and $P_{y}
$ with the time

\[

\frac{DP_{x}}{Dt} = eF\eta,\quad

\frac{DP_{y}}{Dt} = -eF\xi.

\Tag[II.]{(4^{*})}

\]

From this we get, differentiating \Eq[II.]{(1^{*})}~and \Eq[II.]{(2^{*})} with


respect to the time, and remembering that $\alpha$~and~$\omega$

remain constant during the perturbations,

\[

\BreakMath{%

\xi\, \frac{D\xi}{Dt} + \eta\, \frac{D\eta}{Dt}

= -K^{2} \left(P_{x}\, \frac{DP_{x}}{Dt} + P_{y}\, \frac{DP_{y}}{Dt}\right)

= -eFK^{2} (\eta P_{x} - \xi P_{y}),

267
}{%

\begin{aligned}

\xi\, \frac{D\xi}{Dt} + \eta\, \frac{D\eta}{Dt}

&= -K^{2} \left(P_{x}\, \frac{DP_{x}}{Dt} + P_{y}\, \frac{DP_{y}}{Dt}\right)


\\

&= -eFK^{2} (\eta P_{x} - \xi P_{y}),

\end{aligned}

\Tag[II.]{(5^{*})}

\]

where

\[

K = \frac{3}{4\pi m\alpha\omega}.

\Tag[II.]{(6^{*})}

\]

On the other hand we have, differentiating~\Eq[II.]{(3^{*})} and introducing~\


Eq[II.]{(4^{*})},

\[

P_{x}\, \frac{DP_{x}}{Dt} + P_{y}\, \frac{DP_{y}}{Dt} = 0,

\]

which together with~\Eq[II.]{(5^{*})} gives

\[

\frac{D\xi}{Dt} = eFK^{2} P_{y},\quad

268
\frac{D\eta}{Dt} = -eFK^{2} P_{x},

\]

from which we get, by means of~\Eq[II.]{(4^{*})},

\[

\frac{D^{2} \xi}{Dt^{2}} = -e^{2} F^{2} K^{2} \xi,\quad

\frac{D^{2} \eta}{Dt^{2}} = -e^{2} F^{2} K^{2} \eta,

\]

the solution of which is

\[

\xi = \frakA \cos 2\pi(\frakv t + \fraka),\quad

\eta = \frakB \cos 2\pi(\frakv t + \frakb),

\Tag[II.]{(7^{*})}

\]

where $\frakA$,~$\fraka$,~$\frakB$ and~$\frakb$ are constants, and where,


introducing~\Eq[II.]{(6^{*})}, we have

\[

\frakv = \frac{eFK}{2\pi}

= \frac{3eF}{8\pi^{2} m\alpha\omega}.

\Tag[II.]{(8^{*})}

\]

\end{Remark}

\Emph{During the perturbations the electrical centre will thus perform

269
slow harmonic vibrations perpendicular to the direction of the

electric force}, with a frequency which is proportional to the intensity of the

electric field, but, for a given value of~$F$, quite independent of the initial shape
of

the orbit and its position relative to the direction of the field. For the value of

this frequency in the multitude of states of the perturbed system, for which the

mean value of the inner energy is equal to the energy~$E_{n}$ in a stationary state
of

the undisturbed system corresponding to a given value of~$n$, we get from the
above

calculation, introducing for $\alpha$~and~$\omega$ the values of $\alpha_{n}


$~and $\omega_{n}$ given by~\Eq{(41)},

\PageSep{73}

\[

\frakv_{F} = \frac{3hn}{8\pi^{2} Nem}\, F.

\Tag{(76)}

\]

Now\Pagelabel{73} from the periodic motion of the electrical centre we may


conclude that, in the

presence of the field, the system will be able to emit or absorb a radiation of

frequency~$\frakv_{F}$, and that accordingly the possible values of the additional


energy of

the system in the presence of the field will be given directly by \Name{Planck}'s
fundamental

formula~\EqI{(9)}, holding for the possible values of the total energy of a linear

270
harmonic vibrator, if in this formula $\omega$~is replaced by the above
frequency~$\frakv_{F}$. Since

further a circular orbit, perpendicular to the direction of the electric force, will not

undergo secular perturbations during a slow establishment of the field, and


therefore

must be included among the stationary states of the perturbed system, we get

for the total energy of the atom in the presence of the field

\[

E = E_{n} + n\frakv_{F} h

= -\frac{2\pi^{2} N^{2} e^{4} m}{n^{2} h^{2}} + \frac{3h^{2} n\frakn}{8\


pi^{2} Nem}\, F,

\Tag{(77)}

\]

where $\frakn$~is an entire number which in the present case may be taken
positive as

well as negative. From a comparison between \Eq{(75)}~and \Eq{(77)}, we see


that the presence

of the external field imposes the restriction on the motion of the atom in the
stationary

states, that the plane in which the electrical centre of the orbit moves

must have a distance from the nucleus equal to an entire multiple of the \Ord{$n$}
{th}~part

of its maximum distance~$\dfrac{3}{2} \alpha_{n}$.

271
The result, contained in formula~\Eq{(77)}, is in agreement with the expression
for

the total energy in the stationary states, deduced by \Name{Epstein} and \


Name{Schwarzschild}

by means of the general theory of conditionally periodic systems based on the

conditions~\EqI{(22)}. The treatment of these authors rests upon the fact, that, as
mentioned

in Part~I, the equations of motion for the electron in the present problem may be

solved by means of separation of variables in parabolic coordinates (compare \


PageRefI[page]{21}).

Taking for $q_{1}$~and~$q_{2}$ the parameters of the two paraboloids of


revolution, which

pass through the instantaneous position of the electron and which have their foci

at the nucleus and their axes parallel to the direction of the field, and for~$q_{3}$
the

angular distance between the plane through the electron and the axis of the system

and a fixed plane through this axis, the momenta $p_{1}$,~$p_{2}$,~$p_{3}$


will during the motion

depend on the corresponding $q$'s only, and the stationary states will be fixed by

three conditions of the type~\EqI{(22)}. With neglect of small quantities


proportional to

higher powers of~$F$, the final formula for the total energy, obtained by \
Name{Epstein} in

this way, is given by

\[

\BreakMath{%

272
E = -\frac{2\pi^{2} N^{2} e^{4} m}{h^{2} (n_{1} + n_{2} + n_{3})^{2}}

- \frac{3h^{2} (n_{1} + n_{2} + n_{3})(n_{1} - n_{2})}{8\pi^{2} Nem}\, F,\


quad\footnotemark

}{%

\begin{aligned}

E &= -\frac{2\pi^{2} N^{2} e^{4} m}{h^{2} (n_{1} + n_{2} + n_{3})^{2}} \\

&\qquad\qquad- \frac{3h^{2} (n_{1} + n_{2} + n_{3})(n_{1} - n_{2})}{8\


pi^{2} Nem}\, F,\quad\footnotemark

\end{aligned}

\Tag{(78)}

\]

\footnotetext{\Name{P.~Epstein}. Ann.\ d.\ Phys.\ L, p.~508 (1916).}

\PageSep{74}

where $n_{1}$,~$n_{2}$,~$n_{3}$ are the positive entire numbers which occur


as factors to \Name{Planck}'s

constant on the right sides of the mentioned three conditions.

As regards the possible values of the total energy of the hydrogen atom in

the presence of the electric field, it will be seen that \Eq{(78)}~coincides with~\
Eq{(77)} if we

put $n_{1} + n_{2} + n_{3} = n$ and $n_{1} - n_{2} = \frakn$. At the same time
it will be observed,

however, that the motion in the stationary states, as fixed by the procedure
followed

273
by \Name{Epstein}, is more restricted than was necessary in order to secure the
right

relation between the additional energy and the frequency of the secular
perturbations.

Thus, in addition to the condition which fixes the plane in which the electrical

centre moves, \Name{Epstein}'s\Pagelabel{74} theory involves the further


condition, that the angular

momentum of the electron round the axis of the perturbed system is equal to an

entire multiple of~$h/2\pi$; which multiple is seen to be even or uneven, according


as

$n + \frakn$ is an even or an uneven number respectively. This circumstance is


intimately

connected with the fact that, although the perturbed system under consideration is

degenerate if we look apart from small quantities proportional to the square of the

intensity of the external force, the degenerate character of the system does not
reveal

itself from the point of view of the theory of stationary states based on the
conditions~\EqI{(22)},

because the system under consideration allows of separation of variables

only in one set of positional coordinates. On the other hand, this degenerate

character of the system has been emphasised by \Name{Schwarzschild}\footnote

{\Name{K.~Schwarzschild}, Ber.\ Akad.\ Berlin, 1916, p.~548.}

on the basis

of the theory of stationary states based on the introduction of angle variables, in

which the periodicity properties of the motion play an essential part. In a later

274
discussion of this point \Name{Epstein}\footnote

{\Name{P.~Epstein}, Ann.\ d.\ Phys.\ LI, p.~168 (1916).}

calls attention to the fact that, if small quantities

proportional to the square of the electric force are taken into account, the system

appears no more as degenerate; and he finds therein a justification of the fixation


of

the stationary states by means of~\EqI{(22)}. From the point of view of perturbed
systems,

this would mean that the motion in the stationary states of the system in question,

as fixed by~\EqI{(22)}, would certainly be stable for infinitely small disturbances,


but that

we should expect finite deviations from the motion in these states, already if the

system was exposed to a second perturbing field, the intensity of which was only

of the same order as the product of the external electric force with the ratio
between

this force and the attraction from the nucleus. A closer consideration, however, in

which regard is taken to the influence of the relativity modifications, learns that

the degree of stability of the motion in the stationary states, as determined by~\
EqI{(22)},

actually is often much higher, the order of magnitude of the external force,
necessary

to cause finite deviations from this motion, being of the same order as the product

of the attraction from the nucleus with the square of the ratio of the velocity of

the electron and the velocity of light. To this point we shall come back at the end

of this section, when considering the simultaneous perturbing influence on the

275
\PageSep{75}

motion of the election in the hydrogen atom, due to the relativity modifications

and an external electric field.

In the deduction of formula~\Eq{(78)} there is looked apart, not only from the
effect

on the motion of the electron due to the small modifications in the laws of
mechanics

claimed by the theory of relativity, but also from the effect of possible forces

which might act on the electron, corresponding to the reaction from the radiation

in ordinary electrodynamics. If, however, for the moment we exclude all stationary

states for which the angular momentum round the axis of the system would be

equal to zero ($n_{3} = 0$), the total angular momentum of the electron round the

nucleus will during the perturbations always remain larger than or equal to~$h/2\
pi$,

just as in the stationary states considered in the theory of the fine structure; and,

according to the considerations on \PageRef[page]{66}, we shall therefore expect


that the effect

of the neglect of possible ``radiation'' forces will be small compared with the effect

of the relativity modifications. On the other hand, if the intensity of the electric
field is

of the same order of magnitude as that applied in \Name{Stark}'s experiments, the


effect of

these modifications must again be expected to be very small compared with the

total effect of the electric force on the hydrogen lines, since the perturbing effect

276
of this force on the Keplerian motion of the electron will be very large compared

with the corresponding effects of the relativity modifications. If, on the

contrary, we would consider a state of the atom for which $n_{2}$~was equal to
zero,

the orbit would be plane and would during the perturbations assume shapes, for

which the total angular momentum round the nucleus was very small, and in which

the electron during the revolution would pass within a very short distance from

the nucleus. In such a state the effect of the relativity modifications on the motion
of

the electron would be considerable, but quite apart from this a rough calculation
shows

that the amount of energy, which, on ordinary electrodynamics, would be emitted

during the intervals in which the angular momentum during the perturbations of

the orbit remains small, is so large that it would hardly seem justifiable to calculate

the motion and the energy in these states by neglecting all forces corresponding to

the radiation forces in ordinary electrodynamics. We need not, however, enter


more

closely on these difficulties, because, on the general considerations in Part~I about

the a-priori probability of the different stationary states, we are forced to conclude

that, for any value of the external electric field, \Emph{no state which would
correspond

to $n_{3} = 0$ will be physically possible}; since any such state might be

transformed continuously, and without passing through a degenerate system, into

a state which obviously cannot represent a physically realisable stationary state

277
(compare \PageRefI[\Chg{pag.}{page}]{27}). In fact, if we imagine that an
external central field of force,

varying as the inverse cube of the distance from the nucleus, is slowly established,

it would be possible to compensate the secular effect of the relativity modifications

and to obtain orbits in which the electron would pass within any given, however

small, distance from the nucleus. As regards the other stationary states fixed

\PageSep{76}

by~\EqI{(22)}, which correspond to $n_{3} > 1$, we shall according to the


considerations in Part~I

expect that their a-priori probabilities are all equal.\Pagelabel{76}\footnote

{By a simple enumeration it follows from this result, that the total number of
different stationary

states of the hydrogen atom, subject to a small homogeneous electric field, which
corresponds to a stationary

\Typo{states}{state} of the undisturbed atom, characterised by a given value


of~$n$ in the condition $I = nh$,

is equal to $n(n + 1)$. This expression is directly obtained, if we remember that


$n = n_{1} + n_{2} + n_{3}$

and if we count each state, characterised by a given combination of the positive


integers $n_{1}$,~$n_{2}$,~$n_{3}$, as

double, corresponding to the two possible opposite directions of rotation of the


electron round the axis

of the field. With reference to the necessary stability for a small variation of the
external conditions

of the statistical distribution of the values of the energy among a large number of
atoms in temperature

278
equilibrium (see Note on \PageRef[page]{43}), it will be seen that the expression
$n(n + 1)$ may be taken as

a measure for \Emph{the relative value of the a-priori probability of the different
stationary

states of the undisturbed hydrogen atom}, corresponding to different values


of~$n$. The problem of

the determination of this a-priori probability has been discussed by \


Name{K.~Herzfeld} (Ann.\ d.\ Phys.\

LI, p.~261 (1916)) who, by an examination of the volumes of the different


extensions in the phase space

which might be considered as belonging to the different stationary states of the


hydrogen atom, has

arrived at an expression for the a-priori probability of these states which differs
from the above. From

the point of view, as regards the principles of the quantum theory, taken in the
present paper, a consideration

of this kind, however, does not, as explained in Part~I on \PageRefI[page]{26},


afford a rational means

of determining the a-priori probability of the stationary states of an atomic


system.}

As regards the comparison between the theory and the experiments, it will be

remembered that \Name{Stark} found that every hydrogen line in the presence of
an

electric field was split up in a number of polarised components, in a way different

for the different lines. When viewed parallel to the direction of the field, there
appeared

279
a number of components polarised parallel to the field and a number of
components

polarised perpendicular to the field; when viewed in the direction of the

field, only the latter components appeared, but without showing characteristic

polarisation. Apart from the marked symmetry of the resolution of every line,

the distances between successive components and their relative intensities varied

in an apparently irregular way from component to component. As pointed out by

\Name{Epstein} and \Name{Schwarzschild}, however, it is possible by means


of~\Eq{(78)}, in connection

with relation~\EqI{(1)}, to account in a convincing way for \Name{Stark}'s


measurements as regards

the \Emph{frequencies} of the components. Especially a closer examination of


these measurements

showed that all the differences between the frequencies of the components

were equal to entire multipla of a certain quantity, which was the same for all lines

in the spectrum and, within the limits of experimental errors, equal to the
theoretical

value $\dfrac{3hF}{8\pi^{2} Nem}$. On the other hand, the theories of \


Name{Epstein} and \Name{Schwarzschild} gave

no direct information as regards the question of the polarisation and intensity

of the different components. Comparing formula~\Eq{(78)} with \Name{Stark}'s


observations,

\Name{Epstein} pointed out, however, that the \Emph{polarisation} of the


different components

observed could apparently be accounted for by the rule: that a transition between

280
two stationary states gives rise to a component polarised parallel to the field, if
$n_{3}$~remains

unchanged or is changed by an even number of units; while a component,

corresponding to a transition in which $n_{2}$~is changed by an uneven number


of units,

\PageSep{77}

is polarised perpendicular to the field. This result may be simply interpreted on

the basis of the general formal relation between the quantum theory of line

spectra and the ordinary theory of radiation. In fact, it was shown in Part~I

that, for a conditionally periodic system possessing an axis of symmetry, we

shall expect only two types of transitions to be possible. In transitions of the

first type $n_{3}$~remains unchanged, and the emitted radiation is polarised


parallel to

the axis of symmetry, while the transitions of the second type, in which
$n_{3}$~varies

by one unit, give rise to a radiation of circular polarisation in a plane perpendicular

to this axis (see \PageRefI[page]{34}). In order to show that this agrees with the
empirical rule

of \Name{Epstein}, it may be noted in the first place that, for any component
which might

be ascribed to a certain transition in which $n_{3}$~changes by a given entire


number

of units, there exists always another transition which will give rise to a radiation

of the same frequency but in which $n_{3}$~remains unchanged or changes by


one

281
unit, according to whether the given number is even or uneven. Next it will be

seen that, in case of the effect of an electric field on the hydrogen spectrum, we

cannot detect by means of direct observations the circular polarisation of the

radiation corresponding to transitions of the second type; because, for each


transition

giving rise to a radiation of circular polarisation in one direction, there will

exist another transition giving rise to a radiation which possesses the same
frequency

but is polarised in the opposite direction. Besides on the problem of the

polarisations of the different components into which the hydrogen lines are split up

in the presence of the electric field, the general considerations in Part~I allow also
to

throw light on the question of the \Emph{relative intensities} of these components,


by

considering the harmonic vibrations into which the motion of the electron in the

stationary states can be resolved. Compared with the problem of the relative
intensities

of the components of the fine structure of the hydrogen lines, the present

problem is simpler in that respect, that the stationary states may be assumed to

be a-priori equally probable. Since the different components, into which a given

hydrogen line is split up in the electric field, correspond to transitions between


pairs of

states which for all components have very nearly the same values for the total
energy,

these states may therefore be expected to be of approximately equal occurrence in

282
the luminous gas. According to the considerations in Part~I, we shall consequently

assume that for a given hydrogen line the relative intensities of the different \
Name{Stark}

effect components, corresponding to transitions between different pairs of


stationary

states characterised by $n_{1} = n_{1}'$, $n_{2} = n_{2}'$, $n_{3} = n_{3}'$


and $n_{1} = n_{1}''$, $n_{2} = n_{2}''$, $n_{3} = n_{3}''$

respectively, will be intimately connected with the intensities of the radiations of

frequency $(n_{1}' - n_{1}'') \omega_{1} + (n_{2}' - n_{2}'') \omega_{2} +


(n_{3}' - n_{3}'') \omega_{3}$, which on ordinary electrodynamics

would be emitted by the atom in the two states involved in the transition

in question; $\omega_{1}$,~$\omega_{2}$, and~$\omega_{3}$ being the


fundamental frequencies entering in the expression~\EqI{(31)}

for the displacement of the electron. In order to test how far such

a connection is actually brought out by the observations, it is necessary to


determine

the numerical values of the amplitudes of the harmonic vibrations into which the

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd., 8. Række, IV. 1.


