0% found this document useful (0 votes)
22 views81 pages

31812

Uploaded by

krisnaegecan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
22 views81 pages

31812

Uploaded by

krisnaegecan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 81

Download the full version of the ebook at

https://ebookultra.com

Elementary general relativity Macdonald A.

https://ebookultra.com/download/elementary-
general-relativity-macdonald-a/

Explore and download more ebook at https://ebookultra.com


Recommended digital products (PDF, EPUB, MOBI) that
you can download immediately if you are interested.

Relativity an introduction to special and general


relativity 3ed. Edition Hans Stephani

https://ebookultra.com/download/relativity-an-introduction-to-special-
and-general-relativity-3ed-edition-hans-stephani/

ebookultra.com

Conformal Methods in General Relativity 1st Edition Juan


A. Valiente Kroon

https://ebookultra.com/download/conformal-methods-in-general-
relativity-1st-edition-juan-a-valiente-kroon/

ebookultra.com

General Relativity and Gravitation A Centennial


Perspective 1st Edition Abhay Ashtekar

https://ebookultra.com/download/general-relativity-and-gravitation-a-
centennial-perspective-1st-edition-abhay-ashtekar/

ebookultra.com

Modern canonical quantum general relativity 1st Edition


Thomas Thiemann

https://ebookultra.com/download/modern-canonical-quantum-general-
relativity-1st-edition-thomas-thiemann/

ebookultra.com
Introduction to General Relativity and Cosmology 1st
Edition J. Plebanski

https://ebookultra.com/download/introduction-to-general-relativity-
and-cosmology-1st-edition-j-plebanski/

ebookultra.com

An Introduction to General Relativity and Cosmology 1st


Edition Jerzy Plebanski

https://ebookultra.com/download/an-introduction-to-general-relativity-
and-cosmology-1st-edition-jerzy-plebanski/

ebookultra.com

Gravity An Introduction to Einstein s General Relativity


Solutions to Problems 1.1 Edition James B. Hartle

https://ebookultra.com/download/gravity-an-introduction-to-einstein-s-
general-relativity-solutions-to-problems-1-1-edition-james-b-hartle/

ebookultra.com

Calculus 3c 2 Examples of General Elementary Series 1st


edition Edition Mejlbro L.

https://ebookultra.com/download/calculus-3c-2-examples-of-general-
elementary-series-1st-edition-edition-mejlbro-l/

ebookultra.com

Inside Stars A Theory of the Internal Constitution of


Stars and the Sources of Stellar Energy According to
General Relativity Second Edition Larissa Borissova
https://ebookultra.com/download/inside-stars-a-theory-of-the-internal-
constitution-of-stars-and-the-sources-of-stellar-energy-according-to-
general-relativity-second-edition-larissa-borissova/
ebookultra.com
Elementary general relativity Macdonald A. Digital
Instant Download
Author(s): Macdonald A.
ISBN(s): 9789810210786, 9810210787
Edition: lecture notes, web draft v3.35
File Details: PDF, 1.16 MB
Year: 2004
Language: english
Elementary
General
Relativity
Version 3.35

Alan Macdonald
Luther College, Decorah, IA USA
mailto:[email protected]
http://faculty.luther.edu/∼macdonal
c
To Ellen
“The magic of this theory will hardly fail to impose itself on anybody
who has truly understood it.”

Albert Einstein, 1915

“The foundation of general relativity appeared to me then [1915],


and it still does, the greatest feat of human thinking about Nature,
the most amazing combination of philosophical penetration, physical
intuition, and mathematical skill.”

Max Born, 1955

“One of the principal objects of theoretical research in any depart-


ment of knowledge is to find the point of view from which the subject
appears in its greatest simplicity.”

Josiah Willard Gibbs

“There is a widespread indifference to attempts to put accepted the-


ories on better logical foundations and to clarify their experimental
basis, an indifference occasionally amounting to hostility. I am con-
cerned with the effects of our neglect of foundations on the educa-
tion of scientists. It is plain that the clearer the teacher, the more
transparent his logic, the fewer and more decisive the number of ex-
periments to be examined in detail, the faster will the pupil learn
and the surer and sounder will be his grasp of the subject.”

