Download PDF
ads:
(ebook – PDF Science)
IMPORTANT NOTE (please paste this note on the top of all ebooks):
Please use the following classifications in the names of ebooks that
you input or currently have on your hard drive:
1.) Ebooks are Adobe Acrobat PDF files or Tomeraider type only, do not
name txt files ebooks, no matter how long your lines are.
2.) In parenthesis within the name, type (ebook – Type - Class)
3.) In the “Type” space put the format, either PDF or TOMERAIDER. In
the “Class” space, put the classification of the book, classifications
are: Biography, Children, Fiction, Food, Games, Government, Health,
Internet, Martial-Arts, Mathematics, Other, Programming, Reference,
Religious, Science, Sci-Fi, Sex, and Software.
This new standard is accepted by Fink Crew (who do most of the
conversions from plain text to PDF) and MacWarez (One of the largest
groups currently scanning ebooks). It helps people searching for these
files to find them more easily.
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
Relativity: The Special and General Theory
Albert Einstein
ads:
Albert Einstein
Relativity
The Special and General Theory
Written: 1916 (this revised edition: 1924)
Source: Relativity: The Special and General Theory © 1920
Publisher: Methuen & Co Ltd
First Published: December, 1916
Translated: Robert W. Lawson (Authorised translation)
Transcription/Markup: Brian Basgen
Convertion to PDF: Sjoerd Langkemper
Offline Version: Einstein Reference Archive (marxists.org) 1999
Preface
Part I: The Special Theory of Relativity
01. Physical Meaning of Geometrical Propositions
02. The System of Co−ordinates
03. Space and Time in Classical Mechanics
04. The Galileian System of Co−ordinates
05. The Principle of Relativity (in the Restricted Sense)
06. The Theorem of the Addition of Velocities employed in Classical Mechanics
07. The Apparent Incompatability of the Law of Propagation of Light with the Principle of Relativity
08. On the Idea of Time in Physics
09. The Relativity of Simultaneity
10. On the Relativity of the Conception of Distance
11. The Lorentz Transformation
12. The Behaviour of Measuring−Rods and Clocks in Motion
13. Theorem of the Addition of Velocities. The Experiment of Fizeau
14. The Hueristic Value of the Theory of Relativity
15. General Results of the Theory
16. Expereince and the Special Theory of Relativity
17. Minkowski's Four−dimensial Space
Part II: The General Theory of Relativity
18. Special and General Principle of Relativity
19. The Gravitational Field
20. The Equality of Inertial and Gravitational Mass as an Argument for the General Postulate of
Relativity
21. In What Respects are the Foundations of Classical Mechanics and of the Special Theory of
Relativity Unsatisfactory?
Relativity: The Special and General Theory
1
22. A Few Inferences from the General Principle of Relativity
23. Behaviour of Clocks and Measuring−Rods on a Rotating Body of Reference
24. Euclidean and non−Euclidean Continuum
25. Gaussian Co−ordinates
26. The Space−Time Continuum of the Speical Theory of Relativity Considered as a Euclidean
Continuum
27. The Space−Time Continuum of the General Theory of Realtiivty is Not a Eculidean Continuum
28. Exact Formulation of the General Principle of Relativity
29. The Solution of the Problem of Gravitation on the Basis of the General Principle of Relativity
Part III: Considerations on the Universe as a Whole
30. Cosmological Difficulties of Netwon's Theory
31. The Possibility of a "Finite" and yet "Unbounded" Universe
32. The Structure of Space According to the General Theory of Relativity
Appendices:
01. Simple Derivation of the Lorentz Transformation (sup. ch. 11)
02. Minkowski's Four−Dimensional Space ("World") (sup. ch 17)
03. The Experimental Confirmation of the General Theory of Relativity
04. The Structure of Space According to the General Theory of Relativity (sup. ch 32)
05. Relativity and the Problem of Space
Note: The fifth appendix was added by Einstein at the time of the fifteenth re−printing of this book;
and as a result is still under copyright restrictions so cannot be added without the permission of the
publisher.
Einstein Reference Archive
Relativity: The Special and General Theory
2
Albert Einstein
Relativity: The Special and General Theory
Preface
(December, 1916)
The present book is intended, as far as possible, to give an exact insight into the theory of
Relativity to those readers who, from a general scientific and philosophical point of view, are
interested in the theory, but who are not conversant with the mathematical apparatus of theoretical
physics. The work presumes a standard of education corresponding to that of a university
matriculation examination, and, despite the shortness of the book, a fair amount of patience and
force of will on the part of the reader. The author has spared himself no pains in his endeavour to
present the main ideas in the simplest and most intelligible form, and on the whole, in the sequence
and connection in which they actually originated. In the interest of clearness, it appeared to me
inevitable that I should repeat myself frequently, without paying the slightest attention to the
elegance of the presentation. I adhered scrupulously to the precept of that brilliant theoretical
physicist L. Boltzmann, according to whom matters of elegance ought to be left to the tailor and to
the cobbler. I make no pretence of having withheld from the reader difficulties which are inherent to
the subject. On the other hand, I have purposely treated the empirical physical foundations of the
theory in a "step−motherly" fashion, so that readers unfamiliar with physics may not feel like the
wanderer who was unable to see the forest for the trees. May the book bring some one a few
happy hours of suggestive thought!
December, 1916
A. EINSTEIN
Next: The Physical Meaning of Geometrical Propositions
Relativity: The Special and General Theory
Relativity: The Special and General Theory
3
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
Part I
The Special Theory of Relativity
Physical Meaning of Geometrical Propositions
In your schooldays most of you who read this book made acquaintance with the noble building of
Euclid's geometry, and you remember — perhaps with more respect than love — the magnificent
structure, on the lofty staircase of which you were chased about for uncounted hours by
conscientious teachers. By reason of our past experience, you would certainly regard everyone
with disdain who should pronounce even the most out−of−the−way proposition of this science to be
untrue. But perhaps this feeling of proud certainty would leave you immediately if some one were to
ask you: "What, then, do you mean by the assertion that these propositions are true?" Let us
proceed to give this question a little consideration.
Geometry sets out form certain conceptions such as "plane," "point," and "straight line," with which
we are able to associate more or less definite ideas, and from certain simple propositions (axioms)
which, in virtue of these ideas, we are inclined to accept as "true." Then, on the basis of a logical
process, the justification of which we feel ourselves compelled to admit, all remaining propositions
are shown to follow from those axioms, i.e. they are proven. A proposition is then correct ("true")
when it has been derived in the recognised manner from the axioms. The question of "truth" of the
individual geometrical propositions is thus reduced to one of the "truth" of the axioms. Now it has
long been known that the last question is not only unanswerable by the methods of geometry, but
that it is in itself entirely without meaning. We cannot ask whether it is true that only one straight
line goes through two points. We can only say that Euclidean geometry deals with things called
"straight lines," to each of which is ascribed the property of being uniquely determined by two
points situated on it. The concept "true" does not tally with the assertions of pure geometry,
because by the word "true" we are eventually in the habit of designating always the
correspondence with a "real" object; geometry, however, is not concerned with the relation of the
ideas involved in it to objects of experience, but only with the logical connection of these ideas
among themselves.
It is not difficult to understand why, in spite of this, we feel constrained to call the propositions of
geometry "true." Geometrical ideas correspond to more or less exact objects in nature, and these
last are undoubtedly the exclusive cause of the genesis of those ideas. Geometry ought to refrain
from such a course, in order to give to its structure the largest possible logical unity. The practice,
for example, of seeing in a "distance" two marked positions on a practically rigid body is something
which is lodged deeply in our habit of thought. We are accustomed further to regard three points as
being situated on a straight line, if their apparent positions can be made to coincide for observation
with one eye, under suitable choice of our place of observation.
If, in pursuance of our habit of thought, we now supplement the propositions of Euclidean geometry
by the single proposition that two points on a practically rigid body always correspond to the same
distance (line−interval), independently of any changes in position to which we may subject the
body, the propositions of Euclidean geometry then resolve themselves into propositions on the
Relativity: The Special and General Theory
4
possible relative position of practically rigid bodies.
1)
Geometry which has been supplemented in
this way is then to be treated as a branch of physics. We can now legitimately ask as to the "truth"
of geometrical propositions interpreted in this way, since we are justified in asking whether these
propositions are satisfied for those real things we have associated with the geometrical ideas. In
less exact terms we can express this by saying that by the "truth" of a geometrical proposition in
this sense we understand its validity for a construction with rule and compasses.
Of course the conviction of the "truth" of geometrical propositions in this sense is founded
exclusively on rather incomplete experience. For the present we shall assume the "truth" of the
geometrical propositions, then at a later stage (in the general theory of relativity) we shall see that
this "truth" is limited, and we shall consider the extent of its limitation.
Next: The System of Co−ordinates
Notes
1)
It follows that a natural object is associated also with a straight line. Three points A, B and C on a
rigid body thus lie in a straight line when the points A and C being given, B is chosen such that the
sum of the distances AB and BC is as short as possible. This incomplete suggestion will suffice for
the present purpose.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
5
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The System of Co−ordinates
On the basis of the physical interpretation of distance which has been indicated, we are also in a
position to establish the distance between two points on a rigid body by means of measurements.
For this purpose we require a " distance " (rod S) which is to be used once and for all, and which
we employ as a standard measure. If, now, A and B are two points on a rigid body, we can
construct the line joining them according to the rules of geometry ; then, starting from A, we can
mark off the distance S time after time until we reach B. The number of these operations required is
the numerical measure of the distance AB. This is the basis of all measurement of length.
1)
Every description of the scene of an event or of the position of an object in space is based on the
specification of the point on a rigid body (body of reference) with which that event or object
coincides. This applies not only to scientific description, but also to everyday life. If I analyse the
place specification " Times Square, New York,"
[A]
I arrive at the following result. The earth is the
rigid body to which the specification of place refers; " Times Square, New York," is a well−defined
point, to which a name has been assigned, and with which the event coincides in space.
2)
This primitive method of place specification deals only with places on the surface of rigid bodies,
and is dependent on the existence of points on this surface which are distinguishable from each
other. But we can free ourselves from both of these limitations without altering the nature of our
specification of position. If, for instance, a cloud is hovering over Times Square, then we can
determine its position relative to the surface of the earth by erecting a pole perpendicularly on the
Square, so that it reaches the cloud. The length of the pole measured with the standard
measuring−rod, combined with the specification of the position of the foot of the pole, supplies us
with a complete place specification. On the basis of this illustration, we are able to see the manner
in which a refinement of the conception of position has been developed.
(a) We imagine the rigid body, to which the place specification is referred, supplemented in such a
manner that the object whose position we require is reached by. the completed rigid body.
(b) In locating the position of the object, we make use of a number (here the length of the pole
measured with the measuring−rod) instead of designated points of reference.
(c) We speak of the height of the cloud even when the pole which reaches the cloud has not been
erected. By means of optical observations of the cloud from different positions on the ground, and
taking into account the properties of the propagation of light, we determine the length of the pole
we should have required in order to reach the cloud.
From this consideration we see that it will be advantageous if, in the description of position, it
should be possible by means of numerical measures to make ourselves independent of the
existence of marked positions (possessing names) on the rigid body of reference. In the physics of
measurement this is attained by the application of the Cartesian system of co−ordinates.
This consists of three plane surfaces perpendicular to each other and rigidly attached to a rigid
body. Referred to a system of co−ordinates, the scene of any event will be determined (for the main
part) by the specification of the lengths of the three perpendiculars or co−ordinates (x, y, z) which
can be dropped from the scene of the event to those three plane surfaces. The lengths of these
Relativity: The Special and General Theory
6
three perpendiculars can be determined by a series of manipulations with rigid measuring−rods
performed according to the rules and methods laid down by Euclidean geometry.
In practice, the rigid surfaces which constitute the system of co−ordinates are generally not
available ; furthermore, the magnitudes of the co−ordinates are not actually determined by
constructions with rigid rods, but by indirect means. If the results of physics and astronomy are to
maintain their clearness, the physical meaning of specifications of position must always be sought
in accordance with the above considerations.
3)
We thus obtain the following result: Every description of events in space involves the use of a rigid
body to which such events have to be referred. The resulting relationship takes for granted that the
laws of Euclidean geometry hold for "distances;" the "distance" being represented physically by
means of the convention of two marks on a rigid body.
Next: Space and Time in Classical Mechanics
Notes
1)
Here we have assumed that there is nothing left over i.e. that the measurement gives a whole
number. This difficulty is got over by the use of divided measuring−rods, the introduction of which
does not demand any fundamentally new method.
[A]
Einstein used "Potsdamer Platz, Berlin" in the original text. In the authorised translation this was
supplemented with "Tranfalgar Square, London". We have changed this to "Times Square, New
York", as this is the most well known/identifiable location to English speakers in the present day.
[Note by the janitor.]
2)
It is not necessary here to investigate further the significance of the expression "coincidence in
space." This conception is sufficiently obvious to ensure that differences of opinion are scarcely
likely to arise as to its applicability in practice.
3)
A refinement and modification of these views does not become necessary until we come to deal
with the general theory of relativity, treated in the second part of this book.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
7
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
Space and Time in Classical Mechanics
The purpose of mechanics is to describe how bodies change their position in space with "time." I
should load my conscience with grave sins against the sacred spirit of lucidity were I to formulate
the aims of mechanics in this way, without serious reflection and detailed explanations. Let us
proceed to disclose these sins.
It is not clear what is to be understood here by "position" and "space." I stand at the window of a
railway carriage which is travelling uniformly, and drop a stone on the embankment, without
throwing it. Then, disregarding the influence of the air resistance, I see the stone descend in a
straight line. A pedestrian who observes the misdeed from the footpath notices that the stone falls
to earth in a parabolic curve. I now ask: Do the "positions" traversed by the stone lie "in reality" on a
straight line or on a parabola? Moreover, what is meant here by motion "in space" ? From the
considerations of the previous section the answer is self−evident. In the first place we entirely shun
the vague word "space," of which, we must honestly acknowledge, we cannot form the slightest
conception, and we replace it by "motion relative to a practically rigid body of reference." The
positions relative to the body of reference (railway carriage or embankment) have already been
defined in detail in the preceding section. If instead of " body of reference " we insert " system of
co−ordinates," which is a useful idea for mathematical description, we are in a position to say : The
stone traverses a straight line relative to a system of co−ordinates rigidly attached to the carriage,
but relative to a system of co−ordinates rigidly attached to the ground (embankment) it describes a
parabola. With the aid of this example it is clearly seen that there is no such thing as an
independently existing trajectory (lit. "path−curve"
1)
), but only a trajectory relative to a particular
body of reference.
In order to have a complete description of the motion, we must specify how the body alters its
position with time ; i.e. for every point on the trajectory it must be stated at what time the body is
situated there. These data must be supplemented by such a definition of time that, in virtue of this
definition, these time−values can be regarded essentially as magnitudes (results of measurements)
capable of observation. If we take our stand on the ground of classical mechanics, we can satisfy
this requirement for our illustration in the following manner. We imagine two clocks of identical
construction ; the man at the railway−carriage window is holding one of them, and the man on the
footpath the other. Each of the observers determines the position on his own reference−body
occupied by the stone at each tick of the clock he is holding in his hand. In this connection we have
not taken account of the inaccuracy involved by the finiteness of the velocity of propagation of light.
With this and with a second difficulty prevailing here we shall have to deal in detail later.
Next: The Galilean System of Co−ordinates
Relativity: The Special and General Theory
8
Notes
1)
That is, a curve along which the body moves.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
9
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Galileian System of Co−ordinates
As is well known, the fundamental law of the mechanics of Galilei−Newton, which is known as the
law of inertia, can be stated thus: A body removed sufficiently far from other bodies continues in a
state of rest or of uniform motion in a straight line. This law not only says something about the
motion of the bodies, but it also indicates the reference−bodies or systems of coordinates,
permissible in mechanics, which can be used in mechanical description. The visible fixed stars are
bodies for which the law of inertia certainly holds to a high degree of approximation. Now if we use
a system of co−ordinates which is rigidly attached to the earth, then, relative to this system, every
fixed star describes a circle of immense radius in the course of an astronomical day, a result which
is opposed to the statement of the law of inertia. So that if we adhere to this law we must refer
these motions only to systems of coordinates relative to which the fixed stars do not move in a
circle. A system of co−ordinates of which the state of motion is such that the law of inertia holds
relative to it is called a " Galileian system of co−ordinates." The laws of the mechanics of
Galflei−Newton can be regarded as valid only for a Galileian system of co−ordinates.
Next: The Principle of Relativity (in the restricted sense)
Relativity: The Special and General Theory
Relativity: The Special and General Theory
10
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Principle of Relativity
(in the restricted sense)
In order to attain the greatest possible clearness, let us return to our example of the railway
carriage supposed to be travelling uniformly. We call its motion a uniform translation ("uniform"
because it is of constant velocity and direction, " translation " because although the carriage
changes its position relative to the embankment yet it does not rotate in so doing). Let us imagine a
raven flying through the air in such a manner that its motion, as observed from the embankment, is
uniform and in a straight line. If we were to observe the flying raven from the moving railway
carriage. we should find that the motion of the raven would be one of different velocity and
direction, but that it would still be uniform and in a straight line. Expressed in an abstract manner
we may say : If a mass m is moving uniformly in a straight line with respect to a co−ordinate system
K, then it will also be moving uniformly and in a straight line relative to a second co−ordinate
system K
1
provided that the latter is executing a uniform translatory motion with respect to K. In
accordance with the discussion contained in the preceding section, it follows that:
If K is a Galileian co−ordinate system. then every other co−ordinate system K' is a Galileian one,
when, in relation to K, it is in a condition of uniform motion of translation. Relative to K
1
the
mechanical laws of Galilei−Newton hold good exactly as they do with respect to K.
We advance a step farther in our generalisation when we express the tenet thus: If, relative to K,
K
1
is a uniformly moving co−ordinate system devoid of rotation, then natural phenomena run their
course with respect to K
1
according to exactly the same general laws as with respect to K. This
statement is called the principle of relativity (in the restricted sense).
As long as one was convinced that all natural phenomena were capable of representation with the
help of classical mechanics, there was no need to doubt the validity of this principle of relativity. But
in view of the more recent development of electrodynamics and optics it became more and more
evident that classical mechanics affords an insufficient foundation for the physical description of all
natural phenomena. At this juncture the question of the validity of the principle of relativity became
ripe for discussion, and it did not appear impossible that the answer to this question might be in the
negative.
Nevertheless, there are two general facts which at the outset speak very much in favour of the
validity of the principle of relativity. Even though classical mechanics does not supply us with a
sufficiently broad basis for the theoretical presentation of all physical phenomena, still we must
grant it a considerable measure of " truth," since it supplies us with the actual motions of the
heavenly bodies with a delicacy of detail little short of wonderful. The principle of relativity must
therefore apply with great accuracy in the domain of mechanics. But that a principle of such broad
generality should hold with such exactness in one domain of phenomena, and yet should be invalid
for another, is a priori not very probable.