11.

\PageSep{78}

motion of the electron can be resolved. The examination of this problem has been

undertaken by Mr.\ \Name{H.~A. Kramers}, who has deduced complete


expressions for these

amplitudes, by means of which it was found possible, for each of the hydrogen

283
lines $H_{\alpha}$,~$H_{\beta}$,~$H_{\gamma}$ and~$H_{\delta}$, to account
in a convincing way for the apparently capricious

laws which govern the intensities of the components observed by \Name{Stark}.\


footnote

{\Emph{Note added during the proof.} In recent papers \Name{H.~Nyquist}


(Phys.\ Rev.\ X, p.~226 (1917))

and \Name{J.~Stark} (Ann.\ d.\ Physik, LVI, p.~569 (1918)) have published
measurements on the effect of an

electric field on certain lines of the helium spectrum which is given by~\
Eq{(35)}, if in~\Eq{(40)} we put $N = 2$.

As will be seen from~\Eq{(78)}, the differences between the frequencies of the


components into which these

lines are split up will, for the same intensity of the external electric field, be
smaller than for the

hydrogen lines. In conformity with this it was not possible, with the experimental
arrangement used

by the authors mentioned, to observe separately the numerous components to be


expected on the

theory, but only to obtain certain rough features of the resolution of the lines in
question. For the

interpretation of these observations a detailed consideration of the relative


intensities to be expected

for the different theoretical components is therefore essential; and, as it will be


shown in \Name{Kramers}'

paper, it is possible, on the basis of the calculation of the amplitudes of the


harmonic vibrations into

which the motion of the electron in the stationary states can be resolved, to
account satisfactorily for
284
\Name{Nyquist}'s and \Name{Stark}'s results.}

This agreement

offered at the same time a direct experimental support for the conclusions

mentioned above: that there exist no stationary states corresponding to $n_{3} =


0$, while

the stationary states corresponding to other values of~$n_{2}$ are a-priori equally
probable;

and that transitions can only take place between pairs of stationary states

for which $n_{3}$~is the same or differs by one unit. A general discussion of
these problems

will be given by \Name{Kramers} in the paper, mentioned on \PageRef[page]{69}


in the last

section, in which also the problem of the intensity of the fine structure components

is treated in detail.

In the former section and in the present we have seen, how the problems of

the influence of the relativity modifications on the lines of the hydrogen spectrum

and of the influence of an external electric field on this spectrum can be treated,

by regarding the motion of the electron as a perturbed periodic motion, and by

fixing the stationary states on the basis of the relation between the energy and the

frequencies of the secular perturbations. As it was done originally by \


Name{Sommerfeld}

and \Name{Epstein}, both these problems can also be treated by means of the
theory of

the stationary states of conditionally periodic systems which allow of separation

285
of variables in a fixed set of positional coordinates. If, however, we consider the

problem of \Emph{the simultaneous influence on the hydrogen spectrum of

the relativity modifications and a homogeneous electric field} of any

given intensity, there does not exist a set of coordinates for which a separation of

variables can be obtained. On the other hand it is possible, also in this case, to

apply the general considerations about perturbed periodic systems developed in the

preceding. In fact, with reference to the treatment given in~\SecRef{3} of the


problem of

the fine structure of the hydrogen lines, it will be seen that the deviations of the

orbit of the electron from a Keplerian ellipse in the problem under consideration

\PageSep{79}

will be the same as the secular perturbations produced on a Keplerian motion by

the simultaneous influence of an external homogeneous field of force and an


external

central force proportional to the inverse cube of the distance from the nucleus.

Since these two fields together form a perturbing field possessing axial symmetry,

it follows therefore that the secular perturbations, when the relativity \


Typo{modications}{modifications}

are taken into account, will be conditionally periodic and that the problem of the

stationary states may be treated by means of the method mentioned in~\SecRef{2}


on

\PageRef[page]{55}. In this way we obtain in the first place the result, that, for any
value of

the intensity of the external electric field, we must expect that the hydrogen lines

286
will be split up in a number of sharp components. Next, since for any value of

this intensity different from zero the system will be non-degenerate, it follows from

the conditions~\Eq{(61)}, that we must assume that the angular momentum round
the

axis of the field is always equal to an entire multiple of~$h/2\pi$; in consistence


with

the assumption of the validity of the analogous condition involved in the fixation

of the stationary states by means of the method of separation of variables, when

applied to an explanation of the \Name{Stark} effect with neglect of the relativity


modifications

(compare \PageRef[page]{74}). On the basis of the conditions~\Eq{(61)} it is


possible to

predict in detail, how the fine structure of the hydrogen lines will be influenced

by an increasing electric field until, for a sufficiently large intensity of this field,

the phenomenon develops gradually into the ordinary \Name{Stark} effect. The
problem

of this transmutation will be treated in a later paper by Mr.\ \Name{H.~A.


Kramers},\footnote

{Besides the discussion of this problem, the paper in question will contain a
general treatment

of the theory of perturbed periodic systems from the point of view of the
possibility of describing

the motion by means of angle variables (compare Note on \PageRef[page]{58}).}

who

has kindly drawn my attention to this interesting application of the method of

287
perturbations, and has thereby given a valuable impetus to the detailed elaboration

of this method as regard the treatment of more complicate problems.

\Section{5.}{The effect of a magnetic~field on~the~hydrogen~spectrum.}

A theory of the \Name{Zeeman} effect of the hydrogen fines based on the


quantum

theory of line spectra has, as mentioned in the introduction, been given


independently

by \Name{Sommerfeld} and by \Name{Debye}. The calculations of these authors


rest upon

the fact, that it is possible, also in the presence of a magnetic field, to write the

equations of motion of the electron in the canonical Hamiltonian form given by~\
EqI{(4)},

if the momenta $p_{1}$,~$p_{2}$,~$p_{3}$ which are conjugated to the


positional coordinates of

the electron $q_{1}$,~$q_{2}$,~$q_{3}$, are defined in a suitable way. In


complete analogy to the

problem of the fixation of the stationary states of an atomic system when the
relativity

modifications are taken into account, it follows therefore that, if these equations

\PageSep{80}

can be solved by the method of separation of variables, we obtain, by fixing the

288
stationary states by means of the conditions~\EqI{(22)}, a relation between the
total energy

of the atom in the presence of a magnetic field and the fundamental frequencies

characterising the motion of the electron, which is exactly the same as that holding

between the energy and frequencies in the stationary states of an ordinary


conditionally

periodic system. By a procedure analogous to that applied by \Name{Burgers}

in his proof of the mechanical invariance of the relations~\EqI{(22)} for slow


changes of

the external conditions, mentioned in Part~I on \PageRefI[page]{21}, it may


further be proved

that also in the presence of a magnetic field these relations are invariant, when

regard is taken to the effect of the induced electric forces which, according to the

ordinary theory of electrodynamics, will accompany a variation of the magnetic

field. In the following, however, we shall not treat the problem of the influence of

an external magnetic field on the hydrogen spectrum by means of the method of

separation of variables, but in analogy to the treatment of the problems of the fine

structure and of the \Name{Stark} effect of the hydrogen lines, given in the
preceding

sections, we shall treat the problem from the point of view of the theory of
perturbed

periodic systems. Before entering on the detailed discussion of the necessary

modifications to be introduced in the general considerations in~\SecRef{2}, in


order that

they may be applied also to the problem of the fixation of the stationary states of

289
the atom in the presence of external magnetic forces, we shall for the sake of

illustration first show how it is possible in certain cases to treat the problem of

the effect of a homogeneous magnetic field on the hydrogen spectrum in a simple

way, which will be seen to present a close formal analogy with the theory
originally

devised by \Name{Lorentz} on the basis of the classical theory of electrons.

In these considerations we shall make use of a well known theorem of \


Name{Larmor},

which states that, if we look apart from small quantities proportional to the

square of the intensity of the magnetic field, the motion of a system of electrons

moving in a conservative field of force possessing axial symmetry round a fixed

axis will, in the presence of an external homogeneous magnetic field parallel to

this axis, differ from a mechanically possible motion of the system without field,

only by a superposed uniform rotation of the entire system round the axis, the

frequency of which is given by

\[

\frakv_{H} = \frac{e}{4\pi me}H,

\Tag{(79)}

\]

where $H$~is the intensity of the magnetic field and $c$~the velocity of light,
while $-e$~and

$m$ represent the charge and the mass of an electron.\footnote

290
{\Name{J.~Larmor}, Aether~\& Matter, Cambridge 1900, p.~341. This theorem,
which was established in

connection with an attempt to develop a general theory of the \Name{Zeeman}


effect based on the ordinary theory

of electrodynamics, is directly proved by observing that, with the degree of


approximation in question,

the accellerations of the electrons due to the presence of the magnetic field are
equal to the changes in

the accellerations of the particles due to the superposed rotation of the system.}

If the magnetic field

\PageSep{81}

is not constant, but if its intensity increases slowly and uniformly from zero, it is

further simply shown that the electric induction forces, which will accompany the

change in the intensity of the magnetic force, will just effect that a rotation as

that described will be impressed on the original motion of the system.\footnote

{Compare \Name{P.~Langevin}, Ann.\ de Chim\Add{.}\ et de Phys.\ V, p.~70


(1905), who has deduced this result in

connection with his well known theory of the magnetic properties of atomic
systems based on the

classical theory of electrons.}

Moreover,

as regards the effect of the magnetic field on the total energy of the system,\
footnote

{In an earlier paper (Phil.\ Mag.\ XXVII, p.~506 (1914)) the writer had assumed
that the total

291
energy in the stationary states of the hydrogen atom in the presence of a magnetic
field would not be

different from the energy in the corresponding states without field, as far as small
quantities proportional

to the intensity of the magnetic force are concerned; the effect on the kinetic
energy of the electron

due to the superposed rotation being assumed to be compensated by some kind of


``potential'' energy of

the whole atom relative to the magnetic field. This assumption seemed not only
suggested by the absence

of \Emph{paramagnetism} in many elements, the atoms and molecules of which,


according to the

theory to be discussed in Part~IV, must be expected to possess a resultant angular


momentum, but it

was especially thought to be supported by the fact, that the spectrum, emitted by
hydrogen in the

presence of a magnetic field, apparently did not form a combination spectrum of


the type which should

be expected, it the frequency of the radiation, emitted during a transition between


two stationary states

of the atom in the presence of the field, could be calculated directly from the
values of the energy in

these states by means of relation~\EqI{(1)}. As remarked by \Name{Debye}


(Phys.\ Zeitschr.\ XVII, p.~511 (1916)), this

view, however, would not be reconcilable with \Name{Einstein}'s theory of


temperature radiation (see Part~I,

\PageRefI[page]{7}) which implies the general validity of relation~\EqI{(1)};


and, moreover, as will be shown in the following,
292
the \Name{Zeeman} effect of the hydrogen lines may actually be considered, not
as involving a deviation

from the combination principle, but rather as affording an instructive example of


a systematic disappearance

of certain possible combination lines, for which a simple explanation can be


obtained from

a consideration based on the general formal relation between the quantum theory
of line spectra and

the ordinary theory of radiation. Further, with reference to this relation---and


remembering that on

ordinary electrodynamics the magnetic field will not directly influence the
exchange of energy during

a process of radiation, since the forces due to this field, being always
perpendicular to the direction

of the velocity, will not perform work on the moving electron---it seems also
natural to assume that

it is possible, simply from the effect of the superposed rotation on the kinetic
energy of the electron,

to determine the effect of the magnetic field, as regards the \Emph{differences}


between the values of the

energy in the different stationary states of the atom. Now, in a discussion of the
spectrum to be expected

on the quantum theory, we are concerned only with these differences and not with
the

\Emph{absolute values} of the additional energy of the system due to the


presence of the magnetic field.

It would therefore be possible to escape from the difficulty, mentioned above, as


regards the absence
293
of paramagnetism, by assuming that only the energy in the \Typo{socalled}{so
called} ``normal'' state of an atomic

system (\ie\ the stationary state of the system which possesses the smallest value
for the total energy;

see Part~IV) is not altered in the presence of a magnetic field, as far as small
quantities proportional

to the intensity of the magnetic force are concerned. On this view, the absence of
paramagnetism would

thus be a special property of the normal state, connected with the impossibility of
spontaneous transitions

from this state to other stationary states of the system. To this question we shall
come back in

the following parts of this paper; for the sake of simplicity, however, we shall not,
in the considerations

of this section, enter more closely on the consequences of the mentioned


hypothesis, which would

imply small modifications in the form of the following considerations, but would
not affect the results.}

it will\Pagelabel{81}

be observed that the superposed rotation under consideration will not affect the

mutual potential energy of the particles, while, with neglect of small quantities
proportional

to~$H^{2}$, it will produce a change in the kinetic energy equal to~$2\pi P\


frakv_{H}$,

\PageSep{82}

where $P$~represents the total angular momentum of the system round the axis,

taken in the same direction as that of the superposed rotation.


294
From these results it follows that the motion of the electron in any stationary

state of a hydrogen atom, which is exposed to a \Emph{homogeneous magnetic

field}, will---if we look apart from small quantities proportional to the square

of the intensity of the magnetic force and to the product of this intensity with the

ratio between the mass of the electron and that of the nucleus---differ from the

motion in some stationary state of the atom in the absence of the field, only by

a superposed uniform rotation round an axis through the nucleus parallel to the

magnetic force with a frequency given by~\Eq{(79)}. Due to the degenerate


character of

the system formed by the atom in the absence of the magnetic field, it is not
possible,

however, from a consideration of the mechanical effect produced on the

motion of the electron by a slow and uniform establishment of the magnetic field,

to fix the stationary states of the perturbed atom completely, but in order to fix

these states we must consider more closely the relation between the additional

energy of the system due to the presence of the magnetic field and the character

of the secular perturbations produced by this field on the orbit of the electron.

On the basis of \Name{Larmor}'s theorem the discussion of this problem is very


simple.

In fact, since the frequency~$\frakv_{H}$ is independent of the shape and position


of the

orbit, we may proceed in a manner which is completely analogous to that applied

in the fixation of the stationary states of the hydrogen atom in the presence of a

295
homogeneous electric field. Thus, looking apart from the effect of the relativity

modifications, we may conclude at once that the total energy in the stationary

states of the atom will be given by

\[

E = E_{n} + \frakn \frakv_{H}h,

\Tag{(80)}

\]

where $\frakn$~is an entire number which can be positive as well as negative,


while $E_{n}$~will

be equal to the energy in the corresponding stationary state of the undisturbed

atom, which is given by $-W_{n}$~in~\Eq{(41)}. As in the case of the \


Name{Stark} effect,

it will moreover be seen that this formula includes the values of the energy in

such states of the atom, in which the electron moves in a circular orbit
perpendicular

to the direction of the field, and which beforehand must be expected to be

included among the stationary states of the perturbed system, since such orbits

during a slow and uniform establishment of the external field will not undergo

secular perturbations as regards shape and position (compare \PageRef[page]{73}).


In fact,

since in these cases we have $P = ±nh/2\pi$, where $n$~is the entire number
characterising

the stationary states of the undisturbed hydrogen atom, it follows from the

above that the total energy in the special stationary states under consideration will

296
just be represented by the formula~\Eq{(80)}, if we put $\frakn = ±n$. From this
formula it

will be seen at the same time, that the presence of the external magnetic field
imposes

the restriction on the motion in the stationary states of the hydrogen atom,

that, with neglect of small quantities proportional to~$H$, the angular momentum

\PageSep{83}

of the electron round the axis of the field will be equal to an entire multiple

of~$h/2\pi$.

As regards the expression for the total energy of the hydrogen atom in the

presence of the magnetic field, formula~\Eq{(80)} is in agreement with the


formulæ obtained

by \Name{Sommerfeld} and \Name{Debye} on the basis of the conditions~\


EqI{(22)}, holding for conditionally

periodic systems which allow of separation of variables. As shown by these

authors, a system, which consists of an electron moving under the influence of

the attraction from a fixed nucleus and of a homogeneous magnetic field, allows of

separation of variables in polar coordinates, if the polar axis is chosen parallel to

the magnetic field. Looking apart from the effect of the relativity modifications,

and choosing for $q_{1}$,~$q_{2}$, and~$q_{3}$ the length of the radius vector
from the nucleus to

the electron, the angle between this radius vector and the axis of the system, and

the angle which the plane through the electron and this axis makes with a

297
fixed plane through the axis respectively, they obtain the following expression for

the total energy:\footnote

{\Name{A.~Sommerfeld}, Phys.\ Zeitschr.\ XVII, p.~491 (1916) and \


Name{P.~Debye}, Phys.\ Zeitschr.\ XVII, p.~507

(1916). While \Name{Debye} proceeds directly by the application of the


conditions~\EqI{(22)} in a fixed set of positional

polar coordinates, \Name{Sommerfeld} determines the stationary states by


applying these conditions to

the motion of the system relative to a set of coordinates which rotates uniformly
round the polar axis

with the frequency~$\frakv_{H}$; a procedure which in the special case under


consideration is simply shown to

give the same result as the direct application of~\EqI{(22)} to fixed polar
coordinates.}

\[

E = -\frac{2\pi^{2} N^{2} e^{4} m}{h^{2} (n_{1} + n_{2} + n_{3})^{2}}

± \frac{ehn_{3}}{4\pi mc} H,

\Tag{(81)}

\]

where $n_{1}$,~$n_{2}$, and~$n_{3}$ are the integers which appear as factors to


\Name{Planck}'s constant

on the right side of the conditions~\EqI{(22)}. As mentioned this formula gives the
same

result as~\Eq{(80)}; in fact, if we put $n = n_{1} + n_{2} + n_{3}$ and if we look


apart from the

298
small correction due to the finite mass of the nucleus, the first term in~\Eq{(81)} is
seen

to coincide with the expression for~$-W_{n}$ given by~\Eq{(41)}, while the last
term in~\Eq{(81)}

coincides with the last term in~\Eq{(80)}, if we put $\frakn = n_{3}$. It will be
observed, however, that,

while in the theories of \Name{Sommerfeld} and \Name{Debye} the stationary


states are characterised

by three conditions, only two conditions were necessary on the above


considerations

in order to secure the right relation between the energy and frequencies

of the system in the stationary states. Thus, besides the conditions which prescribe

the length of the major axis of the rotating orbit and the value of the angular

momentum of the system round the axis of the field, the theories of the mentioned

authors involve the further condition, that the value of the total angular momentum

of the electron round the nucleus must be equal to an entire multiple of~$h/2\pi$;
and

that consequently the minor axis of the orbit has the same values as in a hydrogen

atom perturbed by a small external central field (compare \PageRef[page]{57}).