Sir Hermann Bondi

“Things should be made as simple as possible, but not simpler.”

Albert Einstein
Contents

Preface

1 Flat Spacetimes
1.1 Spacetimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2 The Inertial Frame Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.3 The Metric Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4 The Geodesic Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2 Curved Spacetimes
2.1 History of Theories of Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 The Key to General Relativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3 The Local Inertial Frame Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.4 The Metric Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.5 The Geodesic Postulate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.6 The Field Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3 Spherically Symmetric Spacetimes


3.1 Stellar Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .49
3.2 The Schwartzschild Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.3 The Solar System Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.4 Kerr Spacetimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.5 The Binary Pulsar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.6 Black Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4 Cosmological Spacetimes
4.1 Our Universe I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .66
4.2 The Robertson-Walker Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.3 The Expansion Redshift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.4 Our Universe II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.5 General Relativity Today . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Preface
The purpose of this book is to provide, with a minimum of mathematical
machinery and in the fewest possible pages, a clear and careful explanation of
the physical principles and applications of classical general relativity. The pre-
requisites are single variable calculus, a few basic facts about partial derivatives
and line integrals, a little matrix algebra, and some basic physics.
The book is for those seeking a conceptual understanding of the theory, not
computational prowess. Despite it’s brevity and modest prerequisites, it is a
serious introduction to the physics and mathematics of general relativity which
demands careful study. The book can stand alone as an introduction to general
relativity or it can be used as an adjunct to standard texts.
Chapter 1 is a self-contained introduction to those parts of special relativity
we require for general relativity. We take a nonstandard approach to the metric,
analogous to the standard approach to the metric in Euclidean geometry. In
geometry, distance is first understood geometrically, independently of any coor-
dinate system. If coordinates are introduced, then distances can be expressed in
terms of coordinate differences: ∆s2 = ∆x2 + ∆y 2 . The formula is important,
but the geometric meaning of the distance is fundamental.
Analogously, we define the spacetime interval of special relativity physically,
independently of any coordinate system. If inertial frame coordinates are in-
troduced, then the interval can be expressed in terms of coordinate differences:
∆s2 = ∆t2 − ∆x2 − ∆y 2 − ∆z 2 . The formula is important, but the physical
meaning of the interval is fundamental. I believe that this approach to the met-
ric provides easier access to and deeper understanding of special relativity, and
facilitates the transition to general relativity.
Chapter 2 introduces the physical principles on which general relativity is
based. The basic concepts of Riemannian geometry are developed in order to
express these principles mathematically as postulates. The purpose of the pos-
tulates is not to achieve complete rigor – which is neither desirable nor possible
in a book at this level – but to state clearly the physical principles, and to
exhibit clearly the relationship to special relativity and the analogy with sur-
faces. The postulates are in one-to-one correspondence with the fundamental
concepts of Riemannian geometry: manifold, metric, geodesic, and curvature.
Concentrating on the physical meaning of the metric greatly simplifies the de-
velopment of general relativity. In particular, tensors are not needed. There
is, however, a brief introcution to tensors in an appendix. (Similarly, modern
elementary differential geometry texts often develop the intrinsic geometry of
curved surfaces by focusing on the geometric meaning of the metric. Tensors
are not used.)
The first two chapters systematically exploit the mathematical analogy which
led to general relativity: a curved spacetime is to a flat spacetime as a curved
surface is to a flat surface. Before introducing a spacetime concept, its analog
for surfaces is presented. This is not a new idea, but it is used here more system-
atically than elsewhere. For example, when the metric ds of general relativity
is introduced, the reader has already seen a metric in three other contexts.
Chapter 3 solves the field equation for a spherically symmetric spacetime
to obtain the Schwartzschild metric. The geodesic equations are then solved
and applied to the classical solar system tests of general relativity. There is a
section on the Kerr metric, including gravitomagnetism and the Gravity Probe
B experiment. The chapter closes with sections on the binary pulsar and black
holes. In this chapter, as elsewhere, I have tried to provide the cleanest possible
calculations.
Chapter 4 applies general relativity to cosmology. We obtain the Robertson-
Walker metric in an elementary manner without using the field equation. We
then solve the field equation with a nonzero cosmological constant for a flat
Robertson-Walker spacetime. WMAP data allow us to specify all parameters in
the solution, giving the new “standard model” of the universe with dark matter
and dark energy.
There have been many spectacular astronomical discoveries and observa-
tions since 1960 which are relevant to general relativity. We describe them at
appropriate places in the book.
Some 50 exercises are scattered throughout. They often serve as examples
of concepts introduced in the text. If they are not done, they should be read.
Some tedious (but always straightforward) calculations have been omitted.
They are best carried out with a computer algebra system. Some material has
been placed in about 20 pages of appendices to keep the main line of development
visible. The appendices occasionally require more background of the reader than
the text. They may be omitted without loss of anything essential. Appendix
1 gives the values of various physical constants. Appendix 2 contains several
approximation formulas used in the text.
Chapter 1