We now proceed to the second argument, to which, moreover, we shall return later. If the principle
of relativity (in the restricted sense) does not hold, then the Galileian co−ordinate systems K, K
1
,
K
2
, etc., which are moving uniformly relative to each other, will not be equivalent for the description
of natural phenomena. In this case we should be constrained to believe that natural laws are
capable of being formulated in a particularly simple manner, and of course only on condition that,
Relativity: The Special and General Theory
11
from amongst all possible Galileian co−ordinate systems, we should have chosen one (K
0
) of a
particular state of motion as our body of reference. We should then be justified (because of its
merits for the description of natural phenomena) in calling this system " absolutely at rest," and all
other Galileian systems K " in motion." If, for instance, our embankment were the system K
0
then
our railway carriage would be a system K, relative to which less simple laws would hold than with
respect to K
0
. This diminished simplicity would be due to the fact that the carriage K would be in
motion (i.e."really")with respect to K
0
. In the general laws of nature which have been formulated
with reference to K, the magnitude and direction of the velocity of the carriage would necessarily
play a part. We should expect, for instance, that the note emitted by an organpipe placed with its
axis parallel to the direction of travel would be different from that emitted if the axis of the pipe were
placed perpendicular to this direction.
Now in virtue of its motion in an orbit round the sun, our earth is comparable with a railway carriage
travelling with a velocity of about 30 kilometres per second. If the principle of relativity were not
valid we should therefore expect that the direction of motion of the earth at any moment would
enter into the laws of nature, and also that physical systems in their behaviour would be dependent
on the orientation in space with respect to the earth. For owing to the alteration in direction of the
velocity of revolution of the earth in the course of a year, the earth cannot be at rest relative to the
hypothetical system K
0
throughout the whole year. However, the most careful observations have
never revealed such anisotropic properties in terrestrial physical space, i.e. a physical
non−equivalence of different directions. This is very powerful argument in favour of the principle of
relativity.
Next: The Theorem of the Addition of Velocities Employed in Classical Mechanics
Relativity: The Special and General Theory
Relativity: The Special and General Theory
12
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Theorem of the
Addition of Velocities
Employed in Classical Mechanics
Let us suppose our old friend the railway carriage to be travelling along the rails with a constant
velocity v, and that a man traverses the length of the carriage in the direction of travel with a
velocity w. How quickly or, in other words, with what velocity W does the man advance relative to
the embankment during the process ? The only possible answer seems to result from the following
consideration: If the man were to stand still for a second, he would advance relative to the
embankment through a distance v equal numerically to the velocity of the carriage. As a
consequence of his walking, however, he traverses an additional distance w relative to the
carriage, and hence also relative to the embankment, in this second, the distance w being
numerically equal to the velocity with which he is walking. Thus in total be covers the distance
W=v+w relative to the embankment in the second considered. We shall see later that this result,
which expresses the theorem of the addition of velocities employed in classical mechanics, cannot
be maintained ; in other words, the law that we have just written down does not hold in reality. For
the time being, however, we shall assume its correctness.
Next: The Apparent Incompatability of the Law of Propagation of Light with the Principle of Relativity
Relativity: The Special and General Theory
Relativity: The Special and General Theory
13
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Apparent Incompatibility of the
Law of Propagation of Light with the
Principle of Relativity
There is hardly a simpler law in physics than that according to which light is propagated in empty
space. Every child at school knows, or believes he knows, that this propagation takes place in
straight lines with a velocity c= 300,000 km./sec. At all events we know with great exactness that
this velocity is the same for all colours, because if this were not the case, the minimum of emission
would not be observed simultaneously for different colours during the eclipse of a fixed star by its
dark neighbour. By means of similar considerations based on observa− tions of double stars, the
Dutch astronomer De Sitter was also able to show that the velocity of propagation of light cannot
depend on the velocity of motion of the body emitting the light. The assumption that this velocity of
propagation is dependent on the direction "in space" is in itself improbable.
In short, let us assume that the simple law of the constancy of the velocity of light c (in vacuum) is
justifiably believed by the child at school. Who would imagine that this simple law has plunged the
conscientiously thoughtful physicist into the greatest intellectual difficulties? Let us consider how
these difficulties arise.
Of course we must refer the process of the propagation of light (and indeed every other process) to
a rigid reference−body (co−ordinate system). As such a system let us again choose our
embankment. We shall imagine the air above it to have been removed. If a ray of light be sent
along the embankment, we see from the above that the tip of the ray will be transmitted with the
velocity c relative to the embankment. Now let us suppose that our railway carriage is again
travelling along the railway lines with the velocity v, and that its direction is the same as that of the
ray of light, but its velocity of course much less. Let us inquire about the velocity of propagation of
the ray of light relative to the carriage. It is obvious that we can here apply the consideration of the
previous section, since the ray of light plays the part of the man walking along relatively to the
carriage. The velocity W of the man relative to the embankment is here replaced by the velocity of
light relative to the embankment. w is the required velocity of light with respect to the carriage, and
we have
w = c−v.
The velocity of propagation ot a ray of light relative to the carriage thus comes cut smaller than c.
But this result comes into conflict with the principle of relativity set forth in Section V. For, like every
other general law of nature, the law of the transmission of light in vacuo [in vacuum] must,
according to the principle of relativity, be the same for the railway carriage as reference−body as
when the rails are the body of reference. But, from our above consideration, this would appear to
be impossible. If every ray of light is propagated relative to the embankment with the velocity
c, then for this reason it would appear that another law of propagation of light must necessarily hold
with respect to the carriage — a result contradictory to the principle of relativity.
In view of this dilemma there appears to be nothing else for it than to abandon either the principle
of relativity or the simple law of the propagation of light in vacuo. Those of you who have carefully
followed the preceding discussion are almost sure to expect that we should retain the principle of
Relativity: The Special and General Theory
14
relativity, which appeals so convincingly to the intellect because it is so natural and simple. The law
of the propagation of light in vacuo would then have to be replaced by a more complicated law
conformable to the principle of relativity. The development of theoretical physics shows, however,
that we cannot pursue this course. The epoch−making theoretical investigations of H. A. Lorentz on
the electrodynamical and optical phenomena connected with moving bodies show that experience
in this domain leads conclusively to a theory of electromagnetic phenomena, of which the law of the
constancy of the velocity of light in vacuo is a necessary conse. quence. Prominent theoretical
physicists were theref ore more inclined to reject the principle of relativity, in spite of the fact that no
empirical data had been found which were contradictory to this principle.
At this juncture the theory of relativity entered the arena. As a result of an analysis of the physical
conceptions of time and space, it became evident that in realily there is not the least incompatibilitiy
between the principle of relativity and the law of propagation of light, and that by systematically
holding fast to both these laws a logically rigid theory could be arrived at. This theory has been
called the special theory of relativity to distinguish it from the extended theory, with which we shall
deal later. In the following pages we shall present the fundamental ideas of the special theory of
relativity.
Next: On the Idea of Time in Physics
Relativity: The Special and General Theory
Relativity: The Special and General Theory
15
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
On the Idea of Time in Physics
Lightning has struck the rails on our railway embankment at two places A and B far distant from
each other. I make the additional assertion that these two lightning flashes occurred
simultaneously. If I ask you whether there is sense in this statement, you will answer my question
with a decided "Yes." But if I now approach you with the request to explain to me the sense of the
statement more precisely, you find after some consideration that the answer to this question is not
so easy as it appears at first sight.
After some time perhaps the following answer would occur to you: "The significance of the
statement is clear in itself and needs no further explanation; of course it would require some
consideration if I were to be commissioned to determine by observations whether in the actual case
the two events took place simultaneously or not." I cannot be satisfied with this answer for the
following reason. Supposing that as a result of ingenious considerations an able meteorologist
were to discover that the lightning must always strike the places A and B simultaneously, then we
should be faced with the task of testing whether or not this theoretical result is in accordance with
the reality. We encounter the same difficulty with all physical statements in which the conception "
simultaneous " plays a part. The concept does not exist for the physicist until he has the possibility
of discovering whether or not it is fulfilled in an actual case. We thus require a definition of
simultaneity such that this definition supplies us with the method by means of which, in the present
case, he can decide by experiment whether or not both the lightning strokes occurred
simultaneously. As long as this requirement is not satisfied, I allow myself to be deceived as a
physicist (and of course the same applies if I am not a physicist), when I imagine that I am able to
attach a meaning to the statement of simultaneity. (I would ask the reader not to proceed farther
until he is fully convinced on this point.)
After thinking the matter over for some time you then offer the following suggestion with which to
test simultaneity. By measuring along the rails, the connecting line AB should be measured up and
an observer placed at the mid−point M of the distance AB. This observer should be supplied with
an arrangement (e.g. two mirrors inclined at 90
0
) which allows him visually to observe both places
A and B at the same time. If the observer perceives the two flashes of lightning at the same time,
then they are simultaneous.
I am very pleased with this suggestion, but for all that I cannot regard the matter as quite settled,
because I feel constrained to raise the following objection:
"Your definition would certainly be right, if only I knew that the light by means of which the observer
at M perceives the lightning flashes travels along the length A M with the same velocity as
along the length B M. But an examination of this supposition would only be possible if we
already had at our disposal the means of measuring time. It would thus appear as though we were
moving here in a logical circle."
After further consideration you cast a somewhat disdainful glance at me — and rightly so — and
you declare:
"I maintain my previous definition nevertheless, because in reality it assumes absolutely nothing
about light. There is only one demand to be made of the definition of simultaneity, namely, that in
Relativity: The Special and General Theory
16
every real case it must supply us with an empirical decision as to whether or not the conception
that has to be defined is fulfilled. That my definition satisfies this demand is indisputable. That light
requires the same time to traverse the path A M as for the path B M is in reality neither a
supposition nor a hypothesis about the physical nature of light, but a stipulation which I can make
of my own freewill in order to arrive at a definition of simultaneity."
It is clear that this definition can be used to give an exact meaning not only to two events, but to as
many events as we care to choose, and independently of the positions of the scenes of the events
with respect to the body of reference
1)
(here the railway embankment). We are thus led also to a
definition of " time " in physics. For this purpose we suppose that clocks of identical construction
are placed at the points A, B and C of the railway line (co−ordinate system) and that they are set in
such a manner that the positions of their pointers are simultaneously (in the above sense) the
same. Under these conditions we understand by the " time " of an event the reading (position of the
hands) of that one of these clocks which is in the immediate vicinity (in space) of the event. In this
manner a time−value is associated with every event which is essentially capable of observation.
This stipulation contains a further physical hypothesis, the validity of which will hardly be doubted
without empirical evidence to the contrary. It has been assumed that all these clocks go at the
same rate if they are of identical construction. Stated more exactly: When two clocks arranged at
rest in different places of a reference−body are set in such a manner that a particular position of the
pointers of the one clock is simultaneous (in the above sense) with the same position, of the
pointers of the other clock, then identical " settings " are always simultaneous (in the sense of the
above definition).
Next: The Relativity of Simultaneity
Footnotes
1)
We suppose further, that, when three events A, B and C occur in different places in such a
manner that A is simultaneous with B and B is simultaneous with C (simultaneous in the sense of
the above definition), then the criterion for the simultaneity of the pair of events A, C is also
satisfied. This assumption is a physical hypothesis about the the of propagation of light: it must
certainly be fulfilled if we are to maintain the law of the constancy of the velocity of light in vacuo.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
17
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Relativity of Simulatneity
Up to now our considerations have been referred to a particular body of reference, which we have
styled a " railway embankment." We suppose a very long train travelling along the rails with the
constant velocity v and in the direction indicated in Fig 1. People travelling in this train will with
advantage uew the train as a rigid reference−body (co−ordinate system); they regard all events in
reference to the train. Then every event which takes place along the line also takes place at a
particular point of the train. Also the definition of simultaneity can be given relative to the train in
exactly the same way as with respect to the embankment. As a natural consequence, however, the
following question arises :
Are two events (e.g. the two strokes of lightning A and B) which are simultaneous with reference to
the railway embankment also simultaneous relatively to the train? We shall show directly that the
answer must be in the negative.
When we say that the lightning strokes A and B are simultaneous with respect to be embankment,
we mean: the rays of light emitted at the places A and B, where the lightning occurs, meet each
other at the mid−point M of the length A B of the embankment. But the events A and B also
correspond to positions A and B on the train. Let M
1
be the mid−point of the distance A B on
the travelling train. Just when the flashes (as judged from the embankment) of lightning occur, this
point M
1
naturally coincides with the point M but it moves towards the right in the diagram with the
velocity v of the train. If an observer sitting in the position M
1
in the train did not possess this
velocity, then he would remain permanently at M, and the light rays emitted by the flashes of
lightning A and B would reach him simultaneously, i.e. they would meet just where he is situated.
Now in reality (considered with reference to the railway embankment) he is hastening towards the
beam of light coming from B, whilst he is riding on ahead of the beam of light coming from A.
Hence the observer will see the beam of light emitted from B earlier than he will see that emitted
from A. Observers who take the railway train as their reference−body must therefore come to the
conclusion that the lightning flash B took place earlier than the lightning flash A. We thus arrive at
the important result:
Events which are simultaneous with reference to the embankment are not simultaneous with
respect to the train, and vice versa (relativity of simultaneity). Every reference−body (co−ordinate
system) has its own particular time ; unless we are told the reference−body to which the statement
of time refers, there is no meaning in a statement of the time of an event.
Relativity: The Special and General Theory
18
Now before the advent of the theory of relativity it had always tacitly been assumed in physics that
the statement of time had an absolute significance, i.e. that it is independent of the state of motion
of the body of reference. But we have just seen that this assumption is incompatible with the most
natural definition of simultaneity; if we discard this assumption, then the conflict between the law of
the propagation of light in vacuo and the principle of relativity (developed in Section 7) disappears.
We were led to that conflict by the considerations of Section 6, which are now no longer tenable. In
that section we concluded that the man in the carriage, who traverses the distance w per
second relative to the carriage, traverses the same distance also with respect to the embankment
in each second of time. But, according to the foregoing considerations, the time required by a
particular occurrence with respect to the carriage must not be considered equal to the duration of
the same occurrence as judged from the embankment (as reference−body). Hence it cannot be
contended that the man in walking travels the distance w relative to the railway line in a time which
is equal to one second as judged from the embankment.
Moreover, the considerations of Section 6 are based on yet a second assumption, which, in the
light of a strict consideration, appears to be arbitrary, although it was always tacitly made even
before the introduction of the theory of relativity.
Next: On the Relativity of the Conception of Distance
Relativity: The Special and General Theory
Relativity: The Special and General Theory
19
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
On the Relativity of the Conception of Distance
Let us consider two particular points on the train
1)
travelling along the embankment with the
velocity v, and inquire as to their distance apart. We already know that it is necessary to have a
body of reference for the measurement of a distance, with respect to which body the distance can
be measured up. It is the simplest plan to use the train itself as reference−body (co−ordinate
system). An observer in the train measures the interval by marking off his measuring−rod in a
straight line (e.g. along the floor of the carriage) as many times as is necessary to take him from
the one marked point to the other. Then the number which tells us how often the rod has to be laid
down is the required distance.
It is a different matter when the distance has to be judged from the railway line. Here the following
method suggests itself. If we call A
1
and B
1
the two points on the train whose distance apart is
required, then both of these points are moving with the velocity v along the embankment. In the first
place we require to determine the points A and B of the embankment which are just being passed
by the two points A
1
and B
1
at a particular time t — judged from the embankment. These points
A and B of the embankment can be determined by applying the definition of time given in Section 8.
The distance between these points A and B is then measured by repeated application of thee
measuring−rod along the embankment.
A priori it is by no means certain that this last measurement will supply us with the same result as
the first. Thus the length of the train as measured from the embankment may be different from that
obtained by measuring in the train itself. This circumstance leads us to a second objection which
must be raised against the apparently obvious consideration of Section 6. Namely, if the man in the
carriage covers the distance w in a unit of time — measured from the train, — then this distance —
as measured from the embankment — is not necessarily also equal to w.
Next: The Lorentz Transformation
Footnotes
1)
e.g. the middle of the first and of the twentieth carriage.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
20
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Lorentz Transformation
The results of the last three sections show that the apparent incompatibility of the law of
propagation of light with the principle of relativity (Section 7) has been derived by means of a
consideration which borrowed two unjustifiable hypotheses from classical mechanics; these are as
follows:
(1) The time−interval (time) between two events is independent of the condition of motion of the
body of reference.
(2) The space−interval (distance) between two points of a rigid body is independent of the condition
of motion of the body of reference.
If we drop these hypotheses, then the dilemma of Section 7 disappears, because the theorem of
the addition of velocities derived in Section 6 becomes invalid. The possibility presents itself that
the law of the propagation of light in vacuo may be compatible with the principle of relativity, and
the question arises: How have we to modify the considerations of Section 6 in order to remove the
apparent disagreement between these two fundamental results of experience? This question leads
to a general one. In the discussion of Section 6 we have to do with places and times relative both to
the train and to the embankment. How are we to find the place and time of an event in relation to
the train, when we know the place and time of the event with respect to the railway embankment ?
Is there a thinkable answer to this question of such a nature that the law of transmission of light in
vacuo does not contradict the principle of relativity ? In other words : Can we conceive of a relation
between place and time of the individual events relative to both reference−bodies, such that every
ray of light possesses the velocity of transmission c relative to the embankment and relative to the
train ? This question leads to a quite definite positive answer, and to a perfectly definite
transformation law for the space−time magnitudes of an event when changing over from one body
of reference to another.
Before we deal with this, we shall introduce the following incidental consideration. Up to the present
we have only considered events taking place along the embankment, which had mathematically to
assume the function of a straight line. In the manner indicated in Section 2 we can imagine this
reference−body supplemented laterally and in a vertical direction by means of a framework of rods,
so that an event which takes place anywhere can be localised with reference to this framework.
Similarly, we can imagine the train travelling with the velocity v to
be continued across the whole of space, so that every event, no matter how far off it may be, could
also be localised with respect to the second framework. Without committing any fundamental error,
Relativity: The Special and General Theory
21
we can disregard the fact that in reality these frameworks would continually interfere with each
other, owing to the impenetrability of solid bodies. In every such framework we imagine three
surfaces perpendicular to each other marked out, and designated as " co−ordinate planes " ("
co−ordinate system "). A co−ordinate system K then corresponds to the embankment, and a
co−ordinate system K' to the train. An event, wherever it may have taken place, would be fixed in
space with respect to K by the three perpendiculars x, y, z on the co−ordinate planes, and with
regard to time by a time value t. Relative to K
1
, the same event would be fixed in respect of space
and time by corresponding values x
1
, y
1
, z
1
, t
1
, which of course are not identical with x, y, z, t. It has
already been set forth in detail how these magnitudes are to be regarded as results of physical
measurements.