This is due to

the circumstance, that the perturbed atom forms a \Emph{degenerate system} if we

look apart from the effect of the relativity modifications, because the secular
perturbations

\PageSep{84}

are simply periodic. From the point of view of separation of variables,

299
this degenerate character of the system is in the present case, in contrast to the

analogous case of the \Name{Stark} effect, also directly revealed by the fact, that a
separation

can be obtained, not only in polar coordinates,\Pagelabel{84} but in any set of


axial elliptical

coordinates for which one focus is placed at the nucleus and the other at

some point on the axis of the field. Just as in the case of the \Name{Stark} effect,
however,

the system is no more degenerate as soon as the relativity modifications are

taken into account, in which case a separation of variables will still be possible

but only in polar coordinates. To this point we shall come back below.

The observations on the \Name{Zeeman} effect of the hydrogen lines show that, if
the

fine structure is neglected, each line is in the presence of a magnetic field split up

in a normal \Name{Lorentz} triplet; \ie\ each line is resolved in three components


of

which the one is undisplaced and polarised parallel to the direction of the field,

while the two other components possess frequencies, which differ from that of the

original line by~$\frakv_{H}$, and are circularly polarised in opposite directions


in a plane

perpendicular to the direction of the field. As pointed out by \Name{Sommerfeld}


and by

\Name{Debye}, the frequencies of a \Name{Lorentz} triplet are included among


the frequencies of the

300
components deduced from~\Eq{(81)} by application of relation~\EqI{(1)}. In
addition to the

observed components, however, we might from \Eq{(81)}~and~\EqI{(1)} expect


the appearance

of a number of components, displaced from the original positions of the lines by

higher multipla of~$\frakv_{H}$. For the non-appearance of these components the


theories of

\Name{Sommerfeld} and \Name{Debye} offered no explanation, no more than for


the polarisation of

the components observed; except that \Name{Sommerfeld} in this connection


draws attention

to the fact, that the law governing the observed polarisations exhibits a certain

analogy to the \Typo{emperical}{empirical} rule of \Name{Epstein} concerning


the observed polarisations of

the components of the \Name{Stark} effect of the hydrogen lines (see \


PageRef[page]{76}). On the other

hand, just as in case of the latter effect, an explanation of the number of the
components

observed and their characteristic polarisations is directly obtained on the

basis of the general formal relation between the quantum theory of line spectra and

the ordinary theory of radiation. In the first place we have at once from \
Name{Larmor}'s

theorem, denoting the frequency of revolution of the electron in a stationary state


of

the undisturbed hydrogen atom by~$\omega$, that the motion of the electron, in a
corresponding

stationary state of the atom in the presence of the field, may be resolved
301
in a number of linear harmonic vibrations parallel to the direction of the magnetic

force with frequencies~$\tau\omega$, where $\tau$~is a positive integer, and in a


number of circular

harmonic rotations perpendicular to this direction with frequencies $\tau\omega + \


frakv_{H}$ or $\tau\omega - \frakv_{H}$,

according as the direction of rotation is the same as or the opposite of that of the

superposed rotation. Next, with neglect of small quantities proportional


to~$H^{2}$ we

have for the difference in the total energy between two neighbouring states of the

perturbed system under consideration

\[

\delta E = \delta E_{0} + \delta \frakE

= \omega\, \delta I + \frakv_{H}\, \delta \frakI,

\Tag{(82)}

\]

\PageSep{85}

where $E_{n}$~and~$\omega$ are the values of the energy and frequency and
$I$~is the value of the

quantity defined by~\EqI{(5)}, all corresponding to the state of the undisturbed


system

which would appear if the magnetic force vanished at a slow and uniform rate,

while $\frakE$~is the additional energy due to the presence of the magnetic field
and $\frakI$~the

angular momentum of the system round the axis of the field multiplied by~$2\pi$

302
and taken in the same direction as that of the superposed rotation. Since \
Eq{(82)}~has

exactly the same form as relation~\Eq{(66)}, and since in the stationary states we
have

$I = nh$ and $\frakI = \frakn h$, we are therefore from a consideration, quite
analogous to

that given in~\SecRef{2} on \PageRef[page]{59}, led to the conclusion, that, in the


presence of

the magnetic field, only two types of transitions between stationary states are

possible. For both types of transitions the integer~$n$ may change by any number

of units, but in transitions of the first type the integer~$\frakn$ will remain
constant and

the emitted radiation will be polarised parallel to the direction of the field, while

in transitions of the second type $\frakn$~will decrease or increase by one unit and
the

emitted radiation will be circularly polarised in a plane perpendicular to the field,


the

direction of the polarisation being the same as or the opposite of that of the
superposed

rotation respectively. Remembering that, with neglect of small quantities


proportional

to the magnetic force, the angular momentum of the system round the

axis of the field remains unaltered in transitions of the first type and changes
by~$h/2\pi$

in transitions of the second type, it will be seen that this conclusion is


independently

supported by a consideration of conservation of angular momentum


303
during the transitions, like that given in Part~I on \PageRefI[page]{34}.

With reference to formula~\Eq{(80)}, it will be seen that the above results are in

complete agreement with the experiments on the \Name{Zeeman} effect of the


hydrogen

lines, as regards the \Emph{frequencies and polarisations} of the observed


components.

On the other hand, the observed \Emph{intensities} are directly accounted for,
independent

of any special theory about the origin of the lines. In fact, from a consideration of

the necessary ``stability'' of spectral phenomena, it follows that the total radiation
of

the components, in which a spectral line, which originally is unpolarised, is split up

in the presence of a small external field, cannot show characteristic polarisation

with respect to any direction. In case of the \Name{Zeeman} effect of the


hydrogen lines, it

is therefore necessary beforehand to expect that the intensity of the radiation,

summed over all directions, corresponding to each of the three components in

which every line is split up must be the same. From the point of view of the

quantum theory of line spectra, it will be seen that by means of considerations

of this kind we may inversely obtain a certain amount of direct quantitative


information

as regards the probabilities of spontaneous transition between different

sets of stationary states, holding also in the region where the integers characterising

these states are not large and where consequently the estimate of the values of

304
these probabilities, based on the formal relation between the quantum theory and

the ordinary theory of radiation, gives results which are only of an approximative

character. This point will be discussed more closely in \Name{Kramers}' paper on


the relative

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd. 8. Række, IV. 1.


12

\PageSep{86}

intensities of the components of the fine structure and the \Name{Stark} effect of
the

hydrogen lines.

A procedure quite analogous to that applied above may be used to treat the

problem of the effect of a homogeneous magnetic field on the hydrogen spectrum,

also when the relativity modifications are taken into account, and when the atoms

at the same time are exposed to a small external field of force of constant potential,

which possesses axial symmetry round an axis through the nucleus parallel to the

magnetic force; because also in this case we can obviously make direct use of

\Name{Larmor}'s theorem. We shall not, however, proceed in this way, but shall
come

back to these questions when we have shown how, by a simple modification

of the general considerations of perturbed periodic systems given in~\SecRef{2}, it


is possible

to represent the theory of the stationary states of the hydrogen atom in the

305
presence of a small magnetic field on a form, which allows to discuss the effect on

the hydrogen spectrum also if the atom is exposed to a magnetic field which is

not homogeneous, or to discuss the effect of a homogeneous magnetic field if


electric

forces, which do not possess axial symmetry round an axis through the nucleus

parallel to the magnetic field, are acting on the atom at the same time.

In order to examine the general problem of the secular perturbations of the

orbit of the electron in the hydrogen atom which take place if the atom is exposed

to small external forces which, entirely or partly, are of magnetic origin,

we shall, as in the usual theory of planetary perturbations, take our starting point

in the equations of motion in their canonical form. Now the equations of motion

of an electron of charge~$-e$, which besides by an electric field of potential~$V$


is

acted upon by a magnetic field of vector potential~$\frakA$ (defined by $\div \


frakA = 0$ and

$\curl \frakA = \frakH$, where $\frakH$~is the magnetic force considered as a


vector), can be written

in the Hamiltonian form given by~\EqI{(4)}, if, just as in the absence of the
magnetic

field, $E$~is taken equal to the sum of the kinetic energy~$T$ of the electron and
its

potential energy~$-eV$ relative to the electric field, while the momenta which are

conjugated to the positional coordinates $q_{1}$,~$q_{2}$,~$q_{3}$ of the


electron in space are

306
defined by the equations\footnote

{See f.\ inst.\ \Name{G.~A. Schott}: Electromagnetic Radiation, App.~F


(Cambridge, 1912).}

\[

p_{k}' = p_{k} - \frac{e}{c}\, \frac{\dd(\frakv \frakA)}{\dd \dot{q}_{k}},

\qquad

(k = 1, 2, 3)

\Tag{(83)}

\]

where the~$p$'s are the momenta defined in the usual way (compare \
PageRefI[page]{10}), and

where $(\frakv \frakA)$~represents the scalar product of the velocity of the


electron~$\frakv$ and the

vector potential~$\frakA$, considered as a function of the~$q$'s and of the


generalised velocities

$\dot{q}_{1}$,~$\dot{q}_{2}$,~$\dot{q}_{3}$. If we now assume that the effect


of the magnetic forces on the

motion of the electron is so small compared with the effect of the electric forces,

that in the calculations we may look apart from all terms proportional to~$\
frakH^{2}$, it

is simply seen that the energy function~$E$ in~\EqI{(4)}, obtained by introducing


the

momenta defined by~\Eq{(83)}, will differ from the corresponding function,


holding in

\PageSep{87}

307
the absence of the magnetic field, only by the addition of a term which is linear

in the momenta and equal to~$\dfrac{e}{c}(\frakv \frakA)$. In fact, denoting


$E$~expressed as a function

of the $q$'s~and~$p$'s by~$\phi(p, q)$, we get from~\Eq{(83)} together with~\


EqI{(4)}, with the approximation

under consideration,

\[

\BreakMath{%

E - \phi(p', q)

= -\sum_{1}^{3} \frac{\dd \phi}{\dd p_{k}'}(p_{k}' - p_{k})

= \sum_{1}^{3} \frac{\dd E}{\dd p_{k}'}\, \frac{e}{c}\, \frac{\dd(\frakv \


frakA)}{\dd \dot{q}_{k}}

= \frac{e}{c} \sum_{1}^{3} \dot{q}_{k}\, \frac{\dd(\frakv \frakA)}{\dd \


dot{q}_{k}}

= \frac{e}{c} (\frakv \frakA).

}{%

\begin{aligned}

E - \phi(p', q)

&= -\sum_{1}^{3} \frac{\dd \phi}{\dd p_{k}'}(p_{k}' - p_{k})

= \sum_{1}^{3} \frac{\dd E}{\dd p_{k}'}\, \frac{e}{c}\, \frac{\dd(\frakv \


frakA)}{\dd \dot{q}_{k}} \\

&= \frac{e}{c} \sum_{1}^{3} \dot{q}_{k}\, \frac{\dd(\frakv \frakA)}{\dd \


dot{q}_{k}}

= \frac{e}{c} (\frakv \frakA).

\end{aligned}

308
}

\]

From this it follows that, with neglect of small quantities proportional to the

square of the magnetic forces, the perturbations of the orbit of the electron in a

hydrogen atom, which besides to a small external electric field of potential is

exposed to a small external magnetic field of vector potential~$\frakA$, are given


by a

set of equations of the same form as~\Eq{(44)} in~\SecRef{2}, but where the $\
alpha$'s~and~$\beta$'s are

replaced by a set of quantities $\alpha_{1}'$,~$\alpha_{2}'$,~$\alpha_{3}'$, $\


beta_{1}'$,~$\beta_{2}'$,~$\beta_{3}'$, which are related to the $q$'s

and~$p'$'s and the time in the same way as the orbital constants $\alpha_{1}$,~$\
alpha_{2}$,~$\alpha_{3}$, $\beta_{1}$,~$\beta_{2}$,~$\beta_{3}$

for the undisturbed atom are related to the $q$'s and~$p$'s and the time through
the

equations~\EqI{(18)}, and where $\Omega$~is replaced by the expression $-e\Phi


+ \dfrac{e}{c} (\frakv \frakA)$, considered

as a function of the $\alpha'$'s~and $\beta'$'s and the time. Since now, at any
moment, the

quantities $\alpha_{1}'$,~$\alpha_{2}'$,~$\alpha_{3}'$, $\beta_{1}'$,~$\


beta_{2}'$,~$\beta_{3}'$ differ from the corresponding orbital constants $\
alpha_{1}$,~$\alpha_{2}$,~$\alpha_{3}$, $\beta_{1}$,~$\beta_{2}$,~$\
beta_{3}$ only by small terms proportional to the intensity of the magnetic

field, we see therefore that, with neglect of small quantities of the same order as

the variation in the orbital constants within a single period, the \Emph{secular
perturbations
309
of the shape and position of the orbit of the electron will

again be given by the equations~\Eq{(46)}}, if in the present case $\psi$~is taken


equal

to the sum of the mean value~$\psi_{E}$ of the potential energy~$-e\Phi$ of the


electron

relative to the external electric forces and the mean value~$\psi_{M}$ of the
quantity~$\dfrac{e}{c} (\frakv \frakA)$,

both taken over an osculating orbit corresponding to some moment during

the revolution and expressed as functions of $\alpha_{1}$,~$\alpha_{2}$,~$\


alpha_{3}$, $\beta_{1}$,~$\beta_{2}$,~$\beta_{3}$.\footnote

{If the relativity modifications are taken into account, the orbit of the electron in
the undisturbed

atom is not strictly periodic, but it will be seen that the secular variations of this
orbit are

still obtained from the equations~\Eq{(46)}, if only, to the expression for~$\psi$


as defined in the text, a term is

added which is equal to the expression for~$\psi$ given by formula~\Eq{(70)}


in~\SecRef{3}.}

The latter mean

value, however, is easily seen to allow of a simple interpretation. In fact, we have

\[

\psi_{M} = \frac{e}{c}\, \frac{1}{\sigma} \int_{0}^{\sigma} (\frakv \frakA)\, dt

= -\frac{e\omega}{c} B,

\Tag{(84)}

\]

310
where $\omega$~is the frequency of revolution of the electron in the osculating
orbit, and

where $B$~represents the total flux of magnetic force through this orbit, taken in

\PageSep{88}

the same direction as that of the magnetic force which would arise from the motion

of the electron according to ordinary electrodynamics.

From the considerations in~\SecRef{2} it follows now in the first place that, with

neglect of small quantities proportional to the square of the external forces, $\psi
= \psi_{E} + \psi_{M}$

will remain \Emph{constant} during the perturbations within a time interval,


sufficiently

long for the perturbing forces to produce a considerable change in the shape and

position of the orbit of the electron; \ie\ in a time interval of the same order as~$\
sigma/\lambda$,

if $\lambda$, just as in~\SecRef{2}, denotes a small quantity of the same order as


the ratio

between the external forces acting on the electron and the attraction from the
nucleus.

From a consideration analogous to that given in~\SecRef{2}, we may further


conclude that,

in the stationary states of the perturbed system, the quantity $\psi = \psi_{E} + \
psi_{M}$ may

be taken equal to the \Emph{additional energy} of the system due to the presence
of the

external fields. In fact, let us imagine that these fields are slowly established at a

311
uniform rate within a time interval from $t = 0$ to $t = \theta$, where $\theta$~is a
quantity

of the same order as~$\sigma/\lambda$. For the total alteration in the inner energy
of the system

during this process we get then, with neglect of small quantities proportional to~$\
lambda^{2}$,

\[

\Delta_{\theta} \alpha_{1}

= e\int_{0}^{\theta} \frac{t}{\theta} \sum_{1}^{3} \frac{\dd \Phi}{\dd


q_{k}}\, \dot{q}_{k}\, dt

- \frac{e}{c} \int_{0}^{\theta} \frac{\omega B}{\theta}\, dt,

\]

where the first term represents the work done on the system by the slowly
increasing

external electric forces, while the second term represents the work performed

by the induced electric forces which accompany the variation in the intensity

of the magnetic field. By partial integration of the first term, we get from

this equation, with the approximation under consideration,

\[

\BreakMath{%

\Delta_{\theta} \alpha_{1} - e\Phi_{\theta}

= -\frac{e}{\theta} \int_{0}^{\theta} \left(\Phi + \frac{\omega}{c}\, B\right) dt

= \frac{1}{\theta} \int_{0}^{\theta} (\psi_{E} + \psi_{M})\, dt

= \frac{1}{\theta} \int_{0}^{\theta} \psi\, dt.

312
}{%

\begin{aligned}

\Delta_{\theta} \alpha_{1} - e\Phi_{\theta}

&= -\frac{e}{\theta} \int_{0}^{\theta} \left(\Phi + \frac{\omega}{c}\, B\right)


dt \\

&= \frac{1}{\theta} \int_{0}^{\theta} (\psi_{E} + \psi_{M})\, dt

= \frac{1}{\theta} \int_{0}^{\theta} \psi\, dt.