Flat Spacetimes

1.1 Spacetimes
The general theory of relativity is our best theory of space, time, and gravity. It
is commonly felt to be the most beautiful of all physical theories. Albert Einstein
created the theory during the decade following the publication, in 1905, of his
special theory of relativity. The special theory is a theory of space and time
which applies when gravity is insignificant. The general theory generalizes the
special theory to include gravity.
In geometry the fundamental entities are points. A point is a specific place.
In relativity the fundamental entities are events. An event is a specific time and
place. For example, the collision of two particles is an event. A concert is an
event (idealizing it to a single time and place). To attend the concert, you must
be at the time and the place of the event.
A flat or curved surface is a set of points. (We shall prefer the term “flat
surface” to “plane”.) Similarly, a spacetime is a set of events. For example,
we might consider the events in a specific room between two specific times. A
flat spacetime is one without significant gravity. Special relativity describes
flat spacetimes. A curved spacetime is one with significant gravity. General
relativity describes curved spacetimes.
There is nothing mysterious about the words “flat” or “curved” attached
to a set of events. They are chosen because of a remarkable analogy, already
hinted at, concerning the mathematical description of a spacetime: a curved
spacetime is to a flat spacetime as a curved surface is to a flat surface. This
analogy will be a major theme of this book; we shall use the mathematics of
flat and curved surfaces to guide our understanding of the mathematics of flat
and curved spacetimes.
We shall explore spacetimes with clocks to measure time and rods (rulers)
to measure space, i.e., distance. However, clocks and rods do not in fact live
up to all that we usually expect of them. In this section we shall see what we
expect of them in relativity.

11
1.1 Spacetimes

Clocks. A curve is a continuous succession of points in a surface. Similarly,


a worldline is a continuous succession of events in a spacetime. A moving
particle or a pulse of light emitted in a single direction is present at a continuous
succession of events, its worldline. Even if a particle is at rest, time passes, and
the particle has a worldline.
The length of a curve between two given points depends on the curve. Sim-
ilarly, the time between two given events measured by a clock moving between
the events depends on the clock’s worldline! J. C. Hafele and R. Keating pro-
vided the most direct verification of this in 1971. They brought two atomic
clocks together, then placed them in separate airplanes which circled the Earth
in opposite directions, and then brought the clocks together again. Thus the
clocks took different worldlines between the event of their separation and the
event of their reunion. They measured different times between the two events.
The difference was small, about 10−7 sec, but was well within the ability of the
clocks to measure. There is no doubt that the effect is real.
Relativity predicts the measured difference. Exercise 1.10 shows that special
relativity predicts a difference between the clocks. Exercise 2.1 shows that
general relativity predicts a further difference. Exercise 3.3 shows that general
relativity predicts the observed difference. Relativity prtedicts large differences
between clocks whose relative velocity is close to the speed of light.
The best answer to the question “How can the clocks in the experiment
possibly disagree?” is the question “Why should they agree?” After all, the
clocks are not connected. According to everyday ideas they should agree because
there is a universal time, common to all observers. It is the duty of clocks to
report this time. The concept of a universal time was abstracted from experience
with small speeds (compared to that of light) and clocks of ordinary accuracy,
where it is nearly valid. The concept permeates our daily lives; there are clocks
everywhere telling us the time. However, the Hafele-Keating experiment shows
that there is no universal time. Time, like distance, is route dependent.
Since clocks on different worldlines between two events measure different
times between the events, we cannot speak of the time between two events.
However, the relative rates of processes – the ticking of a clock, the frequency of
a tuning fork, the aging of an organism, etc. – are the same along all worldlines.
(Unless some adverse physical condition affects a rate.) Twins traveling in the
two airplanes of the Hafele-Keating experiment would each age according to
the clock in their airplane. They would thus be of slightly different ages when
reunited.