Obviously our problem can be exactly formulated in the following manner. What are the values x
1
,
y
1
, z
1
, t
1
, of an event with respect to K
1
, when the magnitudes x, y, z, t, of the same event with
respect to K are given ? The relations must be so chosen that the law of the transmission of light in
vacuo is satisfied for one and the same ray of light (and of course for every ray) with respect to
K and K
1
. For the relative orientation in space of the co−ordinate systems indicated in the diagram
(Fig. 2), this problem is solved by means of the equations :
y
1
= y
z
1
= z
This system of equations is known as the " Lorentz transformation."
1)
If in place of the law of transmission of light we had taken as our basis the tacit assumptions of the
older mechanics as to the absolute character of times and lengths, then instead of the above we
should have obtained the following equations:
x
1
= x − vt
y
1
= y
z
1
= z
t
1
= t
This system of equations is often termed the " Galilei transformation." The Galilei transformation
can be obtained from the Lorentz transformation by substituting an infinitely large value for the
velocity of light c in the latter transformation.
Aided by the following illustration, we can readily see that, in accordance with the Lorentz
transformation, the law of the transmission of light in vacuo is satisfied both for the reference−body
K and for the reference−body K
1
. A light−signal is sent along the positive x−axis, and this
light−stimulus advances in accordance with the equation
x = ct,
i.e. with the velocity c. According to the equations of the Lorentz transformation, this simple relation
between x and t involves a relation between x
1
and t
1
. In point of fact, if we substitute for x the
Relativity: The Special and General Theory
22
value ct in the first and fourth equations of the Lorentz transformation, we obtain:
from which, by division, the expression
x
1
= ct
1
immediately follows. If referred to the system K
1
, the propagation of light takes place according to
this equation. We thus see that the velocity of transmission relative to the reference−body K
1
is also
equal to c. The same result is obtained for rays of light advancing in any other direction
whatsoever. Of cause this is not surprising, since the equations of the Lorentz transformation were
derived conformably to this point of view.
Next: The Behaviour of Measuring−Rods and Clocks in Motion
Footnotes
1)
A simple derivation of the Lorentz transformation is given in Appendix I.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
23
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Behaviour of Measuring−Rods and Clocks in Motion
Place a metre−rod in the x
1
−axis of K
1
in such a manner that one end (the beginning) coincides
with the point x
1
=0 whilst the other end (the end of the rod) coincides with the point x
1
=I. What is
the length of the metre−rod relatively to the system K? In order to learn this, we need only ask
where the beginning of the rod and the end of the rod lie with respect to K at a particular time t of
the system K. By means of the first equation of the Lorentz transformation the values of these two
points at the time t = 0 can be shown to be
the distance between the points being .
But the metre−rod is moving with the velocity v relative to K. It therefore follows that the length of a
rigid metre−rod moving in the direction of its length with a velocity v is of a metre.
The rigid rod is thus shorter when in motion than when at rest, and the more quickly it is moving,
the shorter is the rod. For the velocity v=c we should have ,
and for stiII greater velocities the square−root becomes imaginary. From this we conclude that in
the theory of relativity the velocity c plays the part of a limiting velocity, which can neither be
reached nor exceeded by any real body.
Of course this feature of the velocity c as a limiting velocity also clearly follows from the equations
of the Lorentz transformation, for these became meaningless if we choose values of v greater than
c.
If, on the contrary, we had considered a metre−rod at rest in the x−axis with respect to K, then we
should have found that the length of the rod as judged from K
1
would have been ;
this is quite in accordance with the principle of relativity which forms the basis of our
considerations.
A Priori it is quite clear that we must be able to learn something about the physical behaviour of
measuring−rods and clocks from the equations of transformation, for the magnitudes z, y, x, t, are
nothing more nor less than the results of measurements obtainable by means of measuring−rods
and clocks. If we had based our considerations on the Galileian transformation we should not have
Relativity: The Special and General Theory
24
obtained a contraction of the rod as a consequence of its motion.
Let us now consider a seconds−clock which is permanently situated at the origin (x
1
=0) of K
1
. t
1
=0
and t
1
=I are two successive ticks of this clock. The first and fourth equations of the Lorentz
transformation give for these two ticks :
t = 0
and
As judged from K, the clock is moving with the velocity v; as judged from this reference−body, the
time which elapses between two strokes of the clock is not one second, but
seconds, i.e. a somewhat larger time. As a consequence of its motion the clock goes more slowly
than when at rest. Here also the velocity c plays the part of an unattainable limiting velocity.
Next: Theorem of the Addition of Velocities
Relativity: The Special and General Theory
Relativity: The Special and General Theory
25
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
Theorem of the Addition of Velocities.
The Experiment of Fizeau
Now in practice we can move clocks and measuring−rods only with velocities that are small
compared with the velocity of light; hence we shall hardly be able to compare the results of the
previous section directly with the reality. But, on the other hand, these results must strike you as
being very singular, and for that reason I shall now draw another conclusion from the theory, one
which can easily be derived from the foregoing considerations, and which has been most elegantly
confirmed by experiment.
In Section 6 we derived the theorem of the addition of velocities in one direction in the form which
also results from the hypotheses of classical mechanics− This theorem can also be deduced
readily horn the Galilei transformation (Section 11). In place of the man walking inside the carriage,
we introduce a point moving relatively to the co−ordinate system K
1
in accordance with the
equation
x
1
= wt
1
By means of the first and fourth equations of the Galilei transformation we can express x
1
and t
1
in
terms of x and t, and we then obtain
x = (v + w)t
This equation expresses nothing else than the law of motion of the point with reference to the
system K (of the man with reference to the embankment). We denote this velocity by the symbol W,
and we then obtain, as in Section 6,
W=v+w A)
But we can carry out this consideration just as well on the basis of the theory of relativity. In the
equation
x
1
= wt
1
B)
we must then express x
1
and t
1
in terms of x and t, making use of the first and fourth equations of
the Lorentz transformation. Instead of the equation (A) we then obtain the equation
which corresponds to the theorem of addition for velocities in one direction according to the theory
of relativity. The question now arises as to which of these two theorems is the better in accord with
experience. On this point we axe enlightened by a most important experiment which the brilliant
physicist Fizeau performed more than half a century ago, and which has been repeated since then
Relativity: The Special and General Theory
26
by some of the best experimental physicists, so that there can be no doubt about its result. The
experiment is concerned with the following question. Light travels in a motionless liquid with a
particular velocity w. How quickly does it travel in the direction of the arrow in the tube T (see the
accompanying diagram, Fig. 3) when the liquid above mentioned is flowing through the tube with a
velocity v ?
In accordance with the principle of relativity we shall certainly have to take for granted that the
propagation of light always takes place with the same velocity w with respect to the liquid, whether
the latter is in motion with reference to other bodies or not. The velocity of light relative to the liquid
and the velocity of the latter relative to the tube are thus known, and we require the velocity of light
relative to the tube.
It is clear that we have the problem of Section 6 again before us. The tube plays the part of the
railway embankment or of the co−ordinate system K, the liquid plays the part of the carriage or of
the co−ordinate system K
1
, and finally, the light plays the part of the
man walking along the carriage, or of the moving point in the present section. If we denote the
velocity of the light relative to the tube by W, then this is given by the equation (A) or (B), according
as the Galilei transformation or the Lorentz transformation corresponds to the facts.
Experiment
1)
decides in favour of equation (B) derived from the theory of relativity, and the
agreement is, indeed, very exact. According to recent and most excellent measurements by
Zeeman, the influence of the velocity of flow v on the propagation of light is represented by formula
(B) to within one per cent.
Nevertheless we must now draw attention to the fact that a theory of this phenomenon was given
by H. A. Lorentz long before the statement of the theory of relativity. This theory was of a purely
electrodynamical nature, and was obtained by the use of particular hypotheses as to the
electromagnetic structure of matter. This circumstance, however, does not in the least diminish the
conclusiveness of the experiment as a crucial test in favour of the theory of relativity, for the
electrodynamics of Maxwell−Lorentz, on which the original theory was based, in no way opposes
the theory of relativity. Rather has the latter been developed trom electrodynamics as an
astoundingly simple combination and generalisation of the hypotheses, formerly independent of
each other, on which electrodynamics was built.
Next: The Heuristic Value of the Theory of Relativity
Footnotes
1)
Fizeau found
, where
is the index of refraction of the liquid. On the other hand, owing to the smallness of as
Relativity: The Special and General Theory
27
compared with I,
we can replace (B) in the first place by , or to the same order of
approximation by
, which agrees with Fizeau's result.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
28
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
The Heuristic Value of the Theory of Relativity
Our train of thought in the foregoing pages can be epitomised in the following manner. Experience
has led to the conviction that, on the one hand, the principle of relativity holds true and that on the
other hand the velocity of transmission of light in vacuo has to be considered equal to a constant c.
By uniting these two postulates we obtained the law of transformation for the rectangular
co−ordinates x, y, z and the time t of the events which constitute the processes of nature. In this
connection we did not obtain the Galilei transformation, but, differing from classical mechanics, the
Lorentz transformation.
The law of transmission of light, the acceptance of which is justified by our actual knowledge,
played an important part in this process of thought. Once in possession of the Lorentz
transformation, however, we can combine this with the principle of relativity, and sum up the theory
thus:
Every general law of nature must be so constituted that it is transformed into a law of exactly the
same form when, instead of the space−time variables x, y, z, t of the original coordinate system K,
we introduce new space−time variables x
1
, y
1
, z
1
, t
1
of a co−ordinate system K
1
. In this connection
the relation between the ordinary and the accented magnitudes is given by the Lorentz
transformation. Or in brief : General laws of nature are co−variant with respect to Lorentz
transformations.
This is a definite mathematical condition that the theory of relativity demands of a natural law, and
in virtue of this, the theory becomes a valuable heuristic aid in the search for general laws of
nature. If a general law of nature were to be found which did not satisfy this condition, then at least
one of the two fundamental assumptions of the theory would have been disproved. Let us now
examine what general results the latter theory has hitherto evinced.
Next: General Results of the Theory
Relativity: The Special and General Theory
Relativity: The Special and General Theory
29
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
General Results of the Theory
It is clear from our previous considerations that the (special) theory of relativity has grown out of
electrodynamics and optics. In these fields it has not appreciably altered the predictions of theory,
but it has considerably simplified the theoretical structure, i.e. the derivation of laws, and — what is
incomparably more important — it has considerably reduced the number of independent hypothese
forming the basis of theory. The special theory of relativity has rendered the Maxwell−Lorentz
theory so plausible, that the latter would have been generally accepted by physicists even if
experiment had decided less unequivocally in its favour.
Classical mechanics required to be modified before it could come into line with the demands of the
special theory of relativity. For the main part, however, this modification affects only the laws for
rapid motions, in which the velocities of matter v are not very small as compared with the velocity of
light. We have experience of such rapid motions only in the case of electrons and ions; for other
motions the variations from the laws of classical mechanics are too small to make themselves
evident in practice. We shall not consider the motion of stars until we come to speak of the general
theory of relativity. In accordance with the theory of relativity the kinetic energy of a material point of
mass m is no longer given by the well−known expression
but by the expression
This expression approaches infinity as the velocity v approaches the velocity of light c. The velocity
must therefore always remain less than c, however great may be the energies used to produce the
acceleration. If we develop the expression for the kinetic energy in the form of a series, we obtain
When is small compared with unity, the third of these terms is always small in comparison with
the second,
which last is alone considered in classical mechanics. The first term mc
2
does not contain the
velocity, and requires no consideration if we are only dealing with the question as to how the
energy of a point−mass; depends on the velocity. We shall speak of its essential significance later.
Relativity: The Special and General Theory
30
The most important result of a general character to which the special theory of relativity has led is
concerned with the conception of mass. Before the advent of relativity, physics recognised two
conservation laws of fundamental importance, namely, the law of the canservation of energy and
the law of the conservation of mass these two fundamental laws appeared to be quite independent
of each other. By means of the theory of relativity they have been united into one law. We shall now
briefly consider how this unification came about, and what meaning is to be attached to it.
The principle of relativity requires that the law of the concervation of energy should hold not only
with reference to a co−ordinate system K, but also with respect to every co−ordinate system
K
1
which is in a state of uniform motion of translation relative to K, or, briefly, relative to every "
Galileian " system of co−ordinates. In contrast to classical mechanics; the Lorentz transformation is
the deciding factor in the transition from one such system to another.
By means of comparatively simple considerations we are led to draw the following conclusion from
these premises, in conjunction with the fundamental equations of the electrodynamics of Maxwell:
A body moving with the velocity v, which absorbs
1)
an amount of energy E
0
in the form of radiation
without suffering an alteration in velocity in the process, has, as a consequence, its energy
increased by an amount
In consideration of the expression given above for the kinetic energy of the body, the required
energy of the body comes out to be
Thus the body has the same energy as a body of mass
moving with the velocity v. Hence we can say: If a body takes up an amount of energy E
0
, then its
inertial mass increases by an amount
the inertial mass of a body is not a constant but varies according to the change in the energy of the
body. The inertial mass of a system of bodies can even be regarded as a measure of its energy.
The law of the conservation of the mass of a system becomes identical with the law of the
conservation of energy, and is only valid provided that the system neither takes up nor sends out
energy. Writing the expression for the energy in the form
Relativity: The Special and General Theory
31
we see that the term mc
2
, which has hitherto attracted our attention, is nothing else than the energy
possessed by the body
2)
before it absorbed the energy E
0
.
A direct comparison of this relation with experiment is not possible at the present time (1920; see
Note, p. 48), owing to the fact that the changes in energy E
0
to which we can Subject a system are
not large enough to make themselves perceptible as a change in the inertial mass of the system.
is too small in comparison with the mass m, which was present before the alteration of the energy.
It is owing to this circumstance that classical mechanics was able to establish successfully the
conservation of mass as a law of independent validity.
Let me add a final remark of a fundamental nature. The success of the Faraday−Maxwell
interpretation of electromagnetic action at a distance resulted in physicists becoming convinced
that there are no such things as instantaneous actions at a distance (not involving an intermediary
medium) of the type of Newton's law of gravitation. According to the theory of relativity, action at a
distance with the velocity of light always takes the place of instantaneous action at a distance or of
action at a distance with an infinite velocity of transmission. This is connected with the fact that the
velocity c plays a fundamental role in this theory. In Part II we shall see in what way this result
becomes modified in the general theory of relativity.
Next: Experience and the Special Theory of Relativity
Footnotes
1)
E
0
is the energy taken up, as judged from a co−ordinate system moving with the body.
2)
As judged from a co−ordinate system moving with the body.
[Note]
With the advent of nuclear transformation processes, which result from the bombardment of
elements by ±−particles, protons, deuterous, neutrons or ³−rays, the equivalence of mass and
energy expressed by the ralation E = mc
2
has been amply confirmed. The sum of the reacting
masses, together with the mass equivalent of the kinetic energy of the bombarding particle (or
photon), is always greater than the sum of the resulting masses. The difference is the equivalent
mass of the kinetic energy of the particles generated, or of the released electromagnetic energy
(³−photons). In the same way, the mass of a spontaneously disintegrating radioactive atom is
always greater than the sum of the masses of the resulting atoms by the mass equivalent of the
kinetic energy of the particles generated (or of the photonic energy). Measurements of the energy
of the rays emitted in nuclear reactions, in combination with the equations of such reactions, render
it possible to evaluate atomic weights to a high degree of accuracy. [Note by the translator]
Relativity: The Special and General Theory
32
Relativity: The Special and General Theory
Relativity: The Special and General Theory
33
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
Experience and the Special Theory of Relativity
To what extent is the special theory of relativity supported by experience ? This question is not
easily answered for the reason already mentioned in connection with the fundamental experiment
of Fizeau. The special theory of relativity has crystallised out from the Maxwell−Lorentz theory of
electromagnetic phenomena. Thus all facts of experience which support the electromagnetic theory
also support the theory of relativity. As being of particular importance, I mention here the fact that
the theory of relativity enables us to predict the effects produced on the light reaching us from the
fixed stars. These results are obtained in an exceedingly simple manner, and the effects indicated,
which are due to the relative motion of the earth with reference to those fixed stars are found to be
in accord with experience. We refer to the yearly movement of the apparent position of the fixed
stars resulting from the motion of the earth round the sun (aberration), and to the influence of the
radial components of the relative motions of the fixed stars with respect to the earth on the colour of
the light reaching us from them. The latter effect manifests itself in a slight displacement of the
spectral lines of the light transmitted to us from a fixed star, as compared with the position of the
same spectral lines when they are produced by a terrestrial source of light (Doppler principle). The
experimental arguments in favour of the Maxwell−Lorentz theory, which are at the same time
arguments in favour of the theory of relativity, are too numerous to be set forth here. In reality they
limit the theoretical possibilities to such an extent, that no other theory than that of Maxwell and
Lorentz has been able to hold its own when tested by experience.
But there are two classes of experimental facts hitherto obtained which can be represented in the
Maxwell−Lorentz theory only by the introduction of an auxiliary hypothesis, which in itself —
i.e. without making use of the theory of relativity — appears extraneous.
It is known that cathode rays and the so−called ²−rays emitted by radioactive substances consist of
negatively electrified particles (electrons) of very small inertia and large velocity. By examining the
deflection of these rays under the influence of electric and magnetic fields, we can study the law of
motion of these particles very exactly.
In the theoretical treatment of these electrons, we are faced with the difficulty that electrodynamic
theory of itself is unable to give an account of their nature. For since electrical masses of one sign
repel each other, the negative electrical masses constituting the electron would necessarily be
scattered under the influence of their mutual repulsions, unless there are forces of another kind
operating between them, the nature of which has hitherto remained obscure to us.
1)
If we now
assume that the relative distances between the electrical masses constituting the electron remain
unchanged during the motion of the electron (rigid connection in the sense of classical mechanics),
we arrive at a law of motion of the electron which does not agree with experience. Guided by purely
formal points of view, H. A. Lorentz was the first to introduce the hypothesis that the form of the
electron experiences a contraction in the direction of motion in consequence of that motion. the
contracted length being proportional to the expression
This, hypothesis, which is not justifiable by any electrodynamical facts, supplies us then with that
Relativity: The Special and General Theory
34
particular law of motion which has been confirmed with great precision in recent years.
The theory of relativity leads to the same law of motion, without requiring any special hypothesis
whatsoever as to the structure and the behaviour of the electron. We arrived at a similar conclusion
in Section 13 in connection with the experiment of Fizeau, the result of which is foretold by the
theory of relativity without the necessity of drawing on hypotheses as to the physical nature of the
liquid.