\end{aligned}

\Tag{(85)}

\]

Now the expression on the left side of this equation is equal to the change in the

total energy of the system due to the establishment of the external field. Since the

expression on the right side is seen to be a small quantity of the same order as~$\
lambda \alpha_{1}$,

it follows therefore from\Eq{(85)} in the first place that the secular variations of

$\alpha_{2}$,~$\alpha_{3}$, $\beta_{2}$,~$\beta_{3}$ during the increase of the


fields will, just as in the case considered

in~\SecRef{2} (see \PageRef[page]{47}), be given by a set of equations of the


same form as (46), where

$\psi$~is replaced by~$\dfrac{t}{\theta} \psi$, and where again $\


alpha_{1}$~may be considered as a constant. Also in

the present case it follows therefore that $\psi$~will remain constant during the
establishment

of the external fields, and we see consequently that the expression on the

313
right side of~\Eq{(85)} will be simply equal to~$\psi$, a result which, with
reference to the

principle of the mechanical transformability of the stationary states, leads to the

\PageSep{89}

conclusion mentioned above, that the value of the additional energy in the
stationary

states of the perturbed system is given by the value of~$\psi$ in these states.

From the above considerations it follows that the problem of the stationary

states of the hydrogen atom in the presence of external electric and magnetic forces

may be treated in a manner, which is exactly analogous to that applied in~\


SecRef{2}

in case of a periodic system exposed to a small external field of constant potential.

Thus, if the secular perturbations as determined by~\Eq{(46)} are of conditionally


periodic

type, we shall expect that, with neglect of small quantities proportional to~$\
lambda$, the

cycles of shapes and positions which the orbit of the electron passes through

in the stationary states of the perturbed system will be characterised by the

conditions~\Eq{(55)}, and that the possible values of the additional energy of the

atom in the stationary states will be fixed by these conditions with neglect of small

quantities proportional to~$\lambda^{2}$. We shall therefore conclude that, also


in the presence

of external magnetic forces, the lines of the hydrogen spectrum will, if only the
secular

314
perturbations are of conditionally periodic type, be split up in a number of sharp

components, the frequencies of which are determined by means of the conditions~\


Eq{(67)}

together with relation~\EqI{(1)}. As regards the problem of the intensities and


polarisation

of these components, we may further proceed in a way quite analogous to that


followed

in~\SecRef{2}. In fact, if the secular perturbations are of conditionally periodic


type,

the displacement of the electron in any given direction may be represented as a

sum of harmonic vibrations by an expression of the same type as~\Eq{(65)}.


Moreover it

can be proved that the difference in the total energy of two neighbouring states

of the perturbed atom will again be given by the expression~\Eq{(66)}.\footnote

{Compare Note on \PageRef[page]{58}. Also in the presence of small magnetic


forces, it will be possible

to describe the motion of the perturbed system by means of a suitably chosen set
of angle variables,

if only the secular perturbations are of conditionally periodic type.}

The general

considerations in~\SecRef{2} will therefore apply without alterations to the


problem of

the intensity and polarisation of the components into which the hydrogen lines

are split up in the presence of small external forces, also if these forces are entirely

or partly of magnetic origin. Similarly, it will be seen that the effect on the
spectrum

315
of a perturbed hydrogen atom, which will be due to the presence of a second

external field small compared with the first, also in this case may be discussed

directly by means of the considerations at the end of~\SecRef{2}.

We meet with a direct application of the preceding considerations, if the

hydrogen atom is exposed to \Emph{the simultaneous influence of an external

electric and an external magnetic field, which possess axial symmetry

round a common axis through the nucleus}. Introducing the same

set of orbital constants as described in~\SecRef{2} on \PageRef[page]{54}, we get


in this case that

$\psi_{M}$, as well as~$\psi_{E}$, and consequently the function $\psi = \


psi_{E} + \psi_{M}$ which enters in the

equations~\Eq{(46)}, will, besides on~$\alpha_{1}$, depend on $\alpha_{2}$,~$\


beta_{2}$ and~$\alpha_{3}$ but not on~$\beta_{3}$. The general

character of the secular perturbations of the orbit of the electron will therefore be

the same as in the case, considered in~\SecRef{2}, where the atom is exposed only
to an

\PageSep{90}

electric field of axial symmetry, and the conditions which fix the stationary states
of the

perturbed atom will again be expressed by the relations~\Eq{(61)}. As regards the


question

of the probability of spontaneous transition between the stationary states, we get

moreover, just as in~\SecRef{2}, from a consideration of the harmonic vibrations


into which

316
the motion of the electron can be resolved, that only two types of transitions will

be possible; in transitions of the first type $\frakn_{2}$~remains unaltered, and the


accompanying

radiation is polarised parallel to the direction of the common axis of the

perturbing fields; in transitions of the second type $\frakn_{2}$~decreases or


increases by one

unit, and the accompanying radiation will be circularly polarised in a plane


perpendicular

to this axis. In this connection it may be remarked, however, that the

number of components, into which a given hydrogen line is split up in the presence

of a magnetic field, will in general be double as large as the number of components

which appear in the presence of an external electric field of axial symmetry. In

fact, in the latter case the motions of the electron in two stationary states of the

perturbed atom, corresponding to the same value of~$n$, will be symmetrical with

respect to a plane through the axis, and these states will possess the same values

for the additional energy, if $\frakn_{1}$~is the same while the values of~$\
frakn_{2}$ are numerically

equal but have opposite signs. On the other hand, if the atom is exposed also to

a magnetic field, this will not hold, because the value of the function~$\psi_{M}$,
in contrast

to the value of~$\psi_{E}$, will not possess the same sign for two orbits which
have

the same shape and position relative to the axis, but for which the direction of

revolution of the electron is reversed. Considering two states of the perturbed atom

317
for which the values of~$\frakn_{1}$ are the same and the values of~$\frakn_{2}$
are numerically equal

but have opposite signs, we get therefore, if the atom is exposed only to a magnetic

field of axial symmetry, that the values of the additional energy will be equal

with exception of the sign; while, if the atom is exposed to a magnetic as well

as to an electric field, the additional energy in two such states will in general differ

also as regards its numerical value. In contrast to what in general will take place

if the atom is exposed to an electric field of axial symmetry, it will thus be seen

that, if the hydrogen atom is exposed only to a magnetic field possessing axial

symmetry, the ensemble of components into which a given hydrogen line is


resolved

will be completely symmetrical with respect to the position of the original line, as

regards the frequencies as well as the intensities and polarisations. Moreover it


follows

from the above, that if we consider a hydrogen atom exposed to an electric

field of axial symmetry and imagine that an external magnetic field, which
possesses

symmetry round the same axis, is gradually established, each component which

appears in the presence of the first field only will split up into two components,

in such a way that each component polarised parallel to the axis will split up into

two components of the same polarisation, while each component polarised


perpendicular

to the axis, and which originally showed no polarisation when viewed

in a direction parallel to the axis, will split up into two components showing

318
circular polarisations in opposite directions. If the magnetic field is small, the new

\PageSep{91}

components will be placed symmetrically with respect to the position of the


original

components and their intensities will be approximately equal, but when the
perturbing

influence of the magnetic forces on the motion of the electron becomes of

the same order of magnitude as that of the external electric forces, the components

in question will in general be placed unsymmetrically with respect to their original

position, and their intensities may differ considerably.

An\Pagelabel{91} especially simple example of a magnetic field which possesses


axial symmetry

is afforded by the case of a \Emph{homogeneous magnetic field}, discussed in the

beginning of this section. In this case we have that the total magnetic flux of force

through the orbit of the electron is equal to the product of the intensity~$H$ of the

magnetic field and the area of the projection of the orbit on a plane perpendicular

to this field. Since this area is equal to~$\alpha_{3}/2m\omega$, we get


consequently from~\Eq{(84)}

\[

\psi_{M} = \frac{e\alpha_{3}}{2cm} H.

\Tag{(86)}

\]

319
From the equations~\Eq{(46)} it follows therefore that the effect of a
homogeneous magnetic

field, which acts upon a hydrogen atom which at the same time is exposed

to an external electric field possessing axial symmetry round an axis through the

nucleus parallel to the magnetic force, will consist in a superposition of a uniform

rotation of the orbit round the axis with a frequency equal to

\[

\frakv_{H}

= \frac{1}{2\pi}\, \frac{\dd \psi_{M}}{\dd \alpha_{3}}

= \frac{e}{4\pi mc} H

\]

on the secular perturbations which would take place in the absence of the magnetic

field. This result follows also directly from \Name{Larmor}'s theorem, on which
the simple

considerations about the effect of a homogeneous magnetic field in the beginning

of this section were based. Since a superposed rotation as that in question will

not influence the shape of the orbit of the electron or its position relative to the

axis, it follows from~\Eq{(61)} that the value of~$\psi_{E}$ in the stationary


states of the atom

will not be affected by the presence of the magnetic field, and that consequently
the

effect of this field on the additional energy of the system will simply consist in the

addition of a term given by

\[

320
\psi_{M} = \frac{e}{2mc}\, \frac{\frakn_{2} h}{2\pi} H

= \frakn_{2} \frakv_{H} h.

\Tag{(87)}

\]

This result was also to be expected from a simple consideration of the mechanical

effect produced on the motion by a slow and uniform establishment of the


magnetic

field (compare \PageRef[page]{81}). With reference to the above considerations as


regards

the probability of transition between stationary states, it will be seen to follow


from~\Eq{(87)},

that the presence of the homogeneous magnetic field will leave the components

polarised parallel to the axis unaltered, but will cause every component, which in

the absence of the field was polarised perpendicular to the axis, to split up in a

\PageSep{92}

symmetrical doublet the members of which will show circular polarisation in


opposite

directions, when viewed in the direction of the axis, and will be displaced

from the position of the original component by an amount corresponding to a

frequency difference equal to~$\frakv_{H}$.

A simple application of the last result is afforded by the problem of \Emph{the

simultaneous effect on the hydrogen lines of a homogeneous electric

and a homogeneous magnetic field which have the same direction}.

321
Thus, if the intensities of the fields are so large that we may look apart from the

small modifications claimed by the theory of relativity, we shall from the above
expect

that the effect in question will differ from the ordinary \Name{Stark} effect of the
hydrogen

lines, only therein that every component polarised perpendicular to the field is split
up

in two symmetrical components corresponding to the outer members of a \


Name{Lorentz}

triplet. This seems to agree with some observations of the effect of two such fields

on the hydrogen line~$H_{\alpha}$, published by \Name{Garbasso}.\footnote

{\Name{A.~Garbasso}, Phys.\ Zeitschr.\ XV, p.~123 (1914).}

The problem in question might

also have been treated by means of the method of separation of variables, because,

as may be easily shown, the perturbed system---if the relativity modifications

are neglected---allows of separation of variables in parabolic coordinates, just as

in the presence of the electric field only. If, on the other hand, the relativity
modifications

are taken into account, the method of separation of variables cannot be applied,

but, with reference to the considerations at the end of the last section, it will

be seen that it is possible, also in this case, to predict at once the modification in

the effect of an electric field on the fine structure of the hydrogen lines, which

would result from the simultaneous presence of a parallel magnetic field. Passing

to the limiting case where the intensity of the electric field is equal to zero, it will

322
thus be seen at once from the preceding, that \Emph{the effect of a homogeneous

magnetic field on the fine structure of the hydrogen lines} will consist

in the splitting up of every component in a normal \Name{Lorentz} triplet. As far


as

the frequencies of the components are concerned, this result has been predicted by

\Name{Sommerfeld} and \Name{Debye}, who have treated the problem under


consideration by

means of separation of variables in polar coordinates (compare \PageRef[page]


{84}). In connection

with the fixation of the stationary states in this problem, it may be remarked

that we must assume that no stationary state will exist for which the angular

momentum round an axis through the nucleus parallel to the magnetic field would

be equal to zero. In fact, as seen in~\SecRef{4}, we must assume that, in case of a


hydrogen

atom exposed to a homogeneous electric field, no such states will be possible; and

by imagining that the electric field decreases slowly to zero, while at the same

time a magnetic field parallel to the electric field is slowly established, it would

be possible, without passing through a degenerate system, to obtain a continuous

transformation of the stationary states of the perturbed atom during which the

angular momentum of the electron round the axis would remain unaltered. With

\PageSep{93}

reference to the invariance of the a-priori probability of the stationary states during

such a transformation (see Part~I, \PageRefI[page]{9}~and~\PageRefI[]{27}), we


must therefore conclude that,

323
also in the case of a hydrogen atom in the presence of a magnetic field, no
stationary

states exist for which the angular momentum round the axis would be equal to

zero, although these slates in mechanical respect do not exhibit singularities from

which we might anticipate that they are physically unrealisable.\footnote

{\Emph{Note added during the proof.} In a dissertation which has just


appeared, \Name{J.~M. Burgers} (Het

Atoommodel van \Name{Rutherford}-\Name{Bohr}, Haarlem 1918) has given a


very interesting general survey of the

applications of the quantum theory to the problem of the constitution of atoms,


and has in this connection

entered upon several of the questions discussed in the present paper; for instance
on the

question of the relation between the spectrum of an atomic system, deduced by


application of relation~\EqI{(1)}

from the values of the energy in the stationary states, and the frequencies of the
harmonic vibrations

into which the motion in these states can be resolved; and on the question of the
determination

of the relative values for the a-priori probability of the different stationary states
of an atomic system

by means of \Name{Ehrenfest}'s principle of the invariance of these values


during a continuous transformation

of the system. As an illustration of the latter considerations, \Name{Burgers} has


deduced an expression

for the relative values of the a-priori probability of the different stationary states
of the undisturbed

324
hydrogen atom, by means of an enumeration of the states, determined by the
conditions~\EqI{(22)} when

applied in connection with a separation of variables in polar coordinates, which


correspond to a stationary

state of the undisturbed atom, characterised by a given value of~$n$ in the


condition $I = nh$.

Excluding only such states for which the total angular momentum of the electron
round the nucleus

would be equal to zero, \Name{Burgers} (loc.\ cit.\ p.~259) finds in this way for
the value of the a-priori probability

in question $(n + 1)^{2} - 1$. In connection with the analogous consideration,


given in the Note on

\PageRef[page]{76} of the present paper, which leads to a different result, it may


be of interest to remark that

the necessary conformity between the relative values for the a-priori probability
of the different stationary

states of the undisturbed hydrogen atom, deduced from an enumeration of the


stationary states

of the atom which appear in the presence of a small external electric field or in
the presence of a

small magnetic field respectively, cannot be obtained if in both cases we would


exclude only such

states in which the angular momentum of the electron round the nucleus is always
equal to zero. In

fact, while in case of a magnetic field this would give $(n + 1)^{2} - 1$ different
states corresponding to

a given value of~$n$, it would in case of an electric field give only $(n + 1)^{2} -
2$ such states. On the other
325
hand, if the possible stationary states are selected in the manner explained in the
text, the conformity

in question will obviously be obtained.}

In case we consider the general problem of the effect on a hydrogen atom of

a small electric or magnetic field, which do not possess axial symmetry round an

axis through the nucleus, or of the simultaneous effect of two such fields, which

do not possess such symmetry round a common axis, we must expect that the

secular perturbations of the orbit of the electron will in general not be of


conditionally

periodic type. In such a case we cannot obtain a complete fixation of the

stationary states, and we may conclude that the presence of the external forces will

not give rise to the splitting up of the hydrogen lines into a number of sharp

components but to a diffusion of these lines. With a simple example, in which the

secular perturbations of the atom seem not to be of conditionally periodic type, we

meet if we consider \Emph{the simultaneous effect on the hydrogen spectrum of

an external homogeneous electric field and a homogeneous magnetic

field, the directions of which make an angle with each other}. If

% [** TN: Printer's mark]

% D. K. D. Vidensk. Selsk. Skr., naturvidensk. og mathem. Afd, 8. Række, IV. 1.


13

\PageSep{94}

the effects of the two fields on the motion of the electron are of the same order

of magnitude we may in this case expect that the hydrogen lines will not be

326
resolved into sharp components but will become diffuse. From the considerations

on \PageRef[page]{60} of the effect on the spectrum of a perturbed periodic


system due

to a second external field, the perturbing effect of which is small compared with

that of the first, we may conclude, however, that, if the effect of one of the fields
on

the motion of the electron is large compared with that of the other, the hydrogen

lines will still show a resolution in a number of components, the spectral widths

of which are small compared with the displacements which they have undergone
due to

the presence of the weaker of the external fields. In the discussion of this problem

we shall for simplicity neglect the influence of the relativity modifications,


assuming

that the effect on the spectrum produced by each external field separately is large

compared with the inherent fine structure of the hydrogen lines. Denoting, as in~\
SecRef{2},

by~$\mu$ a small constant of the same order as the ratio between the forces on

the electron due to the weaker of the external fields and those due to the stronger

of these fields, and by~$\lambda$ a small constant of the same order as the ratio
between

the latter forces and the attraction from the nucleus, we have, as shown on \
PageRef[page]{61},

that, with neglect of small quantities of the same order of magnitude as~$\lambda\
mu^{2}$,\footnote

{Rigorously this result holds with neglect of small quantities of the same order of
magnitude

327
as the largest of the quantities $\lambda^{2}$ and~$\lambda\mu^{2}$, but for the
sake of simplicity it is here and in the following

assumed that $\mu$~is not smaller than~$\sqrt{\lambda}$ (compare \


PageRef[page]{61}).}

the change in the additional energy of the atom due to the presence of the weaker

field is, in general, directly obtained by taking the mean value of the function~$\
psi$,

corresponding to the weaker field, over the cycle of shapes and positions which

the orbit of the electron passes through in the stationary states of the atom in the

presence of the stronger field only. In the special case under consideration,
however,

the perturbed system, formed by the atom in the presence of the stronger field

only, is degenerate, the secular perturbations of the orbit of the electron being of a

simple periodic character. The mean value in question will therefore not be
completely

determined, but will be different for the different periodic cycles of shapes and

positions of the orbit, which represent the continuous multitude of stationary


motions

which the electron may perform in each of the stationary states of the atom in

the presence of the stronger field only. In order to fix the stationary states in the

presence of both fields and the change in the additional energy of the atom due to

the presence of the weaker field, it will thus, as mentioned on \PageRef[page]{62},


be necessary

to examine the relation between the mean value in question and the frequency

of the slow periodic ``secular'' variations which the cycles under consideration will

328
undergo under the influence of the weaker of the external fields. Now, in the

special case under consideration this problem may be treated very simply, if we

imagine the weaker field as composed of two homogeneous fields of which the one

is parallel and the other perpendicular to the stronger field, and if we consider

separately the secular effect due to each of these fields. In fact, due to the
symmetry

\PageSep{95}

with respect to the axis of the stronger field, exhibited by the periodic cycle of

shapes and positions which the orbit of the electron would pass through if the

atom were exposed to this field only, it is easily seen that the contribution, which

the perpendicular component of the weaker field gives to the mean value of~$\psi$

corresponding to the latter field, will vanish. From this it follows that the secular
effect

of the weaker field, with neglect of small quantities proportional to~$\mu^{2}$


will be

the same as if only the parallel component of this field was acting on the atom;

and we see consequently that, in the stationary states of the atom in the presence

of both fields, the possible cycles of shapes and positions of the orbit of the
electron

will be characterised in the same way as if the weaker field was parallel to the

stronger. The problem, however, of the fixation of the stationary states of a


hydrogen

atom in the presence of a homogeneous electric field and a homogeneous magnetic

field, which are parallel to each other, is very simple. In fact, as it appears from the

329
considerations on \PageRef[page]{91}, the stationary states will in this case be
fixed completely

by two conditions, of which the one, in the same way as in the simple theory of

the \Name{Stark} effect, defines the position of the plane in which the electrical
centre of

\Typo{he}{the} orbit of the electron moves, while the other defines the value of
the angular

momentum of the electron round the axis of the fields in the same way as in the

simple theory of the \Name{Zeeman} effect. In connection with the problem under
consideration

here, it may be useful for the sake of illustration to note, that, if the

perturbing effect of the electric field is large compared with that of the magnetic,

the second of these conditions may be said to be imposed on the system by the

slow and uniform rotation, which the magnetic field produces on the periodic cycle

of shapes and positions of the orbit of the electron, which would appear if the

atom was exposed to the electric field only. If, on the other hand, the effect of the

magnetic field is large compared with that of the electric field, the first condition

may be said to be imposed on the system by the slow periodic oscillation in the

shape and position relative to the axis, which the electric field produces on the

uniformly rotating orbit which the electron would describe if the atom was exposed

to the magnetic field only.