12
1.1 Spacetimes

Rods. Consider astronauts in interstellar space, where gravity is insignifi-


cant. If their rocket is not firing and their ship is not spinning, then they will
feel no forces acting on them and they can float freely in their cabin. If their
spaceship is accelerating, then they will feel a force pushing them back against
their seat. If the ship turns to the left, then they will feel a force to the right.
If the ship is spinning, they will feel a force outward from the axis of spin. Call
these forces inertial forces.
Accelerometers measure inertial forces. Fig. 1.1
shows a simple accelerometer consisting of a weight held
at the center of a frame by identical springs. Inertial
forces cause the weight to move from the center. An in-
ertial object is one which experiences no inertial forces.
If an object is inertial, then any object moving at a con-
stant velocity with respect to it is also inertial and any
object accelerating with respect to it is not inertial.
In special relativity we make an assumption which
allows us to speak of the distance between two inertial
objects at rest with respect to each other: Inertial rigid Fig. 1.1: An
rods side by side and at rest with respect to two inertial accelerometer.
objects measure the same distance between the objects. The weight is held
More precisely, we assume that any difference is due to at the center by
some adverse physical cause (e.g., thermal expansion) springs. Acceler-
ation causes the
to which an “ideal” rigid rod would not be subject. In
weight to move
particular, the history of a rigid rod does not affect its from the center.
length. Noninertial rods are difficult to deal with in
relativity, and we shall not consider them.
In the next three sections we give three postulates for special relativity.
The inertial frame postulate asserts that certain natural coordinate systems,
called inertial frames, exist for a flat spacetime. The metric postulate asserts
a universal light speed and a slowing of clocks moving in inertial frames. The
geodesic postulate asserts that inertial particles and light move in a straight line
at constant speed in inertial frames.
We shall use the analogy mentioned above to help us understand the pos-
tulates. Imagine two dimensional beings living in a flat surface. These surface
dwellers can no more imagine leaving their two spatial dimensions than we can
imagine leaving our three spatial dimensions. Before introducing a postulate
for a flat spacetime, we introduce the analogous postulate formulated by sur-
face dwellers for a flat surface. The postulates for a flat spacetime use a time
dimension, but those for a flat surface do not.