The second class of facts to which we have alluded has reference to the question whether or not
the motion of the earth in space can be made perceptible in terrestrial experiments. We have
already remarked in Section 5 that all attempts of this nature led to a negative result. Before the
theory of relativity was put forward, it was difficult to become reconciled to this negative result, for
reasons now to be discussed. The inherited prejudices about time and space did not allow any
doubt to arise as to the prime importance of the Galileian transformation for changing over from
one body of reference to another. Now assuming that the Maxwell−Lorentz equations hold for a
reference−body K, we then find that they do not hold for a reference−body K
1
moving uniformly with
respect to K, if we assume that the relations of the Galileian transformstion exist between the
co−ordinates of K and K
1
. It thus appears that, of all Galileian co−ordinate systems, one (K)
corresponding to a particular state of motion is physically unique. This result was interpreted
physically by regarding K as at rest with respect to a hypothetical æther of space. On the other
hand, all coordinate systems K
1
moving relatively to K were to be regarded as in motion with
respect to the æther. To this motion of K
1
against the æther ("æther−drift " relative to K
1
) were
attributed the more complicated laws which were supposed to hold relative to K
1
. Strictly speaking,
such an æther−drift ought also to be assumed relative to the earth, and for a long time the efforts of
physicists were devoted to attempts to detect the existence of an æther−drift at the earth's surface.
In one of the most notable of these attempts Michelson devised a method which appears as though
it must be decisive. Imagine two mirrors so arranged on a rigid body that the reflecting surfaces
face each other. A ray of light requires a perfectly definite time T to pass from one mirror to the
other and back again, if the whole system be at rest with respect to the æther. It is found by
calculation, however, that a slightly different time T
1
is required for this process, if the body,
together with the mirrors, be moving relatively to the æther. And yet another point: it is shown by
calculation that for a given velocity v with reference to the æther, this time T
1
is different when the
body is moving perpendicularly to the planes of the mirrors from that resulting when the motion is
parallel to these planes. Although the estimated difference between these two times is exceedingly
small, Michelson and Morley performed an experiment involving interference in which this
difference should have been clearly detectable. But the experiment gave a negative result — a fact
very perplexing to physicists. Lorentz and FitzGerald rescued the theory from this difficulty by
assuming that the motion of the body relative to the æther produces a contraction of the body in the
direction of motion, the amount of contraction being just sufficient to compensate for the differeace
in time mentioned above. Comparison with the discussion in Section 11 shows that also from the
standpoint of the theory of relativity this solution of the difficulty was the right one. But on the basis
of the theory of relativity the method of interpretation is incomparably more satisfactory. According
to this theory there is no such thing as a " specially favoured " (unique) co−ordinate system to
occasion the introduction of the æther−idea, and hence there can be no æther−drift, nor any
experiment with which to demonstrate it. Here the contraction of moving bodies follows from the
two fundamental principles of the theory, without the introduction of particular hypotheses ; and as
the prime factor involved in this contraction we find, not the motion in itself, to which we cannot
attach any meaning, but the motion with respect to the body of reference chosen in the particular
case in point. Thus for a co−ordinate system moving with the earth the mirror system of Michelson
and Morley is not shortened, but it is shortened for a co−ordinate system which is at rest relatively
to the sun.
Relativity: The Special and General Theory
35
Next: Minkowski's Four−Dimensional Space
Footnotes
1)
The general theory of relativity renders it likely that the electrical masses of an electron are held
together by gravitational forces.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
36
Albert Einstein: Relativity
Part I: The Special Theory of Relativity
Minkowski's Four−Dimensional Space
The non−mathematician is seized by a mysterious shuddering when he hears of
"four−dimensional" things, by a feeling not unlike that awakened by thoughts of the occult. And yet
there is no more common−place statement than that the world in which we live is a
four−dimensional space−time continuum.
Space is a three−dimensional continuum. By this we mean that it is possible to describe the
position of a point (at rest) by means of three numbers (co−ordinales) x, y, z, and that there is an
indefinite number of points in the neighbourhood of this one, the position of which can be described
by co−ordinates such as x
1
, y
1
, z
1
, which may be as near as we choose to the respective values of
the co−ordinates x, y, z, of the first point. In virtue of the latter property we speak of a " continuum,"
and owing to the fact that there are three co−ordinates we speak of it as being "
three−dimensional."
Similarly, the world of physical phenomena which was briefly called " world " by Minkowski is
naturally four dimensional in the space−time sense. For it is composed of individual events, each of
which is described by four numbers, namely, three space co−ordinates x, y, z, and a time
co−ordinate, the time value t. The" world" is in this sense also a continuum; for to every event there
are as many "neighbouring" events (realised or at least thinkable) as we care to choose, the
co−ordinates x
1
, y
1
, z
1
, t
1
of which differ by an indefinitely small amount from those of the event x,
y, z, t originally considered. That we have not been accustomed to regard the world in this sense as
a four−dimensional continuum is due to the fact that in physics, before the advent of the theory of
relativity, time played a different and more independent role, as compared with the space
coordinates. It is for this reason that we have been in the habit of treating time as an independent
continuum. As a matter of fact, according to classical mechanics, time is absolute, i.e. it is
independent of the position and the condition of motion of the system of co−ordinates. We see this
expressed in the last equation of the Galileian transformation (t
1
= t)
The four−dimensional mode of consideration of the "world" is natural on the theory of relativity,
since according to this theory time is robbed of its independence. This is shown by the fourth
equation of the Lorentz transformation:
Moreover, according to this equation the time difference ”t
1
of two events with respect to K
1
does
not in general vanish, even when the time difference ”t
1
of the same events with reference to
K vanishes. Pure " space−distance " of two events with respect to K results in " time−distance " of
the same events with respect to K. But the discovery of Minkowski, which was of importance for the
formal development of the theory of relativity, does not lie here. It is to be found rather in the fact of
his recognition that the four−dimensional space−time continuum of the theory of relativity, in its
most essential formal properties, shows a pronounced relationship to the three−dimensional
Relativity: The Special and General Theory
37
continuum of Euclidean geometrical space.
1)
In order to give due prominence to this relationship,
however, we must replace the usual time co−ordinate t by an imaginary magnitude
proportional to it. Under these conditions, the natural laws satisfying the demands of the
(special) theory of relativity assume mathematical forms, in which the time co−ordinate plays
exactly the same role as the three space co−ordinates. Formally, these four co−ordinates
correspond exactly to the three space co−ordinates in Euclidean geometry. It must be clear even to
the non−mathematician that, as a consequence of this purely formal addition to our knowledge, the
theory perforce gained clearness in no mean measure.
These inadequate remarks can give the reader only a vague notion of the important idea
contributed by Minkowski. Without it the general theory of relativity, of which the fundamental ideas
are developed in the following pages, would perhaps have got no farther than its long clothes.
Minkowski's work is doubtless difficult of access to anyone inexperienced in mathematics, but since
it is not necessary to have a very exact grasp of this work in order to understand the fundamental
ideas of either the special or the general theory of relativity, I shall leave it here at present, and
revert to it only towards the end of Part 2.
Next: Part II: The General Theory of Relativity
Footnotes
1)
Cf. the somewhat more detailed discussion in Appendix II.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
38
Albert Einstein: Relativity
Part II: The General Theory of Relativity
Part II
The General Theory of Relativity
Special and General Principle of Relativity
The basal principle, which was the pivot of all our previous considerations, was the special principle
of relativity, i.e. the principle of the physical relativity of all uniform motion. Let as once more
analyse its meaning carefully.
It was at all times clear that, from the point of view of the idea it conveys to us, every motion must
be considered only as a relative motion. Returning to the illustration we have frequently used of the
embankment and the railway carriage, we can express the fact of the motion here taking place in
the following two forms, both of which are equally justifiable :
(a) The carriage is in motion relative to the embankment,
(b) The embankment is in motion relative to the carriage.
In (a) the embankment, in (b) the carriage, serves as the body of reference in our statement of the
motion taking place. If it is simply a question of detecting or of describing the motion involved, it is
in principle immaterial to what reference−body we refer the motion. As already mentioned, this is
self−evident, but it must not be confused with the much more comprehensive statement called "the
principle of relativity," which we have taken as the basis of our investigations.
The principle we have made use of not only maintains that we may equally well choose the
carriage or the embankment as our reference−body for the description of any event (for this, too, is
self−evident). Our principle rather asserts what follows : If we formulate the general laws of nature
as they are obtained from experience, by making use of
(a) the embankment as reference−body,
(b) the railway carriage as reference−body,
then these general laws of nature (e.g. the laws of mechanics or the law of the propagation of light
in vacuo) have exactly the same form in both cases. This can also be expressed as follows : For
the physical description of natural processes, neither of the reference bodies K, K
1
is unique (lit. "
specially marked out ") as compared with the other. Unlike the first, this latter statement need not of
necessity hold a priori; it is not contained in the conceptions of " motion" and " reference−body "
and derivable from them; only experience can decide as to its correctness or incorrectness.
Up to the present, however, we have by no means maintained the equivalence of all bodies of
reference K in connection with the formulation of natural laws. Our course was more on the
following Iines. In the first place, we started out from the assumption that there exists a
reference−body K, whose condition of motion is such that the Galileian law holds with respect to it :
A particle left to itself and sufficiently far removed from all other particles moves uniformly in a
straight line. With reference to K (Galileian reference−body) the laws of nature were to be as simple
Relativity: The Special and General Theory
39
as possible. But in addition to K, all bodies of reference K
1
should be given preference in this
sense, and they should be exactly equivalent to K for the formulation of natural laws, provided that
they are in a state of uniform rectilinear and non−rotary motion with respect to K ; all these bodies
of reference are to be regarded as Galileian reference−bodies. The validity of the principle of
relativity was assumed only for these reference−bodies, but not for others (e.g. those possessing
motion of a different kind). In this sense we speak of the special principle of relativity, or special
theory of relativity.
In contrast to this we wish to understand by the "general principle of relativity" the following
statement : All bodies of reference K, K
1
, etc., are equivalent for the description of natural
phenomena (formulation of the general laws of nature), whatever may be their state of motion. But
before proceeding farther, it ought to be pointed out that this formulation must be replaced later by
a more abstract one, for reasons which will become evident at a later stage.
Since the introduction of the special principle of relativity has been justified, every intellect which
strives after generalisation must feel the temptation to venture the step towards the general
principle of relativity. But a simple and apparently quite reliable consideration seems to suggest
that, for the present at any rate, there is little hope of success in such an attempt; Let us imagine
ourselves transferred to our old friend the railway carriage, which is travelling at a uniform rate. As
long as it is moving unifromly, the occupant of the carriage is not sensible of its motion, and it is for
this reason that he can without reluctance interpret the facts of the case as indicating that the
carriage is at rest, but the embankment in motion. Moreover, according to the special principle of
relativity, this interpretation is quite justified also from a physical point of view.
If the motion of the carriage is now changed into a non−uniform motion, as for instance by a
powerful application of the brakes, then the occupant of the carriage experiences a correspondingly
powerful jerk forwards. The retarded motion is manifested in the mechanical behaviour of bodies
relative to the person in the railway carriage. The mechanical behaviour is different from that of the
case previously considered, and for this reason it would appear to be impossible that the same
mechanical laws hold relatively to the non−uniformly moving carriage, as hold with reference to the
carriage when at rest or in uniform motion. At all events it is clear that the Galileian law does not
hold with respect to the non−uniformly moving carriage. Because of this, we feel compelled at the
present juncture to grant a kind of absolute physical reality to non−uniform motion, in opposition to
the general principle of relatvity. But in what follows we shall soon see that this conclusion cannot
be maintained.
Next: The Gravitational Field
Relativity: The Special and General Theory
Relativity: The Special and General Theory
40
Albert Einstein: Relativity
Part II: The General Theory of Relativity
The Gravitational Field
"If we pick up a stone and then let it go, why does it fall to the ground ?" The usual answer to this
question is: "Because it is attracted by the earth." Modern physics formulates the answer rather
differently for the following reason. As a result of the more careful study of electromagnetic
phenomena, we have come to regard action at a distance as a process impossible without the
intervention of some intermediary medium. If, for instance, a magnet attracts a piece of iron, we
cannot be content to regard this as meaning that the magnet acts directly on the iron through the
intermediate empty space, but we are constrained to imagine — after the manner of Faraday —
that the magnet always calls into being something physically real in the space around it, that
something being what we call a "magnetic field." In its turn this magnetic field operates on the piece
of iron, so that the latter strives to move towards the magnet. We shall not discuss here the
justification for this incidental conception, which is indeed a somewhat arbitrary one. We shall only
mention that with its aid electromagnetic phenomena can be theoretically represented much more
satisfactorily than without it, and this applies particularly to the transmission of electromagnetic
waves. The effects of gravitation also are regarded in an analogous manner.
The action of the earth on the stone takes place indirectly. The earth produces in its surrounding a
gravitational field, which acts on the stone and produces its motion of fall. As we know from
experience, the intensity of the action on a body dimishes according to a quite definite law, as we
proceed farther and farther away from the earth. From our point of view this means : The law
governing the properties of the gravitational field in space must be a perfectly definite one, in order
correctly to represent the diminution of gravitational action with the distance from operative bodies.
It is something like this: The body (e.g. the earth) produces a field in its immediate neighbourhood
directly; the intensity and direction of the field at points farther removed from the body are thence
determined by the law which governs the properties in space of the gravitational fields themselves.
In contrast to electric and magnetic fields, the gravitational field exhibits a most remarkable
property, which is of fundamental importance for what follows. Bodies which are moving under the
sole influence of a gravitational field receive an acceleration, which does not in the least depend
either on the material or on the physical state of the body. For instance, a piece of lead and a piece
of wood fall in exactly the same manner in a gravitational field (in vacuo), when they start off from
rest or with the same initial velocity. This law, which holds most accurately, can be expressed in a
different form in the light of the following consideration.
According to Newton's law of motion, we have
(Force) = (inertial mass) x (acceleration),
where the "inertial mass" is a characteristic constant of the accelerated body. If now gravitation is
the cause of the acceleration, we then have
(Force) = (gravitational mass) x (intensity of the gravitational field),
where the "gravitational mass" is likewise a characteristic constant for the body. From these two
relations follows:
Relativity: The Special and General Theory
41
If now, as we find from experience, the acceleration is to be independent of the nature and the
condition of the body and always the same for a given gravitational field, then the ratio of the
gravitational to the inertial mass must likewise be the same for all bodies. By a suitable choice of
units we can thus make this ratio equal to unity. We then have the following law: The
gravitational mass of a body is equal to its inertial maw.
It is true that this important law had hitherto been recorded in mechanics, but it had not been
interpreted. A satisfactory interpretation can be obtained only if we recognise the following fact :
The same quality of a body manifests itself according to circumstances as " inertia " or as " weight "
(lit. " heaviness '). In the following section we shall show to what extent this is actually the case,
and how this question is connected with the general postulate of relativity.
Next: The Equality of Inertial and Gravitational Mass as an argument for the General Postule of
Relativity
Relativity: The Special and General Theory
Relativity: The Special and General Theory
42
Albert Einstein: Relativity
Part II: The General Theory of Relativity
The Equality of Inertial and Gravitational Mass
as an argument for the General Postule of Relativity
We imagine a large portion of empty space, so far removed from stars and other appreciable
masses, that we have before us approximately the conditions required by the fundamental law of
Galilei. It is then possible to choose a Galileian reference−body for this part of space (world),
relative to which points at rest remain at rest and points in motion continue permanently in uniform
rectilinear motion. As reference−body let us imagine a spacious chest resembling a room with an
observer inside who is equipped with apparatus. Gravitation naturally does not exist for this
observer. He must fasten himself with strings to the floor, otherwise the slightest impact against the
floor will cause him to rise slowly towards the ceiling of the room.
To the middle of the lid of the chest is fixed externally a hook with rope attached, and now a " being
" (what kind of a being is immaterial to us) begins pulling at this with a constant force. The chest
together with the observer then begin to move "upwards" with a uniformly accelerated motion. In
course of time their velocity will reach unheard−of values — provided that we are viewing all this
from another reference−body which is not being pulled with a rope.
But how does the man in the chest regard the Process ? The acceleration of the chest will be
transmitted to him by the reaction of the floor of the chest. He must therefore take up this pressure
by means of his legs if he does not wish to be laid out full length on the floor. He is then standing in
the chest in exactly the same way as anyone stands in a room of a home on our earth. If he
releases a body which he previously had in his land, the accelertion of the chest will no longer be
transmitted to this body, and for this reason the body will approach the floor of the chest with an
accelerated relative motion. The observer will further convince himself that the acceleration of the
body towards the floor of the chest is always of the same magnitude, whatever kind of body he may
happen to use for the experiment.
Relying on his knowledge of the gravitational field (as it was discussed in the preceding section),
the man in the chest will thus come to the conclusion that he and the chest are in a gravitational
field which is constant with regard to time. Of course he will be puzzled for a moment as to why the
chest does not fall in this gravitational field. just then, however, he discovers the hook in the middle
of the lid of the chest and the rope which is attached to it, and he consequently comes to the
conclusion that the chest is suspended at rest in the gravitational field.
Ought we to smile at the man and say that he errs in his conclusion ? I do not believe we ought to if
we wish to remain consistent ; we must rather admit that his mode of grasping the situation violates
neither reason nor known mechanical laws. Even though it is being accelerated with respect to the
"Galileian space" first considered, we can nevertheless regard the chest as being at rest. We have
thus good grounds for extending the principle of relativity to include bodies of reference which are
accelerated with respect to each other, and as a result we have gained a powerful argument for a
generalised postulate of relativity.
We must note carefully that the possibility of this mode of interpretation rests on the fundamental
property of the gravitational field of giving all bodies the same acceleration, or, what comes to the
same thing, on the law of the equality of inertial and gravitational mass. If this natural law did not
exist, the man in the accelerated chest would not be able to interpret the behaviour of the bodies
Relativity: The Special and General Theory
43
around him on the supposition of a gravitational field, and he would not be justified on the grounds
of experience in supposing his reference−body to be " at rest."
Suppose that the man in the chest fixes a rope to the inner side of the lid, and that he attaches a
body to the free end of the rope. The result of this will be to strech the rope so that it will hang "
vertically " downwards. If we ask for an opinion of the cause of tension in the rope, the man in the
chest will say: "The suspended body experiences a downward force in the gravitational field, and
this is neutralised by the tension of the rope ; what determines the magnitude of the tension of the
rope is the gravitational mass of the suspended body." On the other hand, an observer who is
poised freely in space will interpret the condition of things thus : " The rope must perforce take part
in the accelerated motion of the chest, and it transmits this motion to the body attached to it. The
tension of the rope is just large enough to effect the acceleration of the body. That which
determines the magnitude of the tension of the rope is the inertial mass of the body." Guided by this
example, we see that our extension of the principle of relativity implies the necessity of the law of
the equality of inertial and gravitational mass. Thus we have obtained a physical interpretation of
this law.