If we consider a hydrogen atom which is exposed to the simultaneous influence

of a homogeneous electric field of intensity~$F$ and a homogeneous magnetic

330
field of intensity~$H$, the direction of which makes an angle~$\phi$ with the
direction of

the electric field, it follows from the above that, if the perturbing influence of the

electric field is large compared with that of the magnetic field, the main effect
produced

by the latter field on the spectrum may be described as the splitting up of

each \Name{Stark} effect component, polarised perpendicular to the axis of the


electric field,

into two circularly polarised components, corresponding to the outer members of a

\Name{Lorentz} triplet which would be produced by a magnetic field of intensity


$H\cos \phi$.

On the other hand, if the perturbing effect of the magnetic field is large compared

with that of the electric, it follows that the main effect, produced by the latter field

on the spectrum, may be described as the resolution of the middle component and

\PageSep{96}

of each of the outer components of the normal \Name{Zeeman} effect into a


number of

components, corresponding to the parallel and perpendicular components


respectively

of a \Name{Stark} effect produced by an electric field of intensity $F\cos \phi$.

The effects just described, however, which are the same as would take place

if only the parallel component of the weaker field was acting on the atom, will

not be the only effects of the presence of the weaker field on the spectrum. In

fact, although the perpendicular component of the weaker field, apart from small

331
quantities proportional to~$\mu^{2}$, will not have any secular effect on the cycle
of shapes

and positions which the orbit of the electron would pass through if the atom was

exposed to the stronger field only, it will obviously produce alterations in the

motion of the electron within this cycle which are proportional to~$\mu$. Thus, if
the

weaker field was parallel to the stronger, the motion of the electron in the
perturbed

atom would be composed of a number of linear harmonic vibrations parallel

to the direction of the fields, the frequencies of which are of the type $|\tau \
omega_{P} + \frakt_{1} \frakv_{1}|$,

and of a number of circular harmonic rotations perpendicular to this direction, the

frequencies of which are of the type $|\tau \omega_{P} + \frakt_{1} \frakv_{1} + \


frakv_{2}|$ (compare \PageRef[page]{59}). In the

general case, however, where the weaker field is not parallel to the stronger, there

will, in the expression for the displacement of the electron in any given direction,

in addition appear a number of harmonic vibrations the amplitudes of which are

proportional to~$\mu$ and the frequencies of which, as a closer consideration of


the

perturbations learns, are equal to the sum or difference of the frequency of one of

the harmonic vibrations, in which the motion in this direction could be resolved if

the external fields were parallel to each other, and one of the small frequencies of

type $|\frakt_{1} \frakv_{1} + \frakv_{2}|$, which appear in the expression for the
secular perturbations of the

332
electron in this case. A part of these additional vibrations will again possess
frequencies

of the types $|\tau \omega_{P} + \frakt_{1} \frakv_{1}|$ and $|\tau \omega_{P}


+ \frakt_{1} \frakv_{1} + \frakv_{2}|$, and will cause that the motion,

instead of consisting of vibrations which are exactly linear and exactly circular

as in the case where the external fields are parallel to each other, will be composed

of elliptical harmonic vibrations which partly are nearly linear and parallel

to the direction of the stronger field and partly nearly circular and perpendicular

to this direction. On account of this we shall expect that, due to the

presence of the perpendicular component of the weaker field, the different


components

mentioned above will not be sharply polarised. Further there will, in the

motion of the perturbed atom, also appear a number of circular harmonic rotations

perpendicular to the stronger field, the amplitudes of which are small quantities

proportional to~$\mu$, and the frequencies of which are of the type $|\tau \
omega_{P} + \frakt_{1} \frakv_{1} + 2\frakv_{2}|$.

From this we shall expect the appearance in the spectrum of a number of new

weak components, corresponding to a type of transition between stationary states

which would not be possible if the two external fields were parallel to \
Typo{eachother}{each other}.

When considering more closely the frequencies of these new components, it must

be remembered, however, that, as mentioned above, the present method of treating

the problem of the perturbations assures us of the conditionally periodic character

\PageSep{97}

333
of the motion of the electron within a time interval of the same order of

magnitude as~$\sigma/\lambda$, only if we look apart from small quantities of the


same order as~$\mu^{2}$;

and we must therefore be prepared to find that the frequencies of the vibrations

of small amplitudes will not be defined with the same degree of approximation

as the frequencies of the vibrations of large amplitudes. Thus, while the


frequencies

of the latter vibrations are defined with neglect of small quantities proportional

to~$\lambda\mu^{2}$, the frequencies of the small vibrations under consideration


are

obviously defined only with neglect of small quantities proportional to~$\lambda\


mu$. In intimate

connection with the general want of definition of the energy in the stationary

states for perturbed systems of the type in question, we must accordingly be


prepared

to find that, in contrast to the strong components, for which we may expect

that by far the larger part of the intensity is contained within a spectral interval

of a width proportional to~$\lambda\mu^{2}$, the new components will be


diffused over spectral

intervals of a width proportional to~$\lambda\mu$.\Pagelabel{97}\footnote

{Compare Note on \PageRef[page]{61}. With reference to the general validity of


relation~\EqI{(1)}, it will be seen

that the assumption, that the weak components possess this degree of diffusion,
implies the assumption,

that the corresponding transitions (the probability of occurrence of which is very


small compared with

334
the probability of the transitions responsible for the strong components) will
generally take place

between two states of the perturbed atom, which do not both belong to the well
defined ensemble of

stationary states in which at any moment the great majority among a large number
of atoms will be

present. Thus, in case the effect of the external electric field is large compared
with that of the magnetic

field, we may expect that, in both states involved in the transitions in question, the
positions of

the plane in which the electrical centre moves will coincide with positions of this
plane in states

belonging to the ensemble just mentioned, while the angular momentum of the
electron round the

axis of the electric field will generally change by an amount which will not be
equal to an entire

multiple of~$h/2\pi$. On the other hand, if the effect of the magnetic field is the
larger, the angular

momentum of the electron round the axis of this field will, in the transitions in
question, change by

two times~$h/2\pi$, while we may expect that the plane in which the electrical
centre moves will generally,

in at least one of the states involved in these transitions, differ from the positions
of this plane in the

ensemble of stationary states referred to.}

Thus, in case the effect of the external

electric field is large compared with that of the magnetic field, we might expect at

335
first sight that, on each side of every of the \Name{Stark} effect components
polarised

parallel to the electric force, there would appear a weak component which would
be

circularly polarised and be displaced from this component by an amount twice that

of the displacement of the strong components into which the perpendicularly

polarised \Name{Stark} effect components are split up as a consequence of the


small

magnetic field. We must be prepared, however, to find that these weak components

will be so diffuse, that they are not separated from the weak perpendicular
component

which has the same frequency as the strong parallel components on each

side of which the weak components under consideration would lie, and which

appears as a consequence of the above mentioned want of sharpness as regards the

polarisation of the strong components. On the other hand, if the effect of the
magnetic

field is large compared with that of the electric field, any weak component

of the type under consideration, which corresponds to transitions in which the

\PageSep{98}

angular momentum of the electron round the axis of the magnetic field changes

by two times~$h/2\pi$, will lie at a distance from the original hydrogen line, which
is

approximately twice as large as that of the outer components of the normal \


Name{Zeeman}

effect, and will therefore be distinctly separated from the strong components into

336
which each of the components of the normal \Name{Zeeman} effect is split up in
the presence

of the small electric field. We must be prepared, however, to find that the

weak components will not, as it might be expected at first sight, form two sets of

distinctly separated lines, but that they will only appear as two diffuse lines of
circular

polarisation in opposite directions and of a spectral width proportional to~$\


lambda\mu$.\footnote

{No experiments, which allow to test the preceding results in detail, seem to have
been recorded,

but it would appear that the above considerations afford an explanation of the
general character of the

remarkable deviations from a normal \Name{Zeeman} effect, observed by \


Name{F~Paschen} and \Name{E.~Back} (Ann.\ d.\ Phys.\

XXXIX, p.~897 (1912)) in experiments in which the hydrogen lines were excited
by passing a powerful

condensed discharge through a capillary tube placed at right angles with the
direction of the magnetic

field. Besides the characteristic want of sharpness of the polarisation of the


middle component,

exhibited by all the spectrograms published by \Name{Paschen} and \


Name{Back}, especially one of their photographs

(Tafel VIII, Bild~4) seems to suggest the presence of a weak, perpendicularly


polarised, diffuse line on

each side of the original line and at a distance from it twice that of the outer
components of the

normal effect.}

337
\Section{6.}{The continuous hydrogen~spectrum.}

We shall conclude the considerations of this Part by a brief discussion of the

characteristic continuous spectrum of hydrogen in the ultra violet region, which is

intimately connected with the series spectrum given by~\Eq{(35)}. This spectrum
consists

of a radiation, the frequencies of which are continuously distributed over a spectral

interval extending from the head of the Balmer series in the direction of higher

frequencies.\footnote

{This spectrum has been observed as an emission spectrum in spectra of solar


protuberances

and planetary nebulae (See \Name{J.~Evershed}, Phil.\ Trans.\ Roy.\ Soc.\


197~A, p.~399 (1901) and \Name{W.~H. Wright},

Lick Observatory Bulletin, No.~291 (1917)) as well as in direct laboratory


experiments on spectra excited

by positive rays (See \Name{J.~Stark}, Ann.\ d.\ Phys.\ LII, p.~255 (1917)).
Further it has been observed

as an absorption spectrum in the spectra of several stars (see \


Name{W.~Huggins}, An Atlas of Representative

Stellar Spectra, p.~85 (1899) and \Name{J.~Hartmann}, Phys.\ Zeitschr.\ XVIII


p.~429 (1917)).}

The existence of a continuous spectrum of this type is just what

should be expected from a natural generalisation of the principles underlying the

338
quantum theory of series spectra.\footnote

{Compare \Name{N.~Bohr}, Phil.\ Mag.\ XXVI, p.~17 (1913); and also \


Name{P.~Debye}, Phys.\ Zeitschr.\ XVIII, p.~428

(1917).}

Thus the spectrum under consideration may be

directly explained by application of relation~\EqI{(1)}, if we assume that the


complete

spectrum, emitted by a system consisting of a nucleus and of an electron, originates

not only from radiations, emitted during transitions between two states belonging

to the multitude of stationary states in which the electron describes a closed orbit,

characterised by the condition $I = nh$, but also from radiations emitted during

\PageSep{99}

transitions between two states, one (or both) of which belong to the multitude of

states in which the electron possesses sufficient energy to remove to infinite


distance

from the nucleus. While the electron in the states of the type first mentioned may

be said to be ``bound'' by the nucleus to form an atom, it may in the states of the

last mentioned type be described as ``free''. In order to account for the appearance

of the continuous spectrum, it is necessary to assume that the motions

in the latter states are not restricted by extra-mechanical conditions of the type

holding for the former states, but that all motions, which are consistent with the

application of ordinary mechanics, will represent physically possible states. This

assumption would also seem to present itself naturally from the point of view on

the principles of the quantum theory, taken in the present paper.\footnote

339
{A view contrary to this has been taken by \Name{Epstein}, who in a recent
paper (Ann.\ d.\ Phys.\ L,

p.~815 (1916)) has made an attempt to obtain an explanation of certain


observations on the photoelectric

effect of hydrogen occluded in metals, by applying conditions of the same type


as~\EqI{(22)} to

states of the hydrogen atom in which the electron describes a hyperbolic orbit,
and has tried in a

similar way to develop a theory of the characteristic $\beta$-ray spectra of


radioactive substances.}

Thus it will in

the first place be observed that any attempt to discriminate between the different

states of the type in question, by means of considerations of the mechanical


stability

of stationary states for slow transformations of the external conditions, would fail

on account of the essentially non-periodic character of the motion, which is


irreconcilable

with the idea of invariance of extra-mechanical conditions for such


transformations.

Next, with reference to the formal analogy between the quantum theory

and the ordinary theory of radiation, it will be seen that the fact, that the motion

of a free electron in its hyperbolic orbit cannot be resolved in a sum of harmonic

vibrations of discontinuously varying frequencies but can only be represented by

a Fourier integral extended over a continuous range of frequencies, suggests


beforehand

that the free electron may pass, under emission or absorption of radiation, to

340
any one among a continuous multitude of other states corresponding to a
continuous

multitude of values for the energy of the system. From the preceding
considerations

we may infer, by application of~\EqI{(1)}, that the complete spectrum emitted by


the

hydrogen atom will, besides the series spectrum and the continuous ultra-violet

spectrum mentioned above, which corresponds to transitions from a state in which

the electron is free to a stationary state characterised by $n = 2$ in~\Eq{(41)},


contain a

set of continuous spectra, corresponding to transitions from free states to other


stationary

states, and each extending in the direction of larger frequencies from one

of the values of the frequency, given by~\Eq{(35)} if we put $n' = \infty$.


Moreover, we may

expect the presence of a weak continuous spectrum, extending as a continuous


back

ground over the whole region of frequencies, which will correspond to transitions

between two different states in both of which the electron is free. The relative
intensities

of these different continuous spectra, and the laws according to which the intensity

is distributed within each of them, may be expected to vary to a large extent

according to the different conditions under which the radiation is excited. Thus,

\PageSep{100}

while the continuous spectrum of hydrogen, when observed as emission spectrum

341
in stars, shows a abrupt beginning at the head of the Balmer series, the continuous

spectrum, observed by \Name{Stark} in his experiments referred to above, was

not sharply limited but showed a pronounced maximum in the spectral region

which corresponds to transitions between two states, in the first of which the
velocity

of the free electron relative to the nucleus, before the ``collision'' with the

latter, was of the same order of magnitude as the velocity of the positive rays by

means of which the spectrum was excited.

Besides the series spectrum and the connected continuous spectrum just
considered,

there exists, as well known, another hydrogen spectrum, the \Typo{socalled}{so


called} many-line

spectrum, which on account of its complex structure and its resemblance with

the band spectra, emitted by other elements and combinations of elements, is


generally

ascribed to the hydrogen molecule and not to the atom. This assumption would

also seem to present itself directly from the point of view of the quantum theory,

according to which the simple structure of the series spectrum is directly connected

with the simple periodic character of the motion of the particles in the

atom, while a spectrum of a complexity of the order exhibited by the many-line

spectrum must be assumed to originate from a system the motion of which does

not show such simple periodicity properties. The problem of the constitution of the

342
hydrogen molecule, to be expected on the quantum theory, and the possible
motions

of the particles of this system will be treated in Part~IV\@. In this connection we

shall also consider the problem of dispersion of light in hydrogen gas and the
problem

of the voltage necessary to produce the lines of the series spectrum of hydrogen

by an electric discharge in this gas.

\vfill

\begin{center}

\footnotesize

Færdig fra Trykkeriet d.~30, December 1918.

\end{center}

%[** TN: Catalog text not manually inspected]

\iffalse

Det Kgl. Danske Videnskabernes Selskabs Skrifter.

Naturvidenskabelig og: mathematisk Afdeling,

8de Række.

Kr. 0re

343
I, 1915-1917 10. 75,

1. Prytz, K, og J. N. Nielsen: Undersogelser til Fremstilling af Normaler i


Metersystemet, grundet

paa Sammenligning med de danske Rigsprototyper for Kilogrammet og Meteren.


1915 1. 55.

2. Rasmussen, Hans Baggesgaard : Om Bestemmelse af Nikotin i Tobak og


Tobaksextrakter. En

kritisk Undersogelse. 1916 1- 75.

3. Christiansen, M.: Bakterier af Tyfus-Coligruppen, forekommende i Tarmen hos


sunde Spaed-

kalve og ved disses Tarminfektioner. Sammenlignende Undersogelser. 1916 2. 25.

4. Juel, C: Die elementare Ringflache vierter Ordnung. 1916 » 60.

5. Zeuthen, H. G.: Hvorledes Mathematiken i Tiden fra Platon til Euklid blev en
rationel Viden-

skab. Avec un resume en fran§ais. 1917 8. »

344
II, med 4 Tavler, 1916-1918 '11. 50

1. Jergensen, S. M.: Det kemiske Syrebegrebs Udviklingshistorie indtil 1830.


Efterladt Manuskript,

udgivet af Ove Jergensen og S. P. L. Serensen. 1916 3. 45.