13
1.2 The Inertial Frame Postulate

1.2 The Inertial Frame Postulate


Surface dwellers find it useful to label the points of their flat surface with co-
ordinates. They construct, using identical rigid rods, a square grid and assign
rectangular coordinates (x, y) to the nodes of the grid in the usual way. See
Fig. 1.2. They specify a point by using the coordinates of the node nearest
the point. If more accurate coordinates are required, they build a finer grid.
Surface dwellers call the coordinate system a planar frame. They postulate:
The Planar Frame Postulate for a Flat Surface
A planar frame can be constructed with any point P as origin and
with any orientation.
Similarly, it is useful to label the events in a flat
spacetime with coordinates (t, x, y, z). The coordi-
nates specify when and where the event occurs, i.e.,
they completely specify the event. We now describe
how to attach coordinates to events. The procedure
is idealized, but it gives a clear physical meaning to
the coordinates.
To specify where an event occurs, construct, us-
ing identical rigid rods, an inertial cubical lattice.
See Fig. 1.3. Assign rectangular coordinates (x, y, z)
Fig. 1.2: A planar
to the nodes of the lattice in the usual way. Specify
frame.
where an event occurs by using the coordinates of
the node nearest the event.
To specify when an event occurs, place a clock at
each node of the lattice. Then the times of events at
a given node can be specified by reading the clock at
that node. But to compare meaningfully the times of
events at different nodes, the clocks must be in some
sense synchronized. As we shall see soon, this is not
a trivial matter. (Remember, there is no universal
time.) For now, assume that the clocks have been
synchronized. Then specify when an event occurs by
using the time, t, on the clock at the node nearest the
event. And measure the coordinate time difference
Fig. 1.3: An inertial ∆t between two events using the synchronized clocks
lattice.
at the nodes where the events occur. Note that this
requires two clocks.
The four dimensional coordinate system obtained in this way from an inertial
cubical lattice with synchronized clocks is called an inertial frame. The event
(t, x, y, z) = (0, 0, 0, 0) is the origin of the inertial frame. We postulate:
The Inertial Frame Postulate for a Flat Spacetime
An inertial frame can be constructed with any event E as origin,
with any orientation, and with any inertial object at E at rest in it.

14
1.2 The Inertial Frame Postulate

If we suppress one or two of the spatial coordinates of an inertial frame, then


we can draw a spacetime diagram and depict worldlines. For example, Fig. 1.4
shows the worldlines of two particles. One is at rest on the x-axis and one moves
away from x = 0 and then returns more slowly.

Fig. 1.4: Two worldlines. Fig. 1.5: Worldline of a particle


moving with constant speed.

Exercise 1.1. Show that the worldline of an object moving along the x-axis
at constant speed v is a straight line with slope v. See Fig. 1.5.
Exercise 1.2. Describe the worldline of an object moving in a circle in the
z = 0 plane at constant speed v.
Synchronization. We return to the matter of synchronizing the clocks in
the lattice. What does it mean to say that separated clocks are synchronized?
Einstein realized that the answer to this question is not given to us by Nature;
rather, it must be answered with a definition.
Exercise 1.3. Why not simply bring the clocks together, synchronize them,
move them to the nodes of the lattice, and call them synchronized?
We might try the following definition. Send a signal from a node P of the
lattice at time tP according to the clock at P . Let it arrive at a node Q of
the lattice at time tQ according to the clock at Q. Let the distance between
the nodes be D and the speed of the signal be v. Say that the clocks are
synchronized if
tQ = tP + D/v. (1.1)
Intuitively, the term D/v compensates for the time it takes the signal to get to
Q. This definition is flawed because it contains a logical circle: v is defined by
a rearrangement of Eq. (1.1): v = D/(tQ − tP ). Synchronized clocks cannot be
defined using v because synchronized clocks are needed to define v.
We adopt the following definition, essentially due to Einstein. Emit a pulse
of light from a node P at time tP according to the clock at P . Let it arrive at a
node Q at time tQ according to the clock at Q. Similarly, emit a pulse of light
from Q at time t0Q and let it arrive at P at t0P . The clocks are synchronized if

tQ − tP = t0P − t0Q , (1.2)

i.e., if the times in the two directions are equal.

15
1.2 The Inertial Frame Postulate

Reformulating the definition makes it more transparent. If the pulse from


Q to P is the reflection of the pulse from P to Q, then t0Q = tQ in Eq. (1.2).
Let 2T be the round trip time: 2T = t0P − tP . Substitute this in Eq. (1.2):
t Q = tP + T ; (1.3)
the clocks are synchronized if the pulse arrives at Q in half the time it takes for
the round trip.
Exercise 1.4. Explain why Eq. (1.2) is a satisfactory definition but Eq.
(1.1) is not.
There is a tacit assumption in the definition of synchronized clocks that the
two sides of Eq. (1.2) do not depend on the times that the pulses are sent:
Emit pulses of light from a node R at times tR and t0R according
to a clock at R. Let them arrive at a node S at times tS and t0S
according to a clock at S. Then
t0S − t0R = tS − tR . (1.4)