From our consideration of the accelerated chest we see that a general theory of relativity must yield
important results on the laws of gravitation. In point of fact, the systematic pursuit of the general
idea of relativity has supplied the laws satisfied by the gravitational field. Before proceeding farther,
however, I must warn the reader against a misconception suggested by these considerations. A
gravitational field exists for the man in the chest, despite the fact that there was no such field for the
co−ordinate system first chosen. Now we might easily suppose that the existence of a gravitational
field is always only an apparent one. We might also think that, regardless of the kind of gravitational
field which may be present, we could always choose another reference−body such that
no gravitational field exists with reference to it. This is by no means true for all gravitational fields,
but only for those of quite special form. It is, for instance, impossible to choose a body of reference
such that, as judged from it, the gravitational field of the earth (in its entirety) vanishes.
We can now appreciate why that argument is not convincing, which we brought forward against the
general principle of relativity at theend of Section 18. It is certainly true that the observer in the
railway carriage experiences a jerk forwards as a result of the application of the brake, and that he
recognises, in this the non−uniformity of motion (retardation) of the carriage. But he is compelled by
nobody to refer this jerk to a " real " acceleration (retardation) of the carriage. He might also
interpret his experience thus: " My body of reference (the carriage) remains permanently at rest.
With reference to it, however, there exists (during the period of application of the brakes) a
gravitational field which is directed forwards and which is variable with respect to time. Under the
influence of this field, the embankment together with the earth moves non−uniformly in such a
manner that their original velocity in the backwards direction is continuously reduced."
Next: In What Respects are the Foundations of Classical Mechanics and of the Special Theory of
Relativity Unsatisfactory?
Relativity: The Special and General Theory
Relativity: The Special and General Theory
44
Albert Einstein: Relativity
Part II: The General Theory of Relativity
In What Respects are the Foundations of Classical Mechanics and of the
Special Theory of Relativity Unsatisfactory?
We have already stated several times that classical mechanics starts out from the following law:
Material particles sufficiently far removed from other material particles continue to move uniformly
in a straight line or continue in a state of rest. We have also repeatedly emphasised that this
fundamental law can only be valid for bodies of reference K which possess certain unique states of
motion, and which are in uniform translational motion relative to each other. Relative to other
reference−bodies K the law is not valid. Both in classical mechanics and in the special theory of
relativity we therefore differentiate between reference−bodies K relative to which the recognised "
laws of nature " can be said to hold, and reference−bodies K relative to which these laws do not
hold.
But no person whose mode of thought is logical can rest satisfied with this condition of things. He
asks : " How does it come that certain reference−bodies (or their states of motion) are given priority
over other reference−bodies (or their states of motion) ? What is the reason for this Preference? In
order to show clearly what I mean by this question, I shall make use of a comparison.
I am standing in front of a gas range. Standing alongside of each other on the range are two pans
so much alike that one may be mistaken for the other. Both are half full of water. I notice that steam
is being emitted continuously from the one pan, but not from the other. I am surprised at this, even
if I have never seen either a gas range or a pan before. But if I now notice a luminous something of
bluish colour under the first pan but not under the other, I cease to be astonished, even if I have
never before seen a gas flame. For I can only say that this bluish something will cause the
emission of the steam, or at least possibly it may do so. If, however, I notice the bluish something in
neither case, and if I observe that the one continuously emits steam whilst the other does not, then
I shall remain astonished and dissatisfied until I have discovered some circumstance to which I can
attribute the different behaviour of the two pans.
Analogously, I seek in vain for a real something in classical mechanics (or in the special theory of
relativity) to which I can attribute the different behaviour of bodies considered with respect to the
reference systems K and K
1
.
1)
Newton saw this objection and attempted to invalidate it, but without
success. But E. Mach recognsed it most clearly of all, and because of this objection he claimed that
mechanics must be placed on a new basis. It can only be got rid of by means of a physics which is
conformable to the general principle of relativity, since the equations of such a theory hold for every
body of reference, whatever may be its state of motion.
Next: A Few Inferences from the General Principle of Relativity
Relativity: The Special and General Theory
45
Footnotes
1)
The objection is of importance more especially when the state of motion of the reference−body is
of such a nature that it does not require any external agency for its maintenance, e.g. in the case
when the reference−body is rotating uniformly.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
46
Albert Einstein: Relativity
Part II: The General Theory of Relativity
A Few Inferences from the General Principle of Relativity
The considerations of Section 20 show that the general principle of relativity puts us in a position to
derive properties of the gravitational field in a purely theoretical manner. Let us suppose, for
instance, that we know the space−time " course " for any natural process whatsoever, as regards
the manner in which it takes place in the Galileian domain relative to a Galileian body of reference
K. By means of purely theoretical operations (i.e. simply by calculation) we are then able to find
how this known natural process appears, as seen from a reference−body K
1
which is accelerated
relatively to K. But since a gravitational field exists with respect to this new body of reference K
1
,
our consideration also teaches us how the gravitational field influences the process studied.
For example, we Wayrn that a body which is in a state of uniform rectilinear motion with respect to
K (in accordance with the law of Galilei) is executing an accelerated and in general curvilinear
motion with respect to the accelerated reference−body K
1
(chest). This acceleration or curvature
corresponds to the influence on the moving body of the gravitational field prevailing relatively to
K. It is known that a gravitational field influences the movement of bodies in this way, so that our
consideration supplies us with nothing essentially new.
However, we obtain a new result of fundamental importance when we carry out the analogous
consideration for a ray of light. With respect to the Galileian reference−body K, such a ray of light is
transmitted rectilinearly with the velocity c. It can easily be shown that the path of the same ray of
light is no longer a straight line when we consider it with reference to the accelerated chest
(reference−body K
1
). From this we conclude, that, in general, rays of light are propagated
curvilinearly in gravitational fields. In two respects this result is of great importance.
In the first place, it can be compared with the reality. Although a detailed examination of the
question shows that the curvature of light rays required by the genernal theory of relativity is only
exceedingly small for the gravitational fields at our disposal in practice, its estimated magnitude for
light rays passing the sun at grazing incidence is nevertheless 1.7 seconds of arc. This ought to
manifest itself in the following way. As seen from the earth, certain fixed stars appear to be in the
neighbourhood of the sun, and are thus capable of observation during a total eclipse of the sun. At
such times, these stars ought to appear to be displaced outwards from the sun by an amount
indicated above, as compared with their apparent position in the sky when the sun is situated at
another part of the heavens. The examination of the correctness or otherwise of this deduction is a
problem of the greatest importance, the early solution of which is to be expected of astronomers.
1)
In the second place our result shows that, according to the general theory of relativity, the law of
the constancy of the velocity of light in vacuo, which constitutes one of the two fundamental
assumptions in the special theory of relativity and to which we have already frequently referred,
cannot claim any unlimited validity. A curvature of rays of light can only take place when the
velocity of propagation of light varies with position. Now we might think that as a consequence of
this, the special theory of relativity and with it the whole theory of relativity would be laid in the dust.
But in reality this is not the case. We can only conclude that the special theory of relativity cannot
claim an unlinlited domain of validity ; its results hold only so long as we are able to disregard the
influences of gravitational fields on the phenomena (e.g. of light).
Relativity: The Special and General Theory
47
Since it has often been contended by opponents of the theory of relativity that the special theory of
relativity is overthrown by the general theory of relativity, it is perhaps advisable to make the facts
of the case clearer by means of an appropriate comparison. Before the development of
electrodynamics the laws of electrostatics were looked upon as the laws of electricity. At the
present time we know that electric fields can be derived correctly from electrostatic considerations
only for the case, which is never strictly realised, in which the electrical masses are quite at rest
relatively to each other, and to the co−ordinate system. Should we be justified in saying that for this
reason electrostatics is overthrown by the field−equations of Maxwell in electrodynamics ? Not in
the least. Electrostatics is contained in electrodynamics as a limiting case ; the laws of the latter
lead directly to those of the former for the case in which the fields are invariable with regard to time.
No fairer destiny could be allotted to any physical theory, than that it should of itself point out the
way to the introduction of a more comprehensive theory, in which it lives on as a limiting case.
In the example of the transmission of light just dealt with, we have seen that the general theory of
relativity enables us to derive theoretically the influence of a gravitational field on the course of
natural processes, the Iaws of which are already known when a gravitational field is absent. But the
most attractive problem, to the solution of which the general theory of relativity supplies the key,
concerns the investigation of the laws satisfied by the gravitational field itself. Let us consider this
for a moment.
We are acquainted with space−time domains which behave (approximately) in a " Galileian "
fashion under suitable choice of reference−body, i.e. domains in which gravitational fields are
absent. If we now refer such a domain to a reference−body K
1
possessing any kind of motion, then
relative to K
1
there exists a gravitational field which is variable with respect to space and time.
2)
The
character of this field will of course depend on the motion chosen for K
1
. According to the general
theory of relativity, the general law of the gravitational field must be satisfied for all gravitational
fields obtainable in this way. Even though by no means all gravitationial fields can be produced in
this way, yet we may entertain the hope that the general law of gravitation will be derivable from
such gravitational fields of a special kind. This hope has been realised in the most beautiful
manner. But between the clear vision of this goal and its actual realisation it was necessary to
surmount a serious difficulty, and as this lies deep at the root of things, I dare not withhold it from
the reader. We require to extend our ideas of the space−time continuum still farther.
Next: Behaviour of Clocks and Measuring−Rods on a Rotating Body of Reference
Footnotes
1)
By means of the star photographs of two expeditions equipped by a Joint Committee of the Royal
and Royal Astronomical Societies, the existence of the deflection of light demanded by theory was
first confirmed during the solar eclipse of 29th May, 1919. (Cf. Appendix III.)
2)
This follows from a generalisation of the discussion in Section 20
Relativity: The Special and General Theory
Relativity: The Special and General Theory
48
Albert Einstein: Relativity
Part II: The General Theory of Relativity
Behaviour of Clocks and Measuring−Rods on a Rotating Body of Reference
Hitherto I have purposely refrained from speaking about the physical interpretation of space− and
time−data in the case of the general theory of relativity. As a consequence, I am guilty of a certain
slovenliness of treatment, which, as we know from the special theory of relativity, is far from being
unimportant and pardonable. It is now high time that we remedy this defect; but I would mention at
the outset, that this matter lays no small claims on the patience and on the power of abstraction of
the reader.
We start off again from quite special cases, which we have frequently used before. Let us consider
a space time domain in which no gravitational field exists relative to a reference−body K whose
state of motion has been suitably chosen. K is then a Galileian reference−body as regards the
domain considered, and the results of the special theory of relativity hold relative to K. Let us
supposse the same domain referred to a second body of reference K
1
, which is rotating uniformly
with respect to K. In order to fix our ideas, we shall imagine K
1
to be in the form of a plane circular
disc, which rotates uniformly in its own plane about its centre. An observer who is sitting
eccentrically on the disc K
1
is sensible of a force which acts outwards in a radial direction, and
which would be interpreted as an effect of inertia (centrifugal force) by an observer who was at rest
with respect to the original reference−body K. But the observer on the disc may regard his disc as a
reference−body which is " at rest " ; on the basis of the general principle of relativity he is justified in
doing this. The force acting on himself, and in fact on all other bodies which are at rest relative to
the disc, he regards as the effect of a gravitational field. Nevertheless, the space−distribution of this
gravitational field is of a kind that would not be possible on Newton's theory of gravitation.
1)
But
since the observer believes in the general theory of relativity, this does not disturb him; he is quite
in the right when he believes that a general law of gravitation can be formulated− a law which not
only explains the motion of the stars correctly, but also the field of force experienced by himself.
The observer performs experiments on his circular disc with clocks and measuring−rods. In doing
so, it is his intention to arrive at exact definitions for the signification of time− and space−data with
reference to the circular disc K
1
, these definitions being based on his observations. What will be his
experience in this enterprise ?
To start with, he places one of two identically constructed clocks at the centre of the circular disc,
and the other on the edge of the disc, so that they are at rest relative to it. We now ask ourselves
whether both clocks go at the same rate from the standpoint of the non−rotating Galileian
reference−body K. As judged from this body, the clock at the centre of the disc has no velocity,
whereas the clock at the edge of the disc is in motion relative to K in consequence of the rotation.
According to a result obtained in Section 12, it follows that the latter clock goes at a rate
permanently slower than that of the clock at the centre of the circular disc, i.e. as observed from K.
It is obvious that the same effect would be noted by an observer whom we will imagine sitting
alongside his clock at the centre of the circular disc. Thus on our circular disc, or, to make the case
more general, in every gravitational field, a clock will go more quickly or less quickly, according to
the position in which the clock is situated (at rest). For this reason it is not possible to obtain a
reasonable definition of time with the aid of clocks which are arranged at rest with respect to the
body of reference. A similar difficulty presents itself when we attempt to apply our earlier definition
of simultaneity in such a case, but I do not wish to go any farther into this question.
Relativity: The Special and General Theory
49
Moreover, at this stage the definition of the space co−ordinates also presents insurmountable
difficulties. If the observer applies his standard measuring−rod (a rod which is short as compared
with the radius of the disc) tangentially to the edge of the disc, then, as judged from the Galileian
system, the length of this rod will be less than I, since, according to Section 12, moving bodies
suffer a shortening in the direction of the motion. On the other hand, the measaring−rod will not
experience a shortening in length, as judged from K, if it is applied to the disc in the direction of the
radius. If, then, the observer first measures the circumference of the disc with his measuring−rod
and then the diameter of the disc, on dividing the one by the other, he will not obtain as quotient the
familiar number À = 3.14 . . ., but a larger number,
2)
whereas of course, for a disc which is at rest
with respect to K, this operation would yield À exactly. This proves that the propositions of
Euclidean geometry cannot hold exactly on the rotating disc, nor in general in a gravitational field,
at least if we attribute the length I to the rod in all positions and in every orientation. Hence the idea
of a straight line also loses its meaning. We are therefore not in a position to define exactly the
co−ordinates x, y, z relative to the disc by means of the method used in discussing the special
theory, and as long as the co− ordinates and times of events have not been defined, we cannot
assign an exact meaning to the natural laws in which these occur.
Thus all our previous conclusions based on general relativity would appear to be called in question.
In reality we must make a subtle detour in order to be able to apply the postulate of general
relativity exactly. I shall prepare the reader for this in the following paragraphs.
Next: Euclidean and Non−Euclidean Continuum
Footnotes
1)
The field disappears at the centre of the disc and increases proportionally to the distance from
the centre as we proceed outwards.
2)
Throughout this consideration we have to use the Galileian (non−rotating) system K as
reference−body, since we may only assume the validity of the results of the special theory of
relativity relative to K (relative to K
1
a gravitational field prevails).
Relativity: The Special and General Theory
Relativity: The Special and General Theory
50
Albert Einstein: Relativity
Part II: The General Theory of Relativity
Euclidean and Non−Euclidean Continuum
The surface of a marble table is spread out in front of me. I can get from any one point on this table
to any other point by passing continuously from one point to a " neighbouring " one, and repeating
this process a (large) number of times, or, in other words, by going from point to point without
executing "jumps." I am sure the reader will appreciate with sufficient clearness what I mean here
by " neighbouring " and by " jumps " (if he is not too pedantic). We express this property of the
surface by describing the latter as a continuum.
Let us now imagine that a large number of little rods of equal length have been made, their lengths
being small compared with the dimensions of the marble slab. When I say they are of equal length,
I mean that one can be laid on any other without the ends overlapping. We next lay four of these
little rods on the marble slab so that they constitute a quadrilateral figure (a square), the diagonals
of which are equally long. To ensure the equality of the diagonals, we make use of a little
testing−rod. To this square we add similar ones, each of which has one rod in common with the
first. We proceed in like manner with each of these squares until finally the whole marble slab is
laid out with squares. The arrangement is such, that each side of a square belongs to two squares
and each corner to four squares.
It is a veritable wander that we can carry out this business without getting into the greatest
difficulties. We only need to think of the following. If at any moment three squares meet at a corner,
then two sides of the fourth square are already laid, and, as a consequence, the arrangement of
the remaining two sides of the square is already completely determined. But I am now no longer
able to adjust the quadrilateral so that its diagonals may be equal. If they are equal of their own
accord, then this is an especial favour of the marble slab and of the little rods, about which I can
only be thankfully surprised. We must experience many such surprises if the construction is to be
successful.
If everything has really gone smoothly, then I say that the points of the marble slab constitute a
Euclidean continuum with respect to the little rod, which has been used as a " distance "
(line−interval). By choosing one corner of a square as " origin" I can characterise every other corner
of a square with reference to this origin by means of two numbers. I only need state how many rods
I must pass over when, starting from the origin, I proceed towards the " right " and then " upwards,"
in order to arrive at the corner of the square under consideration. These two numbers are then the "
Cartesian co−ordinates " of this corner with reference to the " Cartesian co−ordinate system" which
is determined by the arrangement of little rods.
By making use of the following modification of this abstract experiment, we recognise that there
must also be cases in which the experiment would be unsuccessful. We shall suppose that the rods
" expand " by in amount proportional to the increase of temperature. We heat the central part of the
marble slab, but not the periphery, in which case two of our little rods can still be brought into
coincidence at every position on the table. But our construction of squares must necessarily come
into disorder during the heating, because the little rods on the central region of the table expand,
whereas those on the outer part do not.
With reference to our little rods — defined as unit lengths — the marble slab is no longer a
Euclidean continuum, and we are also no longer in the position of defining Cartesian co−ordinates
Relativity: The Special and General Theory
51
directly with their aid, since the above construction can no longer be carried out. But since there are
other things which are not influenced in a similar manner to the little rods (or perhaps not at all) by
the temperature of the table, it is possible quite naturally to maintain the point of view that the
marble slab is a " Euclidean continuum." This can be done in a satisfactory manner by making a
more subtle stipulation about the measurement or the comparison of lengths.
But if rods of every kind (i.e. of every material) were to behave in the same way as regards the
influence of temperature when they are on the variably heated marble slab, and if we had no other
means of detecting the effect of temperature than the geometrical behaviour of our rods in
experiments analogous to the one described above, then our best plan would be to assign the
distance one to two points on the slab, provided that the ends of one of our rods could be made to
coincide with these two points ; for how else should we define the distance without our proceeding
being in the highest measure grossly arbitrary ? The method of Cartesian coordinates must then be
discarded, and replaced by another which does not assume the validity of Euclidean geometry for
rigid bodies.
1)
The reader will notice that the situation depicted here corresponds to the one
brought about by the general postitlate of relativity (Section 23).