2. Hansen-Ostenfeld, Carl: De danske Farvandes Plankton i Aarene 1898 - 1901.


Phytoplankton

og Protozoer. 2. Protozoer; Organismer med usikker Stilling; Parasiter i


Phytoplanktonter. Med

4 Figurgrupper og 7 Tabeller i Teksten. Avec un resume en franjais. 1916 2. 75.

3. Jensen, J. L, W. V.: Undersogelser over en Klasse fundamentale Uligheder i de


analytiske Funk-

tioners Theori. I. 1916 » 90.

4 Pedersen, P. O.: Om Poulsen-Buen og dens Teori. En Experimentalundersogelse.


Med 4 Tav-

ler. 1917 2. 90.

5. Juel, C: Die gewundenen Kurven vom Maximalindex auf einer Rcgelflache


zweiter Ordnung. 1917 » 75.

345
6. Warming, Eug. : Om Jordudlobere. With a Resum^ in English. 1918 3. 65.

III, med 14 Kort og 12 Tavler, 1917-1919 26. 00.

1. Wesenberg-Lund, C: Furesostudier. En bathymetrisk Undersogelse af


Molleaaens Soer. Under

Medvirkning af Oberst M. J. Sand, Mag. /. Boye Petersen, Fru A. Seidelin


Raunkicer og Mag. sc.

C. M. Steenberg. Med 7 bathymetriske Kort, 7 Vegetationskort, 8 Tavler og ca. 50


i Texten trykte

Figurer. Avec un resume en frangais. 1917 22. »

2. Lehmann, Alfr.: Stofskifte ved sjaelelig Virksomhed. With a Resume in English.


1918 3. 15.

3. \Name{Kramers}, H. A.: Intensities of Spectral Lines. On the application of the


Quantum Theory to

the problem of the relative intensities of the components of the fine structure and
of the stark

effect of the lines of the hydrogen spectrum. With 4 plates. 1919 9. 50.

IV (under Pressen).

346
1. \Name{Bohr}, N. : On the Quantum Theory of Line-Spectra. Part 1. 1918 2. 25.

- Samme. Part. II. 1918 4. ^0.

V (under Pressen).

1. Bjerrum, Niels und Kirschner, Aage: Die Rhodanide des Goldes und das freie
Rhodan. Mit

einem Anhang iiber das Goldchlorid. 1918 3. 50.

2. Orla-Jensen, S.: The lactic acid Bacteria. With 51 Plates. 1919 46. 00.

VI (under Pressen).

1. Christensen, Carl: A Monograph of the genus Dryopteris. Part II. 1920 8. 25.

\fi

%%%%%%%%%%%%%%%%%%%%%%%%% GUTENBERG LICENSE %%


%%%%%%%%%%%%%%%%%%%%%%%%

\BackMatter

\PGLicense

\begin{PGtext}

347
End of the Project Gutenberg EBook of On the Quantum Theory of Line-Spectra,

Part 1 and 2, by Niels Henrik David Bohr

*** END OF THIS PROJECT GUTENBERG EBOOK QUANTUM THEORY


OF LINE-SPECTRA ***

***** This file should be named 47167-pdf.pdf or 47167-pdf.zip *****

This and all associated files of various formats will be found in:

http://www.gutenberg.org/4/7/1/6/47167/

Produced by Andrew D. Hwang

Updated editions will replace the previous one--the old editions

will be renamed.

Creating the works from public domain print editions means that no

one owns a United States copyright in these works, so the Foundation

(and you!) can copy and distribute it in the United States without

permission and without paying copyright royalties. Special rules,

set forth in the General Terms of Use part of this license, apply to

copying and distributing Project Gutenberg-tm electronic works to

protect the PROJECT GUTENBERG-tm concept and trademark. Project

Gutenberg is a registered trademark, and may not be used if you

348
charge for the eBooks, unless you receive specific permission. If you

do not charge anything for copies of this eBook, complying with the

rules is very easy. You may use this eBook for nearly any purpose

such as creation of derivative works, reports, performances and

research. They may be modified and printed and given away--you may do

practically ANYTHING with public domain eBooks. Redistribution is

subject to the trademark license, especially commercial

redistribution.

*** START: FULL LICENSE ***

THE FULL PROJECT GUTENBERG LICENSE

PLEASE READ THIS BEFORE YOU DISTRIBUTE OR USE THIS WORK

To protect the Project Gutenberg-tm mission of promoting the free

distribution of electronic works, by using or distributing this work

(or any other work associated in any way with the phrase "Project

Gutenberg"), you agree to comply with all the terms of the Full Project

Gutenberg-tm License available with this file or online at

www.gutenberg.org/license.

349
Section 1. General Terms of Use and Redistributing Project Gutenberg-tm

electronic works

1.A. By reading or using any part of this Project Gutenberg-tm

electronic work, you indicate that you have read, understand, agree to

and accept all the terms of this license and intellectual property

(trademark/copyright) agreement. If you do not agree to abide by all

the terms of this agreement, you must cease using and return or destroy

all copies of Project Gutenberg-tm electronic works in your possession.

If you paid a fee for obtaining a copy of or access to a Project

Gutenberg-tm electronic work and you do not agree to be bound by the

terms of this agreement, you may obtain a refund from the person or

entity to whom you paid the fee as set forth in paragraph 1.E.8.

1.B. "Project Gutenberg" is a registered trademark. It may only be

used on or associated in any way with an electronic work by people who

agree to be bound by the terms of this agreement. There are a few

things that you can do with most Project Gutenberg-tm electronic works

even without complying with the full terms of this agreement. See

paragraph 1.C below. There are a lot of things you can do with Project

Gutenberg-tm electronic works if you follow the terms of this agreement

and help preserve free future access to Project Gutenberg-tm electronic


350
works. See paragraph 1.E below.

1.C. The Project Gutenberg Literary Archive Foundation ("the Foundation"

or PGLAF), owns a compilation copyright in the collection of Project

Gutenberg-tm electronic works. Nearly all the individual works in the

collection are in the public domain in the United States. If an

individual work is in the public domain in the United States and you are

located in the United States, we do not claim a right to prevent you from

copying, distributing, performing, displaying or creating derivative

works based on the work as long as all references to Project Gutenberg

are removed. Of course, we hope that you will support the Project

Gutenberg-tm mission of promoting free access to electronic works by

freely sharing Project Gutenberg-tm works in compliance with the terms of

this agreement for keeping the Project Gutenberg-tm name associated with

the work. You can easily comply with the terms of this agreement by

keeping this work in the same format with its attached full Project

Gutenberg-tm License when you share it without charge with others.

1.D. The copyright laws of the place where you are located also govern

what you can do with this work. Copyright laws in most countries are in

a constant state of change. If you are outside the United States, check

the laws of your country in addition to the terms of this agreement

before downloading, copying, displaying, performing, distributing or


351
creating derivative works based on this work or any other Project

Gutenberg-tm work. The Foundation makes no representations concerning

the copyright status of any work in any country outside the United

States.

1.E. Unless you have removed all references to Project Gutenberg:

1.E.1. The following sentence, with active links to, or other immediate

access to, the full Project Gutenberg-tm License must appear prominently

whenever any copy of a Project Gutenberg-tm work (any work on which the

phrase "Project Gutenberg" appears, or with which the phrase "Project

Gutenberg" is associated) is accessed, displayed, performed, viewed,

copied or distributed:

This eBook is for the use of anyone anywhere at no cost and with

almost no restrictions whatsoever. You may copy it, give it away or

re-use it under the terms of the Project Gutenberg License included

with this eBook or online at www.gutenberg.org

1.E.2. If an individual Project Gutenberg-tm electronic work is derived

from the public domain (does not contain a notice indicating that it is

posted with permission of the copyright holder), the work can be copied

and distributed to anyone in the United States without paying any fees
352
or charges. If you are redistributing or providing access to a work

with the phrase "Project Gutenberg" associated with or appearing on the

work, you must comply either with the requirements of paragraphs 1.E.1

through 1.E.7 or obtain permission for the use of the work and the

Project Gutenberg-tm trademark as set forth in paragraphs 1.E.8 or

1.E.9.

1.E.3. If an individual Project Gutenberg-tm electronic work is posted

with the permission of the copyright holder, your use and distribution

must comply with both paragraphs 1.E.1 through 1.E.7 and any additional

terms imposed by the copyright holder. Additional terms will be linked

to the Project Gutenberg-tm License for all works posted with the

permission of the copyright holder found at the beginning of this work.

1.E.4. Do not unlink or detach or remove the full Project Gutenberg-tm

License terms from this work, or any files containing a part of this

work or any other work associated with Project Gutenberg-tm.

1.E.5. Do not copy, display, perform, distribute or redistribute this

electronic work, or any part of this electronic work, without

prominently displaying the sentence set forth in paragraph 1.E.1 with

active links or immediate access to the full terms of the Project

Gutenberg-tm License.
353
1.E.6. You may convert to and distribute this work in any binary,

compressed, marked up, nonproprietary or proprietary form, including any

word processing or hypertext form. However, if you provide access to or

distribute copies of a Project Gutenberg-tm work in a format other than

"Plain Vanilla ASCII" or other format used in the official version

posted on the official Project Gutenberg-tm web site (www.gutenberg.org),

you must, at no additional cost, fee or expense to the user, provide a

copy, a means of exporting a copy, or a means of obtaining a copy upon

request, of the work in its original "Plain Vanilla ASCII" or other

form. Any alternate format must include the full Project Gutenberg-tm

License as specified in paragraph 1.E.1.

1.E.7. Do not charge a fee for access to, viewing, displaying,

performing, copying or distributing any Project Gutenberg-tm works

unless you comply with paragraph 1.E.8 or 1.E.9.

1.E.8. You may charge a reasonable fee for copies of or providing

access to or distributing Project Gutenberg-tm electronic works provided

that

- You pay a royalty fee of 20% of the gross profits you derive from

the use of Project Gutenberg-tm works calculated using the method


354
you already use to calculate your applicable taxes. The fee is

owed to the owner of the Project Gutenberg-tm trademark, but he

has agreed to donate royalties under this paragraph to the

Project Gutenberg Literary Archive Foundation. Royalty payments

must be paid within 60 days following each date on which you

prepare (or are legally required to prepare) your periodic tax

returns. Royalty payments should be clearly marked as such and

sent to the Project Gutenberg Literary Archive Foundation at the

address specified in Section 4, "Information about donations to

the Project Gutenberg Literary Archive Foundation."

- You provide a full refund of any money paid by a user who notifies

you in writing (or by e-mail) within 30 days of receipt that s/he

does not agree to the terms of the full Project Gutenberg-tm

License. You must require such a user to return or

destroy all copies of the works possessed in a physical medium

and discontinue all use of and all access to other copies of

Project Gutenberg-tm works.

- You provide, in accordance with paragraph 1.F.3, a full refund of any

money paid for a work or a replacement copy, if a defect in the

electronic work is discovered and reported to you within 90 days

of receipt of the work.


355
- You comply with all other terms of this agreement for free

distribution of Project Gutenberg-tm works.

1.E.9. If you wish to charge a fee or distribute a Project Gutenberg-tm

electronic work or group of works on different terms than are set

forth in this agreement, you must obtain permission in writing from

both the Project Gutenberg Literary Archive Foundation and Michael

Hart, the owner of the Project Gutenberg-tm trademark. Contact the

Foundation as set forth in Section 3 below.

1.F.

1.F.1. Project Gutenberg volunteers and employees expend considerable

effort to identify, do copyright research on, transcribe and proofread

public domain works in creating the Project Gutenberg-tm

collection. Despite these efforts, Project Gutenberg-tm electronic

works, and the medium on which they may be stored, may contain

"Defects," such as, but not limited to, incomplete, inaccurate or

corrupt data, transcription errors, a copyright or other intellectual

property infringement, a defective or damaged disk or other medium, a

computer virus, or computer codes that damage or cannot be read by

your equipment.
356
1.F.2. LIMITED WARRANTY, DISCLAIMER OF DAMAGES - Except for the
"Right

of Replacement or Refund" described in paragraph 1.F.3, the Project

Gutenberg Literary Archive Foundation, the owner of the Project

Gutenberg-tm trademark, and any other party distributing a Project

Gutenberg-tm electronic work under this agreement, disclaim all

liability to you for damages, costs and expenses, including legal

fees. YOU AGREE THAT YOU HAVE NO REMEDIES FOR NEGLIGENCE,


STRICT

LIABILITY, BREACH OF WARRANTY OR BREACH OF CONTRACT


EXCEPT THOSE

PROVIDED IN PARAGRAPH 1.F.3. YOU AGREE THAT THE


FOUNDATION, THE

TRADEMARK OWNER, AND ANY DISTRIBUTOR UNDER THIS


AGREEMENT WILL NOT BE

LIABLE TO YOU FOR ACTUAL, DIRECT, INDIRECT, CONSEQUENTIAL,


PUNITIVE OR

INCIDENTAL DAMAGES EVEN IF YOU GIVE NOTICE OF THE


POSSIBILITY OF SUCH

DAMAGE.

1.F.3. LIMITED RIGHT OF REPLACEMENT OR REFUND - If you discover a

defect in this electronic work within 90 days of receiving it, you can

receive a refund of the money (if any) you paid for it by sending a

357
written explanation to the person you received the work from. If you

received the work on a physical medium, you must return the medium with

your written explanation. The person or entity that provided you with

the defective work may elect to provide a replacement copy in lieu of a

refund. If you received the work electronically, the person or entity

providing it to you may choose to give you a second opportunity to

receive the work electronically in lieu of a refund. If the second copy

is also defective, you may demand a refund in writing without further

opportunities to fix the problem.

1.F.4. Except for the limited right of replacement or refund set forth

in paragraph 1.F.3, this work is provided to you 'AS-IS', WITH NO OTHER

WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED, INCLUDING BUT


NOT LIMITED TO

WARRANTIES OF MERCHANTABILITY OR FITNESS FOR ANY PURPOSE.

1.F.5. Some states do not allow disclaimers of certain implied

warranties or the exclusion or limitation of certain types of damages.

If any disclaimer or limitation set forth in this agreement violates the

law of the state applicable to this agreement, the agreement shall be

interpreted to make the maximum disclaimer or limitation permitted by

the applicable state law. The invalidity or unenforceability of any

provision of this agreement shall not void the remaining provisions.

358
1.F.6. INDEMNITY - You agree to indemnify and hold the Foundation, the

trademark owner, any agent or employee of the Foundation, anyone

providing copies of Project Gutenberg-tm electronic works in accordance

with this agreement, and any volunteers associated with the production,

promotion and distribution of Project Gutenberg-tm electronic works,

harmless from all liability, costs and expenses, including legal fees,

that arise directly or indirectly from any of the following which you do

or cause to occur: (a) distribution of this or any Project Gutenberg-tm

work, (b) alteration, modification, or additions or deletions to any

Project Gutenberg-tm work, and (c) any Defect you cause.

Section 2. Information about the Mission of Project Gutenberg-tm

Project Gutenberg-tm is synonymous with the free distribution of

electronic works in formats readable by the widest variety of computers

including obsolete, old, middle-aged and new computers. It exists

because of the efforts of hundreds of volunteers and donations from

people in all walks of life.

Volunteers and financial support to provide volunteers with the

assistance they need are critical to reaching Project Gutenberg-tm's


359
goals and ensuring that the Project Gutenberg-tm collection will

remain freely available for generations to come. In 2001, the Project

Gutenberg Literary Archive Foundation was created to provide a secure

and permanent future for Project Gutenberg-tm and future generations.

To learn more about the Project Gutenberg Literary Archive Foundation

and how your efforts and donations can help, see Sections 3 and 4

and the Foundation information page at www.gutenberg.org

Section 3. Information about the Project Gutenberg Literary Archive

Foundation

The Project Gutenberg Literary Archive Foundation is a non profit

501(c)(3) educational corporation organized under the laws of the

state of Mississippi and granted tax exempt status by the Internal

Revenue Service. The Foundation's EIN or federal tax identification

number is 64-6221541. Contributions to the Project Gutenberg

Literary Archive Foundation are tax deductible to the full extent

permitted by U.S. federal laws and your state's laws.

The Foundation's principal office is located at 4557 Melan Dr. S.

Fairbanks, AK, 99712., but its volunteers and employees are scattered

throughout numerous locations. Its business office is located at 809


360
North 1500 West, Salt Lake City, UT 84116, (801) 596-1887. Email

contact links and up to date contact information can be found at the

Foundation's web site and official page at www.gutenberg.org/contact

For additional contact information:

Dr. Gregory B. Newby

Chief Executive and Director

gbnewby@pglaf.org

Section 4. Information about Donations to the Project Gutenberg

Literary Archive Foundation

Project Gutenberg-tm depends upon and cannot survive without wide

spread public support and donations to carry out its mission of

increasing the number of public domain and licensed works that can be

freely distributed in machine readable form accessible by the widest

array of equipment including outdated equipment. Many small donations

($1 to $5,000) are particularly important to maintaining tax exempt

status with the IRS.

The Foundation is committed to complying with the laws regulating

charities and charitable donations in all 50 states of the United

States. Compliance requirements are not uniform and it takes a


361
considerable effort, much paperwork and many fees to meet and keep up

with these requirements. We do not solicit donations in locations

where we have not received written confirmation of compliance. To

SEND DONATIONS or determine the status of compliance for any

particular state visit www.gutenberg.org/donate

While we cannot and do not solicit contributions from states where we

have not met the solicitation requirements, we know of no prohibition

against accepting unsolicited donations from donors in such states who

approach us with offers to donate.

International donations are gratefully accepted, but we cannot make

any statements concerning tax treatment of donations received from

outside the United States. U.S. laws alone swamp our small staff.

Please check the Project Gutenberg Web pages for current donation

methods and addresses. Donations are accepted in a number of other

ways including checks, online payments and credit card donations.

To donate, please visit: www.gutenberg.org/donate

Section 5. General Information About Project Gutenberg-tm electronic

works.
362
Professor Michael S. Hart was the originator of the Project Gutenberg-tm

concept of a library of electronic works that could be freely shared

with anyone. For forty years, he produced and distributed Project

Gutenberg-tm eBooks with only a loose network of volunteer support.