With this assumption we can be sure that synchronized clocks will remain syn-
chronized.
Exercise 1.5. Show that with the assumption Eq. (1.4), T in Eq. (1.3) is
independent of the time the pulse is sent.
A rearrangement of Eq. (1.4) gives
∆so = ∆se , (1.5)
where ∆so = t0S − tS is the time between the observation of the pulses at S and
∆se = t0R − tR is the time between the emission of the pulses at R. (We use ∆s
rather than ∆t to conform to notation used later in more general situations.) If
a clock at R emits pulses of light at regular intervals to S, then Eq. (1.5) states
that an observer at S sees (actually sees) the clock at R going at the same rate
as his clock. Of course, the observer at S will see all physical processes at R
proceed at the same rate they do at S.
Redshifts. We will encounter situations in which ∆so 6= ∆se . Define the
redshift
∆so
z= − 1. (1.6)
∆se
Equations (1.4) and (1.5) correspond to z = 0. If, for example, z = 1 (∆so /∆se
= 2), then the observer at S would see clocks at R, and all other physical
processes at R, proceed at half the rate they do at S.
If the two “pulses” of light in Eq. (1.6) are successive wavecrests of light
emitted at frequency fe = (∆se )−1 and observed at frequency fo = (∆so )−1 ,
then Eq. (1.6) can be written
fe
z= − 1. (1.7)
fo

16
1.2 The Inertial Frame Postulate

In Exercise 1.6 we shall see that Eq. (1.5) is violated, i.e., z 6= 0, if the
emitter and observer are in relative motion in a flat spacetime. This is called
a Doppler redshift. Later we shall see two other kinds of redshift: gravitational
redshifts in Sec. 2.2 and expansion redshifts in Sec. 4.1. The three types of
redshifts have different physical origins and so must be carefully distinguished.
Synchronization. The inertial frame pos-
tulate asserts in part that clocks in an inertial
lattice can be synchronized according to the
definition Eq. (1.2), or, in P. W. Bridgeman’s
descriptive phrase, we can “spread time over
space”. We now prove this with the aid of an
auxiliary assumption. The reader may skip the
proof and turn to the next section without loss
of continuity.
Let 2T be the time, as measured by a clock
at the origin O of the lattice, for light to travel
from O to another node Q and return after
being reflected at Q. Emit a pulse of light at O
toward Q at time tO according to the clock at Fig. 1.6: Light traversing
a triangle in opposite direc-
O. When the pulse arrives at Q set the clock
tions.
there to tQ = tO + T . According to Eq. (1.3)
the clocks at O and Q are now synchronized.
Synchronize all clocks with the one at O in this way.
To show that the clocks at any two nodes P and Q are now synchronized
with each other, we must make this assumption:
The time it takes light to traverse a triangle in the lattice is inde-
pendent of the direction taken around the triangle.
See Fig. 1.6. In an experiment performed in 1963, W. M. Macek and D. T.
M. Davis, Jr. verified the assumption for a square to one part in 1012 . See
Appendix 3.
Reflect a pulse of light around the triangle OP Q. Let the pulse be at
O, P, Q, O at times tO , tP , tQ , tR according to the clock at that node. Similarly,
let the times for a pulse sent around in the other direction be t0O , t0Q , t0P , t0R .
See Fig. 1.6. We have the algebraic identities
tR − tO = (tR − tP ) + (tP − tQ ) + (tQ − tO )
t0R − t0O = (t0R − t0P ) + (t0P − t0Q ) + (t0Q − t0O ). (1.8)
According to our assumption, the left sides of the two equations are equal. Also,
since the clock at O is synchronized with those at P and Q,
tP − tO = t0R − t0P and tR − tQ = t0Q − t0O .
Thus, subtracting the equations Eq. (1.8) shows that the clocks at P and Q are
synchronized:
tQ − tP = t0P − t0Q .

17
1.3 The Metric Postulate

1.3 The Metric Postulate