Next: Gaussian Co−ordinates
Footnotes
1)
Mathematicians have been confronted with our problem in the following form. If we are given a
surface (e.g. an ellipsoid) in Euclidean three−dimensional space, then there exists for this surface a
two−dimensional geometry, just as much as for a plane surface. Gauss undertook the task of
treating this two−dimensional geometry from first principles, without making use of the fact that the
surface belongs to a Euclidean continuum of three dimensions. If we imagine constructions to be
made with rigid rods in the surface (similar to that above with the marble slab), we should find that
different laws hold for these from those resulting on the basis of Euclidean plane geometry. The
surface is not a Euclidean continuum with respect to the rods, and we cannot define Cartesian
co−ordinates in the surface. Gauss indicated the principles according to which we can treat the
geometrical relationships in the surface, and thus pointed out the way to the method of Ricmman of
treating multi−dimensional, non−Euclidean continuum. Thus it is that mathematicians long ago
solved the formal problems to which we are led by the general postulate of relativity.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
52
Albert Einstein: Relativity
Part II: The General Theory of Relativity
Gaussian Co−ordinates
According to Gauss, this combined analytical and geometrical mode of handling the problem can
be arrived at in the following way. We imagine a system of arbitrary curves (see Fig. 4) drawn on
the surface of the table. These we designate as u−curves, and we indicate each of them by means
of a number. The Curves u= 1, u= 2 and u= 3 are drawn in the diagram. Between the curves u= 1
and u= 2 we must imagine an infinitely large number to be drawn, all of which correspond to real
numbers lying between 1 and 2. We have then a system of
u−curves, and this "infinitely dense" system covers the whole surface of the table. These u−curves
must not interseect each other, and through each point of the surface one and only one curve must
pass. Thus a perfectly definite value of u belongs to every point on the surface of the marble slab.
In like manner we imagine a system of v−curves drawn on the surface. These satisfy the same
conditions as the u−curves, they are provided with numbers in a corresponding manner, and they
may likewise be of arbitrary shape. It follows that a value of u and a value of v belong to every point
on the surface of the table. We call these two numbers the co−ordinates of the surface of the table
(Gaussian co−ordinates). For example, the point P in the diagram has the Gaussian co−ordinates
u= 3, v= 1. Two neighbouring points P and P
1
on the surface then correspond to the co−ordinates
P: u,v
P
1
: u + du, v + dv,
where du and dv signify very small numbers. In a similar manner we may indicate the distance
(line−interval) between P and P
1
, as measured with a little rod, by means of the very small number
ds. Then according to Gauss we have
ds
2
= g
11
du
2
+ 2g
12
dudv = g
22
dv
2
where g
11
, g
12
, g
22
, are magnitudes which depend in a perfectly definite way on u and v. The
magnitudes g
11
, g
12
and g
22
, determine the behaviour of the rods relative to the u−curves and
v−curves, and thus also relative to the surface of the table. For the case in which the points of the
surface considered form a Euclidean continuum with reference to the measuring−rods, but only in
this case, it is possible to draw the u−curves and v−curves and to attach numbers to them, in such
a manner, that we simply have :
ds
2
= du
2
+ dv
2
Relativity: The Special and General Theory
53
Under these conditions, the u−curves and v−curves are straight lines in the sense of Euclidean
geometry, and they are perpendicular to each other. Here the Gaussian coordinates are samply
Cartesian ones. It is clear that Gauss co−ordinates are nothing more than an association of two
sets of numbers with the points of the surface considered, of such a nature that numerical values
differing very slightly from each other are associated with neighbouring points " in space."
So far, these considerations hold for a continuum of two dimensions. But the Gaussian method can
be applied also to a continuum of three, four or more dimensions. If, for instance, a continuum of
four dimensions be supposed available, we may represent it in the following way. With every point
of the continuum, we associate arbitrarily four numbers, x
1
, x
2
, x
3
, x
4
, which are known as "
co−ordinates." Adjacent points correspond to adjacent values of the coordinates. If a distance ds is
associated with the adjacent points P and P
1
, this distance being measurable and well defined from
a physical point of view, then the following formula holds:
ds
2
= g
11
dx
1
2
+ 2g
12
dx
1
dx
2
. . . . g
44
dx
4
2
,
where the magnitudes g
11
, etc., have values which vary with the position in the continuum. Only
when the continuum is a Euclidean one is it possible to associate the co−ordinates x
1
. . x
4
. with the
points of the continuum so that we have simply
ds
2
= dx
1
2
+ dx
2
2
+ dx
3
2
+ dx
4
2
.
In this case relations hold in the four−dimensional continuum which are analogous to those holding
in our three−dimensional measurements.
However, the Gauss treatment for ds
2
which we have given above is not always possible. It is only
possible when sufficiently small regions of the continuum under consideration may be regarded as
Euclidean continua. For example, this obviously holds in the case of the marble slab of the table
and local variation of temperature. The temperature is practically constant for a small part of the
slab, and thus the geometrical behaviour of the rods is almost as it ought to be according to the
rules of Euclidean geometry. Hence the imperfections of the construction of squares in the previous
section do not show themselves clearly until this construction is extended over a considerable
portion of the surface of the table.
We can sum this up as follows: Gauss invented a method for the mathematical treatment of
continua in general, in which " size−relations " (" distances " between neighbouring points) are
defined. To every point of a continuum are assigned as many numbers (Gaussian coordinates) as
the continuum has dimensions. This is done in such a way, that only one meaning can be attached
to the assignment, and that numbers (Gaussian coordinates) which differ by an indefinitely small
amount are assigned to adjacent points. The Gaussian coordinate system is a logical
generalisation of the Cartesian co−ordinate system. It is also applicable to non−Euclidean continua,
but only when, with respect to the defined "size" or "distance," small parts of the continuum under
consideration behave more nearly like a Euclidean system, the smaller the part of the continuum
under our notice.
Next: The Space−Time Continuum of the Speical Theory of Relativity Considered as a Euclidean
Continuum
Relativity: The Special and General Theory
Relativity: The Special and General Theory
54
Albert Einstein: Relativity
Part II: The General Theory of Relativity
The Space−Time Continuum of the Speical Theory of Relativity Considered
as a Euclidean Continuum
We are now in a position to formulate more exactly the idea of Minkowski, which was only vaguely
indicated in Section 17. In accordance with the special theory of relativity, certain co−ordinate
systems are given preference for the description of the four−dimensional, space−time continuum.
We called these " Galileian co−ordinate systems." For these systems, the four co−ordinates x, y, z,
t, which determine an event or — in other words, a point of the four−dimensional continuum — are
defined physically in a simple manner, as set forth in detail in the first part of this book. For the
transition from one Galileian system to another, which is moving uniformly with reference to the
first, the equations of the Lorentz transformation are valid. These last form the basis for the
derivation of deductions from the special theory of relativity, and in themselves they are nothing
more than the expression of the universal validity of the law of transmission of light for all Galileian
systems of reference.
Minkowski found that the Lorentz transformations satisfy the following simple conditions. Let us
consider two neighbouring events, the relative position of which in the four−dimensional continuum
is given with respect to a Galileian reference−body K by the space co−ordinate differences dx, dy,
dz and the time−difference dt. With reference to a second Galileian system we shall suppose that
the corresponding differences for these two events are dx
1
, dy
1
, dz
1
, dt
1
. Then these magnitudes
always fulfil the condition
1)
dx
2
+ dy
2
+ dz
2
− c
2
dt
2
= dx
1 2
+ dy
1 2
+ dz
1 2
− c
2
dt
1 2
.
The validity of the Lorentz transformation follows from this condition. We can express this as
follows: The magnitude
ds
2
= dx
2
+ dy
2
+ dz
2
− c
2
dt
2
,
which belongs to two adjacent points of the four−dimensional space−time continuum, has the same
value for all selected (Galileian) reference−bodies. If we replace x, y, z, , by x
1
, x
2
, x
3
, x
4
, we
also obtaill the result that
ds
2
= dx
1
2
+ dx
2
2
+ dx
3
2
+ dx
4
2
.
is independent of the choice of the body of reference. We call the magnitude ds the " distance "
apart of the two events or four−dimensional points.
Thus, if we choose as time−variable the imaginary variable instead of the real quantity t, we
can regard the space−time contintium — accordance with the special theory of relativity — as a ",
Euclidean " four−dimensional continuum, a result which follows from the considerations of the
preceding section.
Next: The Space−Time Continuum of the General Theory of Realtiivty is Not a Eculidean Continuum
Relativity: The Special and General Theory
55
Footnotes
1)
Cf. Appendixes I and 2. The relations which are derived there for the co−ordlnates themselves
are valid also for co−ordinate differences, and thus also for co−ordinate differentials (indefilnitely
small differences).
Relativity: The Special and General Theory
Relativity: The Special and General Theory
56
Albert Einstein: Relativity
Part II: The General Theory of Relativity
The Space−Time Continuum of the General Theory of Realtiivty is Not a
Eculidean Continuum
In the first part of this book we were able to make use of space−time co−ordinates which allowed of
a simple and direct physical interpretation, and which, according to Section 26, can be regarded as
four−dimensional Cartesian co−ordinates. This was possible on the basis of the law of the
constancy of the velocity of tight. But according to Section 21 the general theory of relativity cannot
retain this law. On the contrary, we arrived at the result that according to this latter theory the
velocity of light must always depend on the co−ordinates when a gravitational field is present. In
connection with a specific illustration in Section 23, we found that the presence of a gravitational
field invalidates the definition of the coordinates and the ifine, which led us to our objective in the
special theory of relativity.
In view of the resuIts of these considerations we are led to the conviction that, according to the
general principle of relativity, the space−time continuum cannot be regarded as a Euclidean one,
but that here we have the general case, corresponding to the marble slab with local variations of
temperature, and with which we made acquaintance as an example of a two−dimensional
continuum. Just as it was there impossible to construct a Cartesian co−ordinate system from equal
rods, so here it is impossible to build up a system (reference−body) from rigid bodies and clocks,
which shall be of such a nature that measuring−rods and clocks, arranged rigidly with respect to
one another, shaIll indicate position and time directly. Such was the essence of the difficulty with
which we were confronted in Section 23.
But the considerations of Sections 25 and 26 show us the way to surmount this difficulty. We refer
the fourdimensional space−time continuum in an arbitrary manner to Gauss co−ordinates. We
assign to every point of the continuum (event) four numbers, x
1
, x
2
, x
3
, x
4
(co−ordinates), which
have not the least direct physical significance, but only serve the purpose of numbering the points
of the continuum in a definite but arbitrary manner. This arrangement does not even need to be of
such a kind that we must regard x
1
, x
2
, x
3
, as "space" co−ordinates and x
4
, as a " time "
co−ordinate.
The reader may think that such a description of the world would be quite inadequate. What does it
mean to assign to an event the particular co−ordinates x
1
, x
2
, x
3
, x
4
, if in themselves these
co−ordinates have no significance ? More careful consideration shows, however, that this anxiety is
unfounded. Let us consider, for instance, a material point with any kind of motion. If this point had
only a momentary existence without duration, then it would to described in space−time by a single
system of values x
1
, x
2
, x
3
, x
4
. Thus its permanent existence must be characterised by an infinitely
large number of such systems of values, the co−ordinate values of which are so close together as
to give continuity; corresponding to the material point, we thus have a (uni−dimensional) line in the
four−dimensional continuum. In the same way, any such lines in our continuum correspond to
many points in motion. The only statements having regard to these points which can claim a
physical existence are in reality the statements about their encounters. In our mathematical
treatment, such an encounter is expressed in the fact that the two lines which represent the
motions of the points in question have a particular system of co−ordinate values, x
1
, x
2
, x
3
, x
4
, in
common. After mature consideration the reader will doubtless admit that in reality such encounters
constitute the only actual evidence of a time−space nature with which we meet in physical
statements.
Relativity: The Special and General Theory
57
When we were describing the motion of a material point relative to a body of reference, we stated
nothing more than the encounters of this point with particular points of the reference−body. We can
also determine the corresponding values of the time by the observation of encounters of the body
with clocks, in conjunction with the observation of the encounter of the hands of clocks with
particular points on the dials. It is just the same in the case of space−measurements by means of
measuring−rods, as a litttle consideration will show.
The following statements hold generally : Every physical description resolves itself into a number of
statements, each of which refers to the space−time coincidence of two events A and B. In terms of
Gaussian co−ordinates, every such statement is expressed by the agreement of their four
co−ordinates x
1
, x
2
, x
3
, x
4
. Thus in reality, the description of the time−space continuum by means of
Gauss co−ordinates completely replaces the description with the aid of a body of reference, without
suffering from the defects of the latter mode of description; it is not tied down to the Euclidean
character of the continuum which has to be represented.
Next: Exact Formulation of the General Principle of Relativity
Relativity: The Special and General Theory
Relativity: The Special and General Theory
58
Albert Einstein: Relativity
Part II: The General Theory of Relativity
Exact Formulation of the General Principle of Relativity
We are now in a position to replace the pro. visional formulation of the general principle of relativity
given in Section 18 by an exact formulation. The form there used, "All bodies of reference K, K
1
,
etc., are equivalent for the description of natural phenomena (formulation of the general laws of
nature), whatever may be their state of motion," cannot be maintained, because the use of rigid
reference−bodies, in the sense of the method followed in the special theory of relativity, is in
general not possible in space−time description. The Gauss co−ordinate system has to take the
place of the body of reference. The following statement corresponds to the fundamental idea of the
general principle of relativity: "All Gaussian co−ordinate systems are essentially equivalent for the
formulation of the general laws of nature."
We can state this general principle of relativity in still another form, which renders it yet more clearly
intelligible than it is when in the form of the natural extension of the special principle of relativity.
According to the special theory of relativity, the equations which express the general laws of nature
pass over into equations of the same form when, by making use of the Lorentz transformation, we
replace the space−time variables x, y, z, t, of a (Galileian) reference−body K by the space−time
variables x
1
, y
1
, z
1
, t
1
, of a new reference−body K
1
. According to the general theory of relativity, on
the other hand, by application of arbitrary substitutions of the Gauss variables x
1
, x
2
, x
3
, x
4
, the
equations must pass over into equations of the same form; for every transformation (not only the
Lorentz transformation) corresponds to the transition of one Gauss co−ordinate system into
another.
If we desire to adhere to our "old−time" three−dimensional view of things, then we can characterise
the development which is being undergone by the fundamental idea of the general theory of
relativity as follows : The special theory of relativity has reference to Galileian domains, i.e. to those
in which no gravitational field exists. In this connection a Galileian reference−body serves as body
of reference, i.e. a rigid body the state of motion of which is so chosen that the Galileian law of the
uniform rectilinear motion of "isolated" material points holds relatively to it.
Certain considerations suggest that we should refer the same Galileian domains to
non−Galileian reference−bodies also. A gravitational field of a special kind is then present with
respect to these bodies (cf. Sections 20 and 23).
In gravitational fields there are no such things as rigid bodies with Euclidean properties; thus the
fictitious rigid body of reference is of no avail in the general theory of relativity. The motion of clocks
is also influenced by gravitational fields, and in such a way that a physical definition of time which is
made directly with the aid of clocks has by no means the same degree of plausibility as in the
special theory of relativity.
For this reason non−rigid reference−bodies are used, which are as a whole not only moving in any
way whatsoever, but which also suffer alterations in form ad lib. during their motion. Clocks, for
which the law of motion is of any kind, however irregular, serve for the definition of time. We have
to imagine each of these clocks fixed at a point on the non−rigid reference−body. These clocks
satisfy only the one condition, that the "readings" which are observed simultaneously on adjacent
clocks (in space) differ from each other by an indefinitely small amount. This non−rigid
Relativity: The Special and General Theory
59
reference−body, which might appropriately be termed a "reference−mollusc", is in the main
equivalent to a Gaussian four−dimensional co−ordinate system chosen arbitrarily. That which gives
the "mollusc" a certain comprehensibility as compared with the Gauss co−ordinate system is the
(really unjustified) formal retention of the separate existence of the space co−ordinates as opposed
to the time co−ordinate. Every point on the mollusc is treated as a space−point, and every material
point which is at rest relatively to it as at rest, so long as the mollusc is considered as
reference−body. The general principle of relativity requires that all these molluscs can be used as
reference−bodies with equal right and equal success in the formulation of the general laws of
nature; the laws themselves must be quite independent of the choice of mollusc.
The great power possessed by the general principle of relativity lies in the comprehensive limitation
which is imposed on the laws of nature in consequence of what we have seen above.
Next: The Solution of the Problem of Gravitation on the Basis of the General Principle of Relativity
Relativity: The Special and General Theory
Relativity: The Special and General Theory
60
Albert Einstein: Relativity
Part II: The General Theory of Relativity
The Solution of the Problem of Gravitation on the Basis of the General
Principle of Relativity
If the reader has followed all our previous considerations, he will have no further difficulty in
understanding the methods leading to the solution of the problem of gravitation.
We start off on a consideration of a Galileian domain, i.e. a domain in which there is no
gravitational field relative to the Galileian reference−body K. The behaviour of measuring−rods and
clocks with reference to K is known from the special theory of relativity, likewise the behaviour of
"isolated" material points; the latter move uniformly and in straight lines.
Now let us refer this domain to a random Gauss coordinate system or to a "mollusc" as
reference−body K
1
. Then with respect to K
1
there is a gravitational field G (of a particular kind). We
learn the behaviour of measuring−rods and clocks and also of freely−moving material points with
reference to K
1
simply by mathematical transformation. We interpret this behaviour as the
behaviour of measuring−rods, docks and material points tinder the influence of the gravitational
field G. Hereupon we introduce a hypothesis: that the influence of the gravitational field on
measuringrods, clocks and freely−moving material points continues to take place according to the
same laws, even in the case where the prevailing gravitational field is not derivable from the
Galfleian special care, simply by means of a transformation of co−ordinates.
The next step is to investigate the space−time behaviour of the gravitational field G, which was
derived from the Galileian special case simply by transformation of the coordinates. This behaviour
is formulated in a law, which is always valid, no matter how the reference−body (mollusc) used in
the description may be chosen.
This law is not yet the general law of the gravitational field, since the gravitational field under
consideration is of a special kind. In order to find out the general law−of−field of gravitation we still
require to obtain a generalisation of the law as found above. This can be obtained without caprice,
however, by taking into consideration the following demands:
(a) The required generalisation must likewise satisfy the general postulate of relativity.
(b) If there is any matter in the domain under consideration, only its inertial mass, and thus
according to Section 15 only its energy is of importance for its etfect in exciting a field.
(c) Gravitational field and matter together must satisfy the law of the conservation of energy (and of
impulse).
Finally, the general principle of relativity permits us to determine the influence of the gravitational
field on the course of all those processes which take place according to known laws when a
gravitational field is absent i.e. which have already been fitted into the frame of the special theory of
relativity. In this connection we proceed in principle according to the method which has already
been explained for measuring−rods, clocks and freely moving material points.
The theory of gravitation derived in this way from the general postulate of relativity excels not only
in its beauty ; nor in removing the defect attaching to classical mechanics which was brought to
Relativity: The Special and General Theory
61
light in Section 21; nor in interpreting the empirical law of the equality of inertial and gravitational
mass ; but it has also already explained a result of observation in astronomy, against which
classical mechanics is powerless.
If we confine the application of the theory to the case where the gravitational fields can be regarded
as being weak, and in which all masses move with respect to the coordinate system with velocities
which are small compared with the velocity of light, we then obtain as a first approximation the
Newtonian theory. Thus the latter theory is obtained here without any particular assumption,
whereas Newton had to introduce the hypothesis that the force of attraction between mutually
attracting material points is inversely proportional to the square of the distance between them. If we
increase the accuracy of the calculation, deviations from the theory of Newton make their
appearance, practically all of which must nevertheless escape the test of observation owing to their
smallness.
We must draw attention here to one of these deviations. According to Newton's theory, a planet
moves round the sun in an ellipse, which would permanently maintain its position with respect to
the fixed stars, if we could disregard the motion of the fixed stars themselves and the action of the
other planets under consideration. Thus, if we correct the observed motion of the planets for these
two influences, and if Newton's theory be strictly correct, we ought to obtain for the orbit of the
planet an ellipse, which is fixed with reference to the fixed stars. This deduction, which can be
tested with great accuracy, has been confirmed for all the planets save one, with the precision that
is capable of being obtained by the delicacy of observation attainable at the present time. The sole
exception is Mercury, the planet which lies nearest the sun. Since the time of Leverrier, it has been
known that the ellipse corresponding to the orbit of Mercury, after it has been corrected for the
influences mentioned above, is not stationary with respect to the fixed stars, but that it rotates
exceedingly slowly in the plane of the orbit and in the sense of the orbital motion. The value
obtained for this rotary movement of the orbital ellipse was 43 seconds of arc per century, an
amount ensured to be correct to within a few seconds of arc. This effect can be explained by
means of classical mechanics only on the assumption of hypotheses which have little probability,
and which were devised solely for this purponse.
On the basis of the general theory of relativity, it is found that the ellipse of every planet round the
sun must necessarily rotate in the manner indicated above ; that for all the planets, with the
exception of Mercury, this rotation is too small to be detected with the delicacy of observation
possible at the present time ; but that in the case of Mercury it must amount to 43 seconds of arc
per century, a result which is strictly in agreement with observation.
Apart from this one, it has hitherto been possible to make only two deductions from the theory
which admit of being tested by observation, to wit, the curvature of light rays by the gravitational
field of the sun,
1)
and a displacement of the spectral lines of light reaching us from large stars, as
compared with the corresponding lines for light produced in an analogous manner terrestrially
(i.e. by the same kind of atom).
2)
These two deductions from the theory have both been confirmed.
Next: Part III: Considerations on the Universe as a Whole
Footnotes
1)
First observed by Eddington and others in 1919. (Cf. Appendix III, pp. 126−129).
Relativity: The Special and General Theory
62
2)
Established by Adams in 1924. (Cf. p. 132)
Relativity: The Special and General Theory
Relativity: The Special and General Theory
63
Albert Einstein: Relativity
Part III: Considerations on the Universe as a Whole
Part III
Considerations on the Universe as a Whole
Cosmological Difficulties of Netwon's Theory
Part from the difficulty discussed in Section 21, there is a second fundamental difficulty attending
classical celestial mechanics, which, to the best of my knowledge, was first discussed in detail by
the astronomer Seeliger. If we ponder over the question as to how the universe, considered as a
whole, is to be regarded, the first answer that suggests itself to us is surely this: As regards space
(and time) the universe is infinite. There are stars everywhere, so that the density of matter,
although very variable in detail, is nevertheless on the average everywhere the same. In other
words: However far we might travel through space, we should find everywhere an attenuated
swarm of fixed stars of approrimately the same kind and density.
This view is not in harmony with the theory of Newton. The latter theory rather requires that the
universe should have a kind of centre in which the density of the stars is a maximum, and that as
we proceed outwards from this centre the group−density of the stars should diminish, until finally, at
great distances, it is succeeded by an infinite region of emptiness. The stellar universe ought to be
a finite island in the infinite ocean of space.
1)
This conception is in itself not very satisfactory. It is still less satisfactory because it leads to the
result that the light emitted by the stars and also individual stars of the stellar system are
perpetually passing out into infinite space, never to return, and without ever again coming into
interaction with other objects of nature. Such a finite material universe would be destined to
become gradually but systematically impoverished.
In order to escape this dilemma, Seeliger suggested a modification of Newton's law, in which he
assumes that for great distances the force of attraction between two masses diminishes more
rapidly than would result from the inverse square law. In this way it is possible for the mean density
of matter to be constant everywhere, even to infinity, without infinitely large gravitational fields
being produced. We thus free ourselves from the distasteful conception that the material universe
ought to possess something of the nature of a centre. Of course we purchase our emancipation
from the fundamental difficulties mentioned, at the cost of a modification and complication of
Newton's law which has neither empirical nor theoretical foundation. We can imagine innumerable
laws which would serve the same purpose, without our being able to state a reason why one of
them is to be preferred to the others ; for any one of these laws would be founded just as little on
more general theoretical principles as is the law of Newton.
Next:
Relativity: The Special and General Theory
64
Footnotes
1)
Proof — According to the theory of Newton, the number of "lines of force" which come from
infinity and terminate in a mass m is proportional to the mass m. If, on the average, the Mass
density p
0
is constant throughout tithe universe, then a sphere of volume V will enclose the average
man p
0
V. Thus the number of lines of force passing through the surface F of the sphere into its
interior is proportional to p
0
V. For unit area of the surface of the sphere the number of lines of force
which enters the sphere is thus proportional to p
0
V/F or to p
0
R. Hence the intensity of the field at
the surface would ultimately become infinite with increasing radius R of the sphere, which is
impossible.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
65
Albert Einstein: Relativity
Part III: Considerations on the Universe as a Whole
The Possibility of a "Finite" and yet "Unbounded" Universe
But speculations on the structure of the universe also move in quite another direction. The
development of non−Euclidean geometry led to the recognition of the fact, that we can cast doubt
on the infiniteness of our space without coming into conflict with the laws of thought or with
experience (Riemann, Helmholtz). These questions have already been treated in detail and with
unsurpassable lucidity by Helmholtz and Poincaré, whereas I can only touch on them briefly here.
In the first place, we imagine an existence in two dimensional space. Flat beings with flat
implements, and in particular flat rigid measuring−rods, are free to move in a plane. For them
nothing exists outside of this plane: that which they observe to happen to themselves and to their
flat " things " is the all−inclusive reality of their plane. In particular, the constructions of plane
Euclidean geometry can be carried out by means of the rods e.g. the lattice construction,
considered in Section 24. In contrast to ours, the universe of these beings is two−dimensional; but,
like ours, it extends to infinity. In their universe there is room for an infinite number of identical
squares made up of rods, i.e. its volume (surface) is infinite. If these beings say their universe is "
plane," there is sense in the statement, because they mean that they can perform the constructions
of plane Euclidean geometry with their rods. In this connection the individual rods always represent
the same distance, independently of their position.
Let us consider now a second two−dimensional existence, but this time on a spherical surface
instead of on a plane. The flat beings with their measuring−rods and other objects fit exactly on this
surface and they are unable to leave it. Their whole universe of observation extends exclusively
over the surface of the sphere. Are these beings able to regard the geometry of their universe as
being plane geometry and their rods withal as the realisation of " distance " ? They cannot do this.
For if they attempt to realise a straight line, they will obtain a curve, which we " three−dimensional
beings " designate as a great circle, i.e. a self−contained line of definite finite length, which can be
measured up by means of a measuring−rod. Similarly, this universe has a finite area that can be
compared with the area, of a square constructed with rods. The great charm resulting from this
consideration lies in the recognition of the fact that the universe of these beings is finile and yet has
no limits.
But the spherical−surface beings do not need to go on a world−tour in order to perceive that they
are not living in a Euclidean universe. They can convince themselves of this on every part of their "
world," provided they do not use to small a piece of it. Starting from a point, they draw " straight
lines " (arcs of circles as judged in three dimensional space) of equal length in all directions. They
will call the line joining the free ends of these lines a " circle." For a plane surface, the ratio of the
circumference of a circle to its diameter, both lengths being measured with the same rod, is,
according to Euclidean geometry of the plane, equal to a constant value ¼, which is independent of
the diameter of the circle. On their spherical surface our flat beings would find for this ratio the
value
Relativity: The Special and General Theory
66
i.e. a smaller value than ¼, the difference being the more considerable, the greater is the radius of
the circle in comparison with the radius R of the " world−sphere." By means of this relation the
spherical beings can determine the radius of their universe (" world "), even when only a relatively
small part of their worldsphere is available for their measurements. But if this part is very small
indeed, they will no longer be able to demonstrate that they are on a spherical " world " and not on
a Euclidean plane, for a small part of a spherical surface differs only slightly from a piece of a plane
of the same size.
Thus if the spherical surface beings are living on a planet of which the solar system occupies only a
negligibly small part of the spherical universe, they have no means of determining whether they are
living in a finite or in an infinite universe, because the " piece of universe " to which they have
access is in both cases practically plane, or Euclidean. It follows directly from this discussion, that
for our sphere−beings the circumference of a circle first increases with the radius until the "
circumference of the universe " is reached, and that it thenceforward gradually decreases to zero
for still further increasing values of the radius. During this process the area of the circle continues to
increase more and more, until finally it becomes equal to the total area of the whole "
world−sphere."
Perhaps the reader will wonder why we have placed our " beings " on a sphere rather than on
another closed surface. But this choice has its justification in the fact that, of all closed surfaces, the
sphere is unique in possessing the property that all points on it are equivalent. I admit that the ratio
of the circumference c of a circle to its radius r depends on r, but for a given value of r it is the same
for all points of the " worldsphere "; in other words, the " world−sphere " is a " surface of constant
curvature."
To this two−dimensional sphere−universe there is a three−dimensional analogy, namely, the
three−dimensional spherical space which was discovered by Riemann. its points are likewise all
equivalent. It possesses a finite volume, which is determined by its "radius" (2¼
2
R
3
). Is it possible
to imagine a spherical space? To imagine a space means nothing else than that we imagine an
epitome of our " space " experience, i.e. of experience that we can have in the movement of " rigid
" bodies. In this sense we can imagine a spherical space.
Suppose we draw lines or stretch strings in all directions from a point, and mark off from each of
these the distance r with a measuring−rod. All the free end−points of these lengths lie on a
spherical surface. We can specially measure up the area (F) of this surface by means of a square
made up of measuring−rods. If the universe is Euclidean, then F = 4¼R
2
; if it is spherical, then F is
always less than 4¼R
2
. With increasing values of r, F increases from zero up to a maximum value
which is determined by the " world−radius," but for still further increasing values of r, the area
gradually diminishes to zero. At first, the straight lines which radiate from the starting point diverge
farther and farther from one another, but later they approach each other, and finally they run
together again at a "counter−point" to the starting point. Under such conditions they have traversed
the whole spherical space. It is easily seen that the three−dimensional spherical space is quite
analogous to the two−dimensional spherical surface. It is finite (i.e. of finite volume), and has no
bounds.
It may be mentioned that there is yet another kind of curved space: " elliptical space." It can be
regarded as a curved space in which the two " counter−points " are identical (indistinguishable from
each other). An elliptical universe can thus be considered to some extent as a curved universe
possessing central symmetry.
It follows from what has been said, that closed spaces without limits are conceivable. From
amongst these, the spherical space (and the elliptical) excels in its simplicity, since all points on it
are equivalent. As a result of this discussion, a most interesting question arises for astronomers
and physicists, and that is whether the universe in which we live is infinite, or whether it is finite in
Relativity: The Special and General Theory
67
the manner of the spherical universe. Our experience is far from being sufficient to enable us to
answer this question. But the general theory of relativity permits of our answering it with a moduate
degree of certainty, and in this connection the difficulty mentioned in Section 30 finds its solution.
Next: The Structure of Space According to the General Theory of Relativity
Relativity: The Special and General Theory
Relativity: The Special and General Theory
68
Albert Einstein: Relativity
Part III: Considerations on the Universe as a Whole
The Structure of Space According to the General Theory of Relativity
According to the general theory of relativity, the geometrical properties of space are not
independent, but they are determined by matter. Thus we can draw conclusions about the
geometrical structure of the universe only if we base our considerations on the state of the matter
as being something that is known. We know from experience that, for a suitably chosen
co−ordinate system, the velocities of the stars are small as compared with the velocity of
transmission of light. We can thus as a rough approximation arrive at a conclusion as to the nature
of the universe as a whole, if we treat the matter as being at rest.
We already know from our previous discussion that the behaviour of measuring−rods and clocks is
influenced by gravitational fields, i.e. by the distribution of matter. This in itself is sufficient to
exclude the possibility of the exact validity of Euclidean geometry in our universe. But it is
conceivable that our universe differs only slightly from a Euclidean one, and this notion seems all
the more probable, since calculations show that the metrics of surrounding space is influenced only
to an exceedingly small extent by masses even of the magnitude of our sun. We might imagine
that, as regards geometry, our universe behaves analogously to a surface which is irregularly
curved in its individual parts, but which nowhere departs appreciably from a plane: something like
the rippled surface of a lake. Such a universe might fittingly be called a quasi−Euclidean universe.
As regards its space it would be infinite. But calculation shows that in a quasi−Euclidean universe
the average density of matter would necessarily be nil. Thus such a universe could not be inhabited
by matter everywhere ; it would present to us that unsatisfactory picture which we portrayed in
Section 30.
If we are to have in the universe an average density of matter which differs from zero, however
small may be that difference, then the universe cannot be quasi−Euclidean. On the contrary, the
results of calculation indicate that if matter be distributed uniformly, the universe would necessarily
be spherical (or elliptical). Since in reality the detailed distribution of matter is not uniform, the real
universe will deviate in individual parts from the spherical, i.e. the universe will be quasi−spherical.
But it will be necessarily finite. In fact, the theory supplies us with a simple connection
1)
between
the space−expanse of the universe and the average density of matter in it.
Footnotes
1)
For the radius R of the universe we obtain the equation
The use of the C.G.S. system in this equation gives 2/k = 1
.
08.10
27
; p is the average density of the
matter and k is a constant connected with the Newtonian constant of gravitation.
Relativity: The Special and General Theory
69
Appendix I: Simple Derivation of the Lorentz Transformation
Relativity: The Special and General Theory
Relativity: The Special and General Theory
70
Albert Einstein: Relativity
Appendix
Appendix I
Simple Derivation of the Lorentz Transformation
(Supllementary to Section 11)
For the relative orientation of the co−ordinate systems indicated in Fig. 2, the x−axes of both
systems pernumently coincide. In the present case we can divide the problem into parts by
considering first only events which are localised on the x−axis. Any such event is represented with
respect to the co−ordinate system K by the abscissa x and the time t, and with respect to the
system K
1
by the abscissa x' and the time t'. We require to find x' and t' when x and t are given.
A light−signal, which is proceeding along the positive axis of x, is transmitted according to the
equation
x = ct
or
x − ct = 0 . . . (1).
Since the same light−signal has to be transmitted relative to K
1
with the velocity c, the propagation
relative to the system K
1
will be represented by the analogous formula
x' − ct' = O . . . (2)
Those space−time points (events) which satisfy (x) must also satisfy (2). Obviously this will be the
case when the relation
(x' − ct') = » (x − ct) . . . (3).
is fulfilled in general, where » indicates a constant ; for, according to (3), the disappearance of (x −
ct) involves the disappearance of (x' − ct').
If we apply quite similar considerations to light rays which are being transmitted along the negative
x−axis, we obtain the condition
(x' + ct') = µ(x + ct) . . . (4).
By adding (or subtracting) equations (3) and (4), and introducing for convenience the constants
a and b in place of the constants » and µ, where
and
Relativity: The Special and General Theory
71
we obtain the equations
We should thus have the solution of our problem, if the constants a and b were known. These
result from the following discussion.
For the origin of K
1
we have permanently x' = 0, and hence according to the first of the equations
(5)
If we call v the velocity with which the origin of K
1
is moving relative to K, we then have
The same value v can be obtained from equations (5), if we calculate the velocity of another point
of K
1
relative to K, or the velocity (directed towards the negative x−axis) of a point of K with respect
to K'. In short, we can designate v as the relative velocity of the two systems.
Furthermore, the principle of relativity teaches us that, as judged from K, the length of a unit
measuring−rod which is at rest with reference to K
1
must be exactly the same as the length, as
judged from K', of a unit measuring−rod which is at rest relative to K. In order to see how the points
of the x−axis appear as viewed from K, we only require to take a " snapshot " of K
1
from K; this
means that we have to insert a particular value of t (time of K), e.g. t = 0. For this value of t we then
obtain from the first of the equations (5)
x' = ax
Two points of the x'−axis which are separated by the distance ”x' = I when measured in the
K
1
system are thus separated in our instantaneous photograph by the distance
But if the snapshot be taken from K'(t' = 0), and if we eliminate t from the equations (5), taking into
account the expression (6), we obtain
From this we conclude that two points on the x−axis separated by the distance I (relative to K) will
be represented on our snapshot by the distance
Relativity: The Special and General Theory
72
But from what has been said, the two snapshots must be identical; hence ”x in (7) must be equal to
”x' in (7a), so that we obtain
The equations (6) and (7b) determine the constants a and b. By inserting the values of these
constants in (5), we obtain the first and the fourth of the equations given in Section 11.
Thus we have obtained the Lorentz transformation for events on the x−axis. It satisfies the
condition
x'
2
− c
2
t'
2
= x
2
− c
2
t
2
. . . (8a).
The extension of this result, to include events which take place outside the x−axis, is obtained by
retaining equations (8) and supplementing them by the relations
In this way we satisfy the postulate of the constancy of the velocity of light in vacuo for rays of light
of arbitrary direction, both for the system K and for the system K'. This may be shown in the
following manner.
We suppose a light−signal sent out from the origin of K at the time t = 0. It will be propagated
according to the equation
or, if we square this equation, according to the equation
x
2
+ y
2
+ z
2
= c
2
t
2
= 0 . . . (10).
It is required by the law of propagation of light, in conjunction with the postulate of relativity, that the
transmission of the signal in question should take place — as judged from K
1
— in accordance with
the corresponding formula
r' = ct'
Relativity: The Special and General Theory
73
or,
x'
2
+ y'
2
+ z'
2
− c
2
t'
2
= 0 . . . (10a).
In order that equation (10a) may be a consequence of equation (10), we must have
x'
2
+ y'
2
+ z'
2
− c
2
t'
2
= Ã (x
2
+ y
2
+ z
2
− c
2
t
2
) (11).
Since equation (8a) must hold for points on the x−axis, we thus have à = I. It is easily seen that the
Lorentz transformation really satisfies equation (11) for à = I; for (11) is a consequence of (8a) and
(9), and hence also of (8) and (9). We have thus derived the Lorentz transformation.
The Lorentz transformation represented by (8) and (9) still requires to be generalised. Obviously it
is immaterial whether the axes of K
1
be chosen so that they are spatially parallel to those of K. It is
also not essential that the velocity of translation of K
1
with respect to K should be in the direction of
the x−axis. A simple consideration shows that we are able to construct the Lorentz transformation
in this general sense from two kinds of transformations, viz. from Lorentz transformations in the
special sense and from purely spatial transformations. which corresponds to the replacement of the
rectangular co−ordinate system by a new system with its axes pointing in other directions.
Mathematically, we can characterise the generalised Lorentz transformation thus :
It expresses x', y', x', t', in terms of linear homogeneous functions of x, y, x, t, of such a kind that the
relation
x'
2
+ y'
2
+ z'
2
− c
2
t'
2
= x
2
+ y
2
+ z
2
− c
2
t
2
(11a).
is satisficd identically. That is to say: If we substitute their expressions in x, y, x, t, in place of x', y',
x', t', on the left−hand side, then the left−hand side of (11a) agrees with the right−hand side.
Next: Appendix II: Minkowski's Four Dimensional Space
Relativity: The Special and General Theory
Relativity: The Special and General Theory
74
Albert Einstein: Relativity
Appendix
Appendix II
Minkowski's Four−Dimensional Space ("World")
(supplementary to section 17)
We can characterise the Lorentz transformation still more simply if we introduce the imaginary
in place of t, as time−variable. If, in accordance with this, we insert
x
1
= x
x
2
= y
x
3
= z
x
4
=
and similarly for the accented system K
1
, then the condition which is identically satisfied by the
transformation can be expressed thus :
x
1
'
2
+ x
2
'
2
+ x
3
'
2
+ x
4
'
2
= x
1
2
+ x
2
2
+ x
3
2
+ x
4
2
(12).
That is, by the afore−mentioned choice of " coordinates," (11a) [see the end of Appendix II] is
transformed into this equation.
We see from (12) that the imaginary time co−ordinate x
4
, enters into the condition of transformation
in exactly the same way as the space co−ordinates x
1
, x
2
, x
3
. It is due to this fact that, according to
the theory of relativity, the " time "x
4
, enters into natural laws in the same form as the space co
ordinates x
1
, x
2
, x
3
.
A four−dimensional continuum described by the "co−ordinates" x
1
, x
2
, x
3
, x
4
, was called "world" by
Minkowski, who also termed a point−event a " world−point." From a "happening" in
three−dimensional space, physics becomes, as it were, an " existence " in the four−dimensional "
world."
This four−dimensional " world " bears a close similarity to the three−dimensional " space " of
(Euclidean) analytical geometry. If we introduce into the latter a new Cartesian co−ordinate system
(x'
1
, x'
2
, x'
3
) with the same origin, then x'
1
, x'
2
, x'
3
, are linear homogeneous functions of x
1
, x
2
,
x
3
which identically satisfy the equation
x'
1
2
+ x'
2
2
+ x'
3
2
= x
1
2
+ x
2
2
+ x
3
2
The analogy with (12) is a complete one. We can regard Minkowski's " world " in a formal manner
as a four−dimensional Euclidean space (with an imaginary time coordinate) ; the Lorentz
transformation corresponds to a " rotation " of the co−ordinate system in the fourdimensional "
world."
Next: The Experimental Confirmation of the General Theory of Relativity
Relativity: The Special and General Theory
75
Relativity: The Special and General Theory
Relativity: The Special and General Theory
76
Albert Einstein: Relativity
Appendix
Appendix III
The Experimental Confirmation of the General Theory of
Relativity
From a systematic theoretical point of view, we may imagine the process of evolution of an
empirical science to be a continuous process of induction. Theories are evolved and are expressed
in short compass as statements of a large number of individual observations in the form of
empirical laws, from which the general laws can be ascertained by comparison. Regarded in this
way, the development of a science bears some resemblance to the compilation of a classified
catalogue. It is, as it were, a purely empirical enterprise.
But this point of view by no means embraces the whole of the actual process ; for it slurs over the
important part played by intuition and deductive thought in the development of an exact science. As
soon as a science has emerged from its initial stages, theoretical advances are no longer achieved
merely by a process of arrangement. Guided by empirical data, the investigator rather develops a
system of thought which, in general, is built up logically from a small number of fundamental
assumptions, the so−called axioms. We call such a system of thought a theory. The theory finds
the justification for its existence in the fact that it correlates a large number of single observations,
and it is just here that the " truth " of the theory lies.
Corresponding to the same complex of empirical data, there may be several theories, which differ
from one another to a considerable extent. But as regards the deductions from the theories which
are capable of being tested, the agreement between the theories may be so complete that it
becomes difficult to find any deductions in which the two theories differ from each other. As an
example, a case of general interest is available in the province of biology, in the Darwinian theory
of the development of species by selection in the struggle for existence, and in the theory of
development which is based on the hypothesis of the hereditary transmission of acquired
characters.
We have another instance of far−reaching agreement between the deductions from two theories in
Newtonian mechanics on the one hand, and the general theory of relativity on the other. This
agreement goes so far, that up to the preseat we have been able to find only a few deductions from
the general theory of relativity which are capable of investigation, and to which the physics of
pre−relativity days does not also lead, and this despite the profound difference in the fundamental
assumptions of the two theories. In what follows, we shall again consider these important
deductions, and we shall also discuss the empirical evidence appertaining to them which has
hitherto been obtained.
(a) Motion of the Perihelion of Mercury
According to Newtonian mechanics and Newton's law of gravitation, a planet which is revolving
round the sun would describe an ellipse round the latter, or, more correctly, round the common
centre of gravity of the sun and the planet. In such a system, the sun, or the common centre of
gravity, lies in one of the foci of the orbital ellipse in such a manner that, in the course of a
planet−year, the distance sun−planet grows from a minimum to a maximum, and then decreases
Relativity: The Special and General Theory
77
again to a minimum. If instead of Newton's law we insert a somewhat different law of attraction into
the calculation, we find that, according to this new law, the motion would still take place in such a
manner that the distance sun−planet exhibits periodic variations; but in this case the angle
described by the line joining sun and planet during such a period (from perihelion—closest
proximity to the sun—to perihelion) would differ from 360
0
. The line of the orbit would not then be a
closed one but in the course of time it would fill up an annular part of the orbital plane, viz. between
the circle of least and the circle of greatest distance of the planet from the sun.
According also to the general theory of relativity, which differs of course from the theory of Newton,
a small variation from the Newton−Kepler motion of a planet in its orbit should take place, and in
such away, that the angle described by the radius sun−planet between one perhelion and the next
should exceed that corresponding to one complete revolution by an amount given by
(N.B. — One complete revolution corresponds to the angle 2À in the absolute angular measure
customary in physics, and the above expression giver the amount by which the radius sun−planet
exceeds this angle during the interval between one perihelion and the next.) In this expression
a represents the major semi−axis of the ellipse, e its eccentricity, c the velocity of light, and T the
period of revolution of the planet. Our result may also be stated as follows : According to the
general theory of relativity, the major axis of the ellipse rotates round the sun in the same sense as
the orbital motion of the planet. Theory requires that this rotation should amount to 43 seconds of
arc per century for the planet Mercury, but for the other Planets of our solar system its magnitude
should be so small that it would necessarily escape detection.
1)
In point of fact, astronomers have found that the theory of Newton does not suffice to calculate the
observed motion of Mercury with an exactness corresponding to that of the delicacy of observation
attainable at the present time. After taking account of all the disturbing influences exerted on
Mercury by the remaining planets, it was found (Leverrier: 1859; and Newcomb: 1895) that an
unexplained perihelial movement of the orbit of Mercury remained over, the amount of which does
not differ sensibly from the above mentioned +43 seconds of arc per century. The uncertainty of the
empirical result amounts to a few seconds only.
(b) Deflection of Light by a Gravitational Field
In Section 22 it has been already mentioned that according to the general theory of relativity, a ray
of light will experience a curvature of its path when passing through a gravitational field, this
curvature being similar to that experienced by the path of a body which is projected through a
gravitational field. As a result of this theory, we should expect that a ray of light which is passing
close to a heavenly body would be deviated towards the latter. For a ray of light which passes the
sun at a distance of ” sun−radii from its centre, the angle of deflection (a) should amount to
Relativity: The Special and General Theory
78
It may be added that, according to the theory, half of this deflection is produced
by the Newtonian field of attraction of the sun, and the other half by the geometrical modification ("
curvature ") of space caused by the sun.
This result admits of an experimental test by means of the photographic registration of stars during
a total eclipse of the sun. The only reason why we must wait for a total eclipse is because at every
other time the atmosphere is so strongly illuminated by the light from the sun that the stars situated
near the sun's disc are invisible. The predicted effect can be seen clearly from the accompanying
diagram. If the sun (S) were not present, a star which is practically infinitely distant would be seen
in the direction D
1
, as observed front the earth. But as a consequence of the deflection of light from
the star by the sun, the star will be seen in the direction D
2
, i.e. at a somewhat greater distance
from the centre of the sun than corresponds to its real position.
In practice, the question is tested in the following way. The stars in the neighbourhood of the sun
are photographed during a solar eclipse. In addition, a second photograph of the same stars is
taken when the sun is situated at another position in the sky, i.e. a few months earlier or later. As
compared whh the standard photograph, the positions of the stars on the eclipse−photograph
ought to appear displaced radially outwards (away from the centre of the sun) by an amount
corresponding to the angle a.
We are indebted to the [British] Royal Society and to the Royal Astronomical Society for the
investigation of this important deduction. Undaunted by the [first world] war and by difficulties of
both a material and a psychological nature aroused by the war, these societies equipped two
expeditions — to Sobral (Brazil), and to the island of Principe (West Africa) — and sent several of
Britain's most celebrated astronomers (Eddington, Cottingham, Crommelin, Davidson), in order to
obtain photographs of the solar eclipse of 29th May, 1919. The relative discrepancies to be
expected between the stellar photographs obtained during the eclipse and the comparison
photographs amounted to a few hundredths of a millimetre only. Thus great accuracy was
necessary in making the adjustments required for the taking of the photographs, and in their
subsequent measurement.
The results of the measurements confirmed the theory in a thoroughly satisfactory manner. The
rectangular components of the observed and of the calculated deviations of the stars (in seconds of
arc) are set forth in the following table of results :
Relativity: The Special and General Theory
79
(c) Displacement of Spectral Lines Towards the Red
In Section 23 it has been shown that in a system K
1
which is in rotation with regard to a Galileian
system K, clocks of identical construction, and which are considered at rest with respect to the
rotating reference−body, go at rates which are dependent on the positions of the clocks. We shall
now examine this dependence quantitatively. A clock, which is situated at a distance r from the
centre of the disc, has a velocity relative to K which is given by
V = wr
where w represents the angular velocity of rotation of the disc K
1
with respect to K. If v
0
, represents
the number of ticks of the clock per unit time (" rate " of the clock) relative to K when the clock is at
rest, then the " rate " of the clock (v) when it is moving relative to K with a velocity V, but at rest with
respect to the disc, will, in accordance with Section 12, be given by
or with sufficient accuracy by
This expression may also be stated in the following form:
If we represent the difference of potential of the centrifugal force between the position of the clock
and the centre of the disc by Æ, i.e. the work, considered negatively, which must be performed on
the unit of mass against the centrifugal force in order to transport it from the position of the clock on
the rotating disc to the centre of the disc, then we have
From this it follows that
In the first place, we see from this expression that two clocks of identical construction will go at
different rates when situated at different distances from the centre of the disc. This result is aiso
valid from the standpoint of an observer who is rotating with the disc.
Now, as judged from the disc, the latter is in a gravititional field of potential Æ, hence the result we
have obtained will hold quite generally for gravitational fields. Furthermore, we can regard an atom
which is emitting spectral lines as a clock, so that the following statement will hold:
An atom absorbs or emits light of a frequency which is dependent on the potential of the
gravitational field in which it is situated.
Relativity: The Special and General Theory
80
The frequency of an atom situated on the surface of a heavenly body will be somewhat less than
the frequency of an atom of the same element which is situated in free space (or on the surface of
a smaller celestial body).
Now Æ = − K (M/r), where K is Newton's constant of gravitation, and M is the mass of the heavenly
body. Thus a displacement towards the red ought to take place for spectral lines produced at the
surface of stars as compared with the spectral lines of the same element produced at the surface of
the earth, the amount of this displacement being
For the sun, the displacement towards the red predicted by theory amounts to about two millionths
of the wave−length. A trustworthy calculation is not possible in the case of the stars, because in
general neither the mass M nor the radius r are known.
It is an open question whether or not this effect exists, and at the present time (1920) astronomers
are working with great zeal towards the solution. Owing to the smallness of the effect in the case of
the sun, it is difficult to form an opinion as to its existence. Whereas Grebe and Bachem (Bonn), as
a result of their own measurements and those of Evershed and Schwarzschild on the cyanogen
bands, have placed the existence of the effect almost beyond doubt, while other investigators,
particularly St. John, have been led to the opposite opinion in consequence of their measurements.
Mean displacements of lines towards the less refrangible end of the spectrum are certainly
revealed by statistical investigations of the fixed stars ; but up to the present the examination of the
available data does not allow of any definite decision being arrived at, as to whether or not these
displacements are to be referred in reality to the effect of gravitation. The results of observation
have been collected together, and discussed in detail from the standpoint of the question which has
been engaging our attention here, in a paper by E. Freundlich entitled "Zur Prüfung der
allgemeinen Relativitâts−Theorie" (Die Naturwissenschaften, 1919, No. 35, p. 520: Julius Springer,
Berlin).
At all events, a definite decision will be reached during the next few years. If the displacement of
spectral lines towards the red by the gravitational potential does not exist, then the general theory
of relativity will be untenable. On the other hand, if the cause of the displacement of spectral lines
be definitely traced to the gravitational potential, then the study of this displacement will furnish us
with important information as to the mass of the heavenly bodies.
[A]
Next: Appendix IV: The Structure of Space According to the General Theory of Relativity
Footnotes
1)
Especially since the next planet Venus has an orbit that is almost an exact circle, which makes it
more difficult to locate the perihelion with precision.
[A]
The displacentent of spectral lines towards the red end of the spectrum was definitely
established by Adams in 1924, by observations on the dense companion of Sirius, for which the
effect is about thirty times greater than for the Sun. R.W.L. — translator
Relativity: The Special and General Theory
81
Relativity: The Special and General Theory
Relativity: The Special and General Theory
82
Albert Einstein: Relativity
Appendix
Appendix IV
The Structure of Space According to the General Theory of
Relativity
(Supplementary to Section 32)
Since the publication of the first edition of this little book, our knowledge about the structure of
space in the large (" cosmological problem ") has had an important development, which ought to be
mentioned even in a popular presentation of the subject.
My original considerations on the subject were based on two hypotheses:
(1) There exists an average density of matter in the whole of space which is everywhere the same
and different from zero.
(2) The magnitude (" radius ") of space is independent of time.
Both these hypotheses proved to be consistent, according to the general theory of relativity, but
only after a hypothetical term was added to the field equations, a term which was not required by
the theory as such nor did it seem natural from a theoretical point of view (" cosmological term of
the field equations ").
Hypothesis (2) appeared unavoidable to me at the time, since I thought that one would get into
bottomless speculations if one departed from it.
However, already in the 'twenties, the Russian mathematician Friedman showed that a different
hypothesis was natural from a purely theoretical point of view. He realized that it was possible to
preserve hypothesis (1) without introducing the less natural cosmological term into the field
equations of gravitation, if one was ready to drop hypothesis (2). Namely, the original field
equations admit a solution in which the " world radius " depends on time (expanding space). In that
sense one can say, according to Friedman, that the theory demands an expansion of space.
A few years later Hubble showed, by a special investigation of the extra−galactic nebulae (" milky
ways "), that the spectral lines emitted showed a red shift which increased regularly with the
distance of the nebulae. This can be interpreted in regard to our present knowledge only in the
sense of Doppler's principle, as an expansive motion of the system of stars in the large — as
required, according to Friedman, by the field equations of gravitation. Hubble's discovery can,
therefore, be considered to some extent as a confirmation of the theory.
There does arise, however, a strange difficulty. The interpretation of the galactic line−shift
discovered by Hubble as an expansion (which can hardly be doubted from a theoretical point of
view), leads to an origin of this expansion which lies " only " about 10
9
years ago, while physical
astronomy makes it appear likely that the development of individual stars and systems of stars
takes considerably longer. It is in no way known how this incongruity is to be overcome.
Relativity: The Special and General Theory
83
I further want to rernark that the theory of expanding space, together with the empirical data of
astronomy, permit no decision to be reached about the finite or infinite character of
(three−dimensional) space, while the original " static " hypothesis of space yielded the closure
(finiteness) of space.
Relativity: The Special and General Theory
Relativity: The Special and General Theory
84
Livros Grátis
( http://www.livrosgratis.com.br )
Milhares de Livros para Download:
Baixar livros de Administração
Baixar livros de Agronomia
Baixar livros de Arquitetura
Baixar livros de Artes
Baixar livros de Astronomia
Baixar livros de Biologia Geral
Baixar livros de Ciência da Computação
Baixar livros de Ciência da Informação
Baixar livros de Ciência Política
Baixar livros de Ciências da Saúde
Baixar livros de Comunicação
Baixar livros do Conselho Nacional de Educação - CNE
Baixar livros de Defesa civil
Baixar livros de Direito
Baixar livros de Direitos humanos
Baixar livros de Economia
Baixar livros de Economia Doméstica
Baixar livros de Educação
Baixar livros de Educação - Trânsito
Baixar livros de Educação Física
Baixar livros de Engenharia Aeroespacial
Baixar livros de Farmácia
Baixar livros de Filosofia
Baixar livros de Física
Baixar livros de Geociências
Baixar livros de Geografia
Baixar livros de História
Baixar livros de Línguas
Baixar livros de Literatura
Baixar livros de Literatura de Cordel
Baixar livros de Literatura Infantil
Baixar livros de Matemática
Baixar livros de Medicina
Baixar livros de Medicina Veterinária
Baixar livros de Meio Ambiente
Baixar livros de Meteorologia
Baixar Monografias e TCC
Baixar livros Multidisciplinar
Baixar livros de Música
Baixar livros de Psicologia
Baixar livros de Química
Baixar livros de Saúde Coletiva
Baixar livros de Serviço Social
Baixar livros de Sociologia
Baixar livros de Teologia
Baixar livros de Trabalho
Baixar livros de Turismo