Project Gutenberg-tm eBooks are often created from several printed

editions, all of which are confirmed as Public Domain in the U.S.

unless a copyright notice is included. Thus, we do not necessarily

keep eBooks in compliance with any particular paper edition.

Most people start at our Web site which has the main PG search facility:

www.gutenberg.org

This Web site includes information about Project Gutenberg-tm,

including how to make donations to the Project Gutenberg Literary

Archive Foundation, how to help produce our new eBooks, and how to

subscribe to our email newsletter to hear about new eBooks.

\end{PGtext}

% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %

363
% %

% End of the Project Gutenberg EBook of On the Quantum Theory of Line-


Spectra,

% Part 1 and 2, by Niels Henrik David Bohr %

% %

% *** END OF THIS PROJECT GUTENBERG EBOOK QUANTUM THEORY


OF LINE-SPECTRA ***

% %

% ***** This file should be named 47167-t.tex or 47167-t.zip ***** %

% This and all associated files of various formats will be found in: %

% http://www.gutenberg.org/4/7/1/6/47167/ %

% %

% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %

\end{document}

###

@ControlwordReplace = (

['\\eg', 'e.g.'],

['\\ie', 'i.e.']

);

@ControlwordArguments = (

['\\label', 1, 0, '', ''],

364
['\\hyperref', 0, 0, '', '', 1, 1, '', ''],

['\\Graphic', 1, 0, '', '', 1, 0, '<GRAPHIC>', ''],

['\\Figure', 0, 0, '', '', 1, 0, '', '', 1, 0, '<GRAPHIC>', ''],

['\\Figures', 0, 0, '', '', 1, 0, '', '', 1, 0, '', '', 1, 0, '<GRAPHIC>', ''],

['\\FigureH', 0, 0, '', '', 1, 0, '', '', 1, 0, '<GRAPHIC>', ''],

['\\FiguresH', 0, 0, '', '', 1, 0, '', '', 1, 0, '', '', 1, 0, '<GRAPHIC>', ''],

['\\BookMark', 1, 0, '', '', 1, 0, '', ''],

['\\First', 1, 1, '', ''],

['\\Part', 1, 1, 'Part ', ' ', 1, 1, '', ''],

['\\Chapter', 1, 1, '', ''],

['\\Section', 1, 1, '', ' ', 1, 1, '', ''],

['\\SecRef', 0, 0, '', '', 1, 1, '§', ''],

['\\SecNum', 1, 1, '', ''],

['\\Fig', 1, 1, 'Fig. ', ''],

['\\FigNum', 1, 1, '', ''],

['\\Gloss', 0, 0, '', '', 1, 1, '', ''],

['\\Term', 1, 1, '', ', ', 1, 1, '', '', 1, 0, '', ''],

['\\Pagelabel', 1, 0, '', ''],

['\\PageRef', 0, 1, '', ' ', 1, 1, '', ''],

['\\PageRefI', 0, 1, '', ' ', 1, 1, '', ''],

['\\Eq', 0, 0, '', '', 1, 1, '', ''],

['\\Typo', 1, 0, '', '', 1, 1, '', ''],

['\\Note', 1, 0, '', ''],


365
['\\Add', 1, 1, '', ''],

['\\Chg', 1, 0, '', '', 1, 1, '', '']

);

$PageSeparator = qr/^\\PageSep/;

$CustomClean = 'print "\\nCustom cleaning in progress...";

my $cline = 0;

while ($cline <= $#file) {

$file[$cline] =~ s/--------[^\n]*\n//; # strip page separators

$cline++

print "done\\n";';

###

This is pdfTeX, Version 3.1415926-2.5-1.40.14 (TeX Live 2013/Debian)


(format=pdflatex 2014.9.6) 22 NOV 2014 05:22

entering extended mode

%&-line parsing enabled.

**47167-t.tex

(./47167-t.tex

LaTeX2e <2011/06/27>

Babel <3.9h> and hyphenation patterns for 78 languages loaded.

(/usr/share/texlive/texmf-dist/tex/latex/base/book.cls

Document Class: book 2007/10/19 v1.4h Standard LaTeX document class

(/usr/share/texlive/texmf-dist/tex/latex/base/bk12.clo

366
File: bk12.clo 2007/10/19 v1.4h Standard LaTeX file (size option)

\c@part=\count79

\c@chapter=\count80

\c@section=\count81

\c@subsection=\count82

\c@subsubsection=\count83

\c@paragraph=\count84

\c@subparagraph=\count85

\c@figure=\count86

\c@table=\count87

\abovecaptionskip=\skip41

\belowcaptionskip=\skip42

\bibindent=\dimen102

) (/usr/share/texlive/texmf-dist/tex/latex/base/inputenc.sty

Package: inputenc 2008/03/30 v1.1d Input encoding file

\inpenc@prehook=\toks14

\inpenc@posthook=\toks15

(/usr/share/texlive/texmf-dist/tex/latex/base/latin1.def

File: latin1.def 2008/03/30 v1.1d Input encoding file

)) (/usr/share/texlive/texmf-dist/tex/latex/base/ifthen.sty

Package: ifthen 2001/05/26 v1.1c Standard LaTeX ifthen package (DPC)

) (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsmath.sty
367
Package: amsmath 2013/01/14 v2.14 AMS math features

\@mathmargin=\skip43

For additional information on amsmath, use the `?' option.

(/usr/share/texlive/texmf-dist/tex/latex/amsmath/amstext.sty

Package: amstext 2000/06/29 v2.01

(/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsgen.sty

File: amsgen.sty 1999/11/30 v2.0

\@emptytoks=\toks16

\ex@=\dimen103

)) (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsbsy.sty

Package: amsbsy 1999/11/29 v1.2d

\pmbraise@=\dimen104

) (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsopn.sty

Package: amsopn 1999/12/14 v2.01 operator names

\inf@bad=\count88

LaTeX Info: Redefining \frac on input line 210.

\uproot@=\count89

\leftroot@=\count90

LaTeX Info: Redefining \overline on input line 306.

\classnum@=\count91

\DOTSCASE@=\count92

LaTeX Info: Redefining \ldots on input line 378.


368
LaTeX Info: Redefining \dots on input line 381.

LaTeX Info: Redefining \cdots on input line 466.

\Mathstrutbox@=\box26

\strutbox@=\box27

\big@size=\dimen105

LaTeX Font Info: Redeclaring font encoding OML on input line 566.

LaTeX Font Info: Redeclaring font encoding OMS on input line 567.

\macc@depth=\count93

\c@MaxMatrixCols=\count94

\dotsspace@=\muskip10

\c@parentequation=\count95

\dspbrk@lvl=\count96

\tag@help=\toks17

\row@=\count97

\column@=\count98

\maxfields@=\count99

\andhelp@=\toks18

\eqnshift@=\dimen106

\alignsep@=\dimen107

\tagshift@=\dimen108

\tagwidth@=\dimen109

\totwidth@=\dimen110

\lineht@=\dimen111
369
\@envbody=\toks19

\multlinegap=\skip44

\multlinetaggap=\skip45

\mathdisplay@stack=\toks20

LaTeX Info: Redefining \[ on input line 2665.

LaTeX Info: Redefining \] on input line 2666.

) (/usr/share/texlive/texmf-dist/tex/latex/amsfonts/amssymb.sty

Package: amssymb 2013/01/14 v3.01 AMS font symbols

(/usr/share/texlive/texmf-dist/tex/latex/amsfonts/amsfonts.sty

Package: amsfonts 2013/01/14 v3.01 Basic AMSFonts support

\symAMSa=\mathgroup4

\symAMSb=\mathgroup5

LaTeX Font Info: Overwriting math alphabet `\mathfrak' in version `bold'

(Font) U/euf/m/n --> U/euf/b/n on input line 106.

)) (/usr/share/texlive/texmf-dist/tex/latex/base/alltt.sty

Package: alltt 1997/06/16 v2.0g defines alltt environment

) (/usr/share/texlive/texmf-dist/tex/latex/tools/indentfirst.sty

Package: indentfirst 1995/11/23 v1.03 Indent first paragraph (DPC)

) (/usr/share/texlive/texmf-dist/tex/latex/footmisc/footmisc.sty

Package: footmisc 2011/06/06 v5.5b a miscellany of footnote facilities

\FN@temptoken=\toks21

\footnotemargin=\dimen112

\c@pp@next@reset=\count100
370
\c@@fnserial=\count101

Package footmisc Info: Declaring symbol style bringhurst on input line 855.

Package footmisc Info: Declaring symbol style chicago on input line 863.

Package footmisc Info: Declaring symbol style wiley on input line 872.

Package footmisc Info: Declaring symbol style lamport-robust on input line 883.

Package footmisc Info: Declaring symbol style lamport* on input line 903.

Package footmisc Info: Declaring symbol style lamport*-robust on input line 924

) (/usr/share/texlive/texmf-dist/tex/latex/tools/calc.sty

Package: calc 2007/08/22 v4.3 Infix arithmetic (KKT,FJ)

\calc@Acount=\count102

\calc@Bcount=\count103

\calc@Adimen=\dimen113

\calc@Bdimen=\dimen114

\calc@Askip=\skip46

\calc@Bskip=\skip47

LaTeX Info: Redefining \setlength on input line 76.

LaTeX Info: Redefining \addtolength on input line 77.

\calc@Ccount=\count104

\calc@Cskip=\skip48

) (/usr/share/texlive/texmf-dist/tex/latex/fancyhdr/fancyhdr.sty

\fancy@headwidth=\skip49
371
\f@ncyO@elh=\skip50

\f@ncyO@erh=\skip51

\f@ncyO@olh=\skip52

\f@ncyO@orh=\skip53

\f@ncyO@elf=\skip54

\f@ncyO@erf=\skip55

\f@ncyO@olf=\skip56

\f@ncyO@orf=\skip57

) (/usr/share/texlive/texmf-dist/tex/latex/geometry/geometry.sty

Package: geometry 2010/09/12 v5.6 Page Geometry

(/usr/share/texlive/texmf-dist/tex/latex/graphics/keyval.sty

Package: keyval 1999/03/16 v1.13 key=value parser (DPC)

\KV@toks@=\toks22

) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/ifpdf.sty

Package: ifpdf 2011/01/30 v2.3 Provides the ifpdf switch (HO)

Package ifpdf Info: pdfTeX in PDF mode is detected.

) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/ifvtex.sty

Package: ifvtex 2010/03/01 v1.5 Detect VTeX and its facilities (HO)

Package ifvtex Info: VTeX not detected.

) (/usr/share/texlive/texmf-dist/tex/generic/ifxetex/ifxetex.sty

Package: ifxetex 2010/09/12 v0.6 Provides ifxetex conditional

\Gm@cnth=\count105
372
\Gm@cntv=\count106

\c@Gm@tempcnt=\count107

\Gm@bindingoffset=\dimen115

\Gm@wd@mp=\dimen116

\Gm@odd@mp=\dimen117

\Gm@even@mp=\dimen118

\Gm@layoutwidth=\dimen119

\Gm@layoutheight=\dimen120

\Gm@layouthoffset=\dimen121

\Gm@layoutvoffset=\dimen122

\Gm@dimlist=\toks23

) (/usr/share/texlive/texmf-dist/tex/latex/hyperref/hyperref.sty

Package: hyperref 2012/11/06 v6.83m Hypertext links for LaTeX

(/usr/share/texlive/texmf-dist/tex/generic/oberdiek/hobsub-hyperref.sty

Package: hobsub-hyperref 2012/05/28 v1.13 Bundle oberdiek, subset hyperref


(HO)

(/usr/share/texlive/texmf-dist/tex/generic/oberdiek/hobsub-generic.sty

Package: hobsub-generic 2012/05/28 v1.13 Bundle oberdiek, subset generic (HO)

Package: hobsub 2012/05/28 v1.13 Construct package bundles (HO)

Package: infwarerr 2010/04/08 v1.3 Providing info/warning/error messages (HO)

Package: ltxcmds 2011/11/09 v1.22 LaTeX kernel commands for general use (HO)

Package: ifluatex 2010/03/01 v1.3 Provides the ifluatex switch (HO)

373
Package ifluatex Info: LuaTeX not detected.

Package hobsub Info: Skipping package `ifvtex' (already loaded).

Package: intcalc 2007/09/27 v1.1 Expandable calculations with integers (HO)

Package hobsub Info: Skipping package `ifpdf' (already loaded).

Package: etexcmds 2011/02/16 v1.5 Avoid name clashes with e-TeX commands
(HO)

Package etexcmds Info: Could not find \expanded.

(etexcmds) That can mean that you are not using pdfTeX 1.50 or

(etexcmds) that some package has redefined \expanded.

(etexcmds) In the latter case, load this package earlier.

Package: kvsetkeys 2012/04/25 v1.16 Key value parser (HO)

Package: kvdefinekeys 2011/04/07 v1.3 Define keys (HO)

Package: pdftexcmds 2011/11/29 v0.20 Utility functions of pdfTeX for LuaTeX


(HO

Package pdftexcmds Info: LuaTeX not detected.

Package pdftexcmds Info: \pdf@primitive is available.

Package pdftexcmds Info: \pdf@ifprimitive is available.

Package pdftexcmds Info: \pdfdraftmode found.

Package: pdfescape 2011/11/25 v1.13 Implements pdfTeX's escape features (HO)

Package: bigintcalc 2012/04/08 v1.3 Expandable calculations on big integers (HO

Package: bitset 2011/01/30 v1.1 Handle bit-vector datatype (HO)

374
Package: uniquecounter 2011/01/30 v1.2 Provide unlimited unique counter (HO)

Package hobsub Info: Skipping package `hobsub' (already loaded).

Package: letltxmacro 2010/09/02 v1.4 Let assignment for LaTeX macros (HO)

Package: hopatch 2012/05/28 v1.2 Wrapper for package hooks (HO)

Package: xcolor-patch 2011/01/30 xcolor patch

Package: atveryend 2011/06/30 v1.8 Hooks at the very end of document (HO)

Package atveryend Info: \enddocument detected (standard20110627).

Package: atbegshi 2011/10/05 v1.16 At begin shipout hook (HO)

Package: refcount 2011/10/16 v3.4 Data extraction from label references (HO)

Package: hycolor 2011/01/30 v1.7 Color options for hyperref/bookmark (HO)

) (/usr/share/texlive/texmf-dist/tex/latex/oberdiek/auxhook.sty

Package: auxhook 2011/03/04 v1.3 Hooks for auxiliary files (HO)

) (/usr/share/texlive/texmf-dist/tex/latex/oberdiek/kvoptions.sty

Package: kvoptions 2011/06/30 v3.11 Key value format for package options (HO)

\@linkdim=\dimen123

\Hy@linkcounter=\count108

\Hy@pagecounter=\count109

(/usr/share/texlive/texmf-dist/tex/latex/hyperref/pd1enc.def

File: pd1enc.def 2012/11/06 v6.83m Hyperref: PDFDocEncoding definition (HO)

\Hy@SavedSpaceFactor=\count110
375
(/usr/share/texlive/texmf-dist/tex/latex/latexconfig/hyperref.cfg

File: hyperref.cfg 2002/06/06 v1.2 hyperref configuration of TeXLive

Package hyperref Info: Option `hyperfootnotes' set `false' on input line 4319.

Package hyperref Info: Option `bookmarks' set `true' on input line 4319.

Package hyperref Info: Option `linktocpage' set `false' on input line 4319.

Package hyperref Info: Option `pdfdisplaydoctitle' set `true' on input line 431

9.

Package hyperref Info: Option `pdfpagelabels' set `true' on input line 4319.

Package hyperref Info: Option `bookmarksopen' set `true' on input line 4319.

Package hyperref Info: Option `colorlinks' set `true' on input line 4319.

Package hyperref Info: Hyper figures OFF on input line 4443.

Package hyperref Info: Link nesting OFF on input line 4448.

Package hyperref Info: Hyper index ON on input line 4451.

Package hyperref Info: Plain pages OFF on input line 4458.

Package hyperref Info: Backreferencing OFF on input line 4463.

Package hyperref Info: Implicit mode ON; LaTeX internals redefined.

Package hyperref Info: Bookmarks ON on input line 4688.

\c@Hy@tempcnt=\count111

(/usr/share/texlive/texmf-dist/tex/latex/url/url.sty

\Urlmuskip=\muskip11

Package: url 2013/09/16 ver 3.4 Verb mode for urls, etc.

)
376
LaTeX Info: Redefining \url on input line 5041.

\XeTeXLinkMargin=\dimen124

\Fld@menulength=\count112

\Field@Width=\dimen125

\Fld@charsize=\dimen126

Package hyperref Info: Hyper figures OFF on input line 6295.

Package hyperref Info: Link nesting OFF on input line 6300.

Package hyperref Info: Hyper index ON on input line 6303.

Package hyperref Info: backreferencing OFF on input line 6310.

Package hyperref Info: Link coloring ON on input line 6313.

Package hyperref Info: Link coloring with OCG OFF on input line 6320.

Package hyperref Info: PDF/A mode OFF on input line 6325.

LaTeX Info: Redefining \ref on input line 6365.

LaTeX Info: Redefining \pageref on input line 6369.

\Hy@abspage=\count113

\c@Item=\count114

Package hyperref Message: Driver: hpdftex.

(/usr/share/texlive/texmf-dist/tex/latex/hyperref/hpdftex.def

File: hpdftex.def 2012/11/06 v6.83m Hyperref driver for pdfTeX

\Fld@listcount=\count115
377
\c@bookmark@seq@number=\count116

(/usr/share/texlive/texmf-dist/tex/latex/oberdiek/rerunfilecheck.sty

Package: rerunfilecheck 2011/04/15 v1.7 Rerun checks for auxiliary files (HO)

Package uniquecounter Info: New unique counter `rerunfilecheck' on input line 2

82.

\Hy@SectionHShift=\skip58

\c@PartNo=\count117

\c@SecNo=\count118

(./47167-t.aux)

\openout1 = `47167-t.aux'.

LaTeX Font Info: Checking defaults for OML/cmm/m/it on input line 410.

LaTeX Font Info: ... okay on input line 410.

LaTeX Font Info: Checking defaults for T1/cmr/m/n on input line 410.

LaTeX Font Info: ... okay on input line 410.

LaTeX Font Info: Checking defaults for OT1/cmr/m/n on input line 410.

LaTeX Font Info: ... okay on input line 410.

LaTeX Font Info: Checking defaults for OMS/cmsy/m/n on input line 410.

LaTeX Font Info: ... okay on input line 410.

LaTeX Font Info: Checking defaults for OMX/cmex/m/n on input line 410.

LaTeX Font Info: ... okay on input line 410.


378
LaTeX Font Info: Checking defaults for U/cmr/m/n on input line 410.

LaTeX Font Info: ... okay on input line 410.

LaTeX Font Info: Checking defaults for PD1/pdf/m/n on input line 410.

LaTeX Font Info: ... okay on input line 410.

*geometry* driver: auto-detecting

*geometry* detected driver: pdftex

*geometry* verbose mode - [ preamble ] result:

* driver: pdftex

* paper: <default>

* layout: <same size as paper>

* layoutoffset:(h,v)=(0.0pt,0.0pt)

* hratio: 1:1

* modes: includehead includefoot twoside

* h-part:(L,W,R)=(9.03375pt, 307.14749pt, 9.03375pt)

* v-part:(T,H,B)=(1.26749pt, 466.58623pt, 1.90128pt)

* \paperwidth=325.215pt

* \paperheight=469.75499pt

* \textwidth=307.14749pt

* \textheight=404.71243pt

* \oddsidemargin=-63.23624pt

* \evensidemargin=-63.23624pt

* \topmargin=-71.0025pt

* \headheight=12.0pt
379
* \headsep=19.8738pt

* \topskip=12.0pt

* \footskip=30.0pt

* \marginparwidth=98.0pt

* \marginparsep=7.0pt

* \columnsep=10.0pt

* \skip\footins=10.8pt plus 4.0pt minus 2.0pt

* \hoffset=0.0pt

* \voffset=0.0pt

* \mag=1000

* \@twocolumnfalse

* \@twosidetrue

* \@mparswitchtrue

* \@reversemarginfalse

* (1in=72.27pt=25.4mm, 1cm=28.453pt)

\AtBeginShipoutBox=\box28

(/usr/share/texlive/texmf-dist/tex/latex/graphics/color.sty

Package: color 2005/11/14 v1.0j Standard LaTeX Color (DPC)

(/usr/share/texlive/texmf-dist/tex/latex/latexconfig/color.cfg

File: color.cfg 2007/01/18 v1.5 color configuration of teTeX/TeXLive

Package color Info: Driver file: pdftex.def on input line 130.


380
(/usr/share/texlive/texmf-dist/tex/latex/pdftex-def/pdftex.def

File: pdftex.def 2011/05/27 v0.06d Graphics/color for pdfTeX

\Gread@gobject=\count119

(/usr/share/texlive/texmf-dist/tex/context/base/supp-pdf.mkii

[Loading MPS to PDF converter (version 2006.09.02).]

\scratchcounter=\count120

\scratchdimen=\dimen127

\scratchbox=\box29

\nofMPsegments=\count121

\nofMParguments=\count122

\everyMPshowfont=\toks24

\MPscratchCnt=\count123

\MPscratchDim=\dimen128

\MPnumerator=\count124

\makeMPintoPDFobject=\count125

\everyMPtoPDFconversion=\toks25

)))

Package hyperref Info: Link coloring ON on input line 410.

(/usr/share/texlive/texmf-dist/tex/latex/hyperref/nameref.sty

Package: nameref 2012/10/27 v2.43 Cross-referencing by name of section

(/usr/share/texlive/texmf-dist/tex/generic/oberdiek/gettitlestring.sty

Package: gettitlestring 2010/12/03 v1.4 Cleanup title references (HO)

)
381
\c@section@level=\count126

LaTeX Info: Redefining \ref on input line 410.

LaTeX Info: Redefining \pageref on input line 410.

LaTeX Info: Redefining \nameref on input line 410.

(./47167-t.out) (./47167-t.out)

\@outlinefile=\write3

\openout3 = `47167-t.out'.

Overfull \hbox (160.35922pt too wide) in paragraph at lines 417--417

[]\OT1/cmtt/m/n/8 The Project Gutenberg EBook of On the Quantum Theory of


Line-

Spectra, Part 1 and 2, by Niels Henrik David Bohr[]

[]

Overfull \hbox (7.35703pt too wide) in paragraph at lines 419--419

[]\OT1/cmtt/m/n/8 This eBook is for the use of anyone anywhere in the United St

ates and most[]

[]

382
Overfull \hbox (7.35703pt too wide) in paragraph at lines 421--421

[]\OT1/cmtt/m/n/8 whatsoever. You may copy it, give it away or re-use it under

the terms of[]

[]

Overfull \hbox (15.85715pt too wide) in paragraph at lines 423--423

[]\OT1/cmtt/m/n/8 www.gutenberg.org. If you are not located in the United Stat

es, you'll have[]

[]

Overfull \hbox (28.60733pt too wide) in paragraph at lines 424--424

[]\OT1/cmtt/m/n/8 to check the laws of the country where you are located before

using this ebook.[]

[]

Overfull \hbox (15.85715pt too wide) in paragraph at lines 438--438

[]\OT1/cmtt/m/n/8 *** START OF THIS PROJECT GUTENBERG EBOOK


QUANTUM THEORY OF L

INE-SPECTRA ***[]

[]

383
LaTeX Font Info: Try loading font information for U+msa on input line 440.

(/usr/share/texlive/texmf-dist/tex/latex/amsfonts/umsa.fd

File: umsa.fd 2013/01/14 v3.01 AMS symbols A

LaTeX Font Info: Try loading font information for U+msb on input line 440.

(/usr/share/texlive/texmf-dist/tex/latex/amsfonts/umsb.fd

File: umsb.fd 2013/01/14 v3.01 AMS symbols B

) [1

{/var/lib/texmf/fonts/map/pdftex/updmap/pdftex.map}] [2] [1

LaTeX Font Info: Try loading font information for OMS+cmr on input line 537.

(/usr/share/texlive/texmf-dist/tex/latex/base/omscmr.fd

File: omscmr.fd 1999/05/25 v2.5h Standard LaTeX font definitions

LaTeX Font Info: Font shape `OMS/cmr/m/it' in size <10> not available

(Font) Font shape `OMS/cmsy/m/n' tried instead on input line 537.

[2] [1

384
] [2] [3] [4]

LaTeX Font Info: Font shape `OMS/cmr/bx/n' in size <12> not available

(Font) Font shape `OMS/cmsy/b/n' tried instead on input line 635.

[5

] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [

22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36]

LaTeX Font Info: Font shape `OMS/cmr/m/n' in size <12> not available

(Font) Font shape `OMS/cmsy/m/n' tried instead on input line 1470.

[37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51]

LaTeX Font Info: Font shape `OMS/cmr/m/n' in size <10> not available

(Font) Font shape `OMS/cmsy/m/n' tried instead on input line 1913.

Underfull \hbox (badness 2376) in paragraph at lines 1913--1913

[]\OT1/cmr/m/n/10 In this con-nec-tion it may be men-tioned that the method of

[]

[52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67]

[68] [69]pdfTeX warning (ext4): destination with the same identifier (name{SecN

o.1}) has been already used, duplicate ignored


385
<to be read again>

\relax

l.2290 ...ory of the series~spectrum~of~hydrogen.}

[70

] [71] [72] [73] [74] [75] [76] [77] [78]pdfTeX warning (ext4): destination wit

h the same identifier (name{SecNo.2}) has been already used, duplicate ignored

<to be read again>

\relax

l.2526 ... states of a perturbed~periodic~system.}

[79] [80]

LaTeX Font Info: Try loading font information for U+euf on input line 2576.

(/usr/share/texlive/texmf-dist/tex/latex/amsfonts/ueuf.fd

File: ueuf.fd 2013/01/14 v3.01 Euler Fraktur

) [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] [9

6] [97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [

110] [111] [112] [113] [114] [115] [116] [117] [118] [119] [120] [121] [122]pdf

TeX warning (ext4): destination with the same identifier (name{SecNo.3}) has be

en already used, duplicate ignored

<to be read again>

\relax

l.3600 ...e fine structure of the~hydrogen~lines.}

[123] [124] [125] [126] [127


386
] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] [139] [140]

[141] [142] [143] [144] [145] [146] [147] [148] [149] [150] [151] [152]

Overfull \vbox (0.32088pt too high) has occurred while \output is active []

[153] [154] [155] [156] [157] [158] [159] [160] [161] [162]

Underfull \hbox (badness 1038) in paragraph at lines 4582--4601

\OT1/cmr/m/n/12 mula [][]$(80)$[][] is in agree-ment with the for-mul^^Z ob-tai

ned

[]

[163] [164] [165] [166] [167] [168] [169] [170] [171] [172] [173] [174] [175] [

176] [177] [178] [179] [180] [181] [182] [183] [184] [185] [186] [187] [188] [1

89] [190] [191] [192] [193] [194] [195] [196] [197] [198] [199]

Overfull \hbox (15.85715pt too wide) in paragraph at lines 5495--5495

[]\OT1/cmtt/m/n/8 End of the Project Gutenberg EBook of On the Quantum


Theory o

f Line-Spectra,[]

[]

Overfull \hbox (7.35703pt too wide) in paragraph at lines 5498--5498

[]\OT1/cmtt/m/n/8 *** END OF THIS PROJECT GUTENBERG EBOOK


QUANTUM THEORY OF LIN

387
E-SPECTRA ***[]

[]

[1

Overfull \hbox (3.10696pt too wide) in paragraph at lines 5565--5565

[]\OT1/cmtt/m/n/8 1.C. The Project Gutenberg Literary Archive Foundation ("the

Foundation"[]

[]

Overfull \hbox (3.10696pt too wide) in paragraph at lines 5570--5570

[]\OT1/cmtt/m/n/8 located in the United States, we do not claim a right to prev

ent you from[]

[]

Overfull \hbox (3.10696pt too wide) in paragraph at lines 5575--5575

[]\OT1/cmtt/m/n/8 freely sharing Project Gutenberg-tm works in compliance with

the terms of[]


388
[]

[2] [3]

Overfull \hbox (3.10696pt too wide) in paragraph at lines 5638--5638

[]\OT1/cmtt/m/n/8 posted on the official Project Gutenberg-tm web site (www.gut

enberg.org),[]

[]

[4] [5] [6] [7] [8]

Package atveryend Info: Empty hook `BeforeClearDocument' on input line 5867.

[9]

Package atveryend Info: Empty hook `AfterLastShipout' on input line 5867.

(./47167-t.aux)

Package atveryend Info: Executing hook `AtVeryEndDocument' on input line


5867.

*File List*

book.cls 2007/10/19 v1.4h Standard LaTeX document class

bk12.clo 2007/10/19 v1.4h Standard LaTeX file (size option)

inputenc.sty 2008/03/30 v1.1d Input encoding file

latin1.def 2008/03/30 v1.1d Input encoding file

ifthen.sty 2001/05/26 v1.1c Standard LaTeX ifthen package (DPC)

amsmath.sty 2013/01/14 v2.14 AMS math features

389
amstext.sty 2000/06/29 v2.01

amsgen.sty 1999/11/30 v2.0

amsbsy.sty 1999/11/29 v1.2d

amsopn.sty 1999/12/14 v2.01 operator names

amssymb.sty 2013/01/14 v3.01 AMS font symbols

amsfonts.sty 2013/01/14 v3.01 Basic AMSFonts support

alltt.sty 1997/06/16 v2.0g defines alltt environment

indentfirst.sty 1995/11/23 v1.03 Indent first paragraph (DPC)

footmisc.sty 2011/06/06 v5.5b a miscellany of footnote facilities

calc.sty 2007/08/22 v4.3 Infix arithmetic (KKT,FJ)

fancyhdr.sty

geometry.sty 2010/09/12 v5.6 Page Geometry

keyval.sty 1999/03/16 v1.13 key=value parser (DPC)

ifpdf.sty 2011/01/30 v2.3 Provides the ifpdf switch (HO)

ifvtex.sty 2010/03/01 v1.5 Detect VTeX and its facilities (HO)

ifxetex.sty 2010/09/12 v0.6 Provides ifxetex conditional

hyperref.sty 2012/11/06 v6.83m Hypertext links for LaTeX

hobsub-hyperref.sty 2012/05/28 v1.13 Bundle oberdiek, subset hyperref (HO)

hobsub-generic.sty 2012/05/28 v1.13 Bundle oberdiek, subset generic (HO)

hobsub.sty 2012/05/28 v1.13 Construct package bundles (HO)

infwarerr.sty 2010/04/08 v1.3 Providing info/warning/error messages (HO)

ltxcmds.sty 2011/11/09 v1.22 LaTeX kernel commands for general use (HO)

ifluatex.sty 2010/03/01 v1.3 Provides the ifluatex switch (HO)


390
intcalc.sty 2007/09/27 v1.1 Expandable calculations with integers (HO)

etexcmds.sty 2011/02/16 v1.5 Avoid name clashes with e-TeX commands (HO)

kvsetkeys.sty 2012/04/25 v1.16 Key value parser (HO)

kvdefinekeys.sty 2011/04/07 v1.3 Define keys (HO)

pdftexcmds.sty 2011/11/29 v0.20 Utility functions of pdfTeX for LuaTeX (HO)

pdfescape.sty 2011/11/25 v1.13 Implements pdfTeX's escape features (HO)

bigintcalc.sty 2012/04/08 v1.3 Expandable calculations on big integers (HO)

bitset.sty 2011/01/30 v1.1 Handle bit-vector datatype (HO)

uniquecounter.sty 2011/01/30 v1.2 Provide unlimited unique counter (HO)

letltxmacro.sty 2010/09/02 v1.4 Let assignment for LaTeX macros (HO)

hopatch.sty 2012/05/28 v1.2 Wrapper for package hooks (HO)

xcolor-patch.sty 2011/01/30 xcolor patch

atveryend.sty 2011/06/30 v1.8 Hooks at the very end of document (HO)

atbegshi.sty 2011/10/05 v1.16 At begin shipout hook (HO)

refcount.sty 2011/10/16 v3.4 Data extraction from label references (HO)

hycolor.sty 2011/01/30 v1.7 Color options for hyperref/bookmark (HO)

auxhook.sty 2011/03/04 v1.3 Hooks for auxiliary files (HO)

kvoptions.sty 2011/06/30 v3.11 Key value format for package options (HO)

pd1enc.def 2012/11/06 v6.83m Hyperref: PDFDocEncoding definition (HO)

hyperref.cfg 2002/06/06 v1.2 hyperref configuration of TeXLive

url.sty 2013/09/16 ver 3.4 Verb mode for urls, etc.

hpdftex.def 2012/11/06 v6.83m Hyperref driver for pdfTeX

rerunfilecheck.sty 2011/04/15 v1.7 Rerun checks for auxiliary files (HO)


391
color.sty 2005/11/14 v1.0j Standard LaTeX Color (DPC)

color.cfg 2007/01/18 v1.5 color configuration of teTeX/TeXLive

pdftex.def 2011/05/27 v0.06d Graphics/color for pdfTeX

supp-pdf.mkii

nameref.sty 2012/10/27 v2.43 Cross-referencing by name of section

gettitlestring.sty 2010/12/03 v1.4 Cleanup title references (HO)

47167-t.out

47167-t.out

umsa.fd 2013/01/14 v3.01 AMS symbols A

umsb.fd 2013/01/14 v3.01 AMS symbols B

omscmr.fd 1999/05/25 v2.5h Standard LaTeX font definitions

ueuf.fd 2013/01/14 v3.01 Euler Fraktur

***********

Package atveryend Info: Executing hook `AtEndAfterFileList' on input line 5867.

Package rerunfilecheck Info: File `47167-t.out' has not changed.

(rerunfilecheck) Checksum:
14EA603261988BEA2BBCAC95B4BE4629;1097.

Package atveryend Info: Empty hook `AtVeryVeryEnd' on input line 5867.

Here is how much of TeX's memory you used:

7430 strings out of 493304

392
102167 string characters out of 6139872

223006 words of memory out of 5000000

10300 multiletter control sequences out of 15000+600000

18356 words of font info for 71 fonts, out of 8000000 for 9000

957 hyphenation exceptions out of 8191

28i,20n,43p,304b,483s stack positions out of 5000i,500n,10000p,200000b,80000s

</usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmbsy10.pfb></u

sr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmbx12.pfb></usr/sha

re/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmcsc10.pfb></usr/share/te

xlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmex10.pfb></usr/share/texlive/

texmf-dist/fonts/type1/public/amsfonts/cm/cmmi10.pfb></usr/share/texlive/texmf-

dist/fonts/type1/public/amsfonts/cm/cmmi12.pfb></usr/share/texlive/texmf-dist/f

onts/type1/public/amsfonts/cm/cmmi5.pfb></usr/share/texlive/texmf-dist/fonts/ty

pe1/public/amsfonts/cm/cmmi6.pfb></usr/share/texlive/texmf-dist/fonts/type1/pub

lic/amsfonts/cm/cmmi7.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/ams

fonts/cm/cmmi8.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/c

m/cmr10.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmr12

.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmr17.pfb></

usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmr5.pfb></usr/shar

e/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmr6.pfb></usr/share/texliv

e/texmf-dist/fonts/type1/public/amsfonts/cm/cmr7.pfb></usr/share/texlive/texmf-

dist/fonts/type1/public/amsfonts/cm/cmr8.pfb></usr/share/texlive/texmf-dist/fon

ts/type1/public/amsfonts/cm/cmsy10.pfb></usr/share/texlive/texmf-dist/fonts/typ
393
e1/public/amsfonts/cm/cmsy6.pfb></usr/share/texlive/texmf-dist/fonts/type1/publ

ic/amsfonts/cm/cmsy7.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsf

onts/cm/cmsy8.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm

/cmti10.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmti1

2.pfb></usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmti8.pfb><

/usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmtt10.pfb></usr/s

hare/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmtt8.pfb></usr/share/te

xlive/texmf-dist/fonts/type1/public/amsfonts/euler/eufm10.pfb>

Output written on 47167-t.pdf (212 pages, 766804 bytes).

PDF statistics:

2006 PDF objects out of 2073 (max. 8388607)

1747 compressed objects within 18 object streams

509 named destinations out of 1000 (max. 500000)

145 words of extra memory for PDF output out of 10000 (max. 10000000)

394

You might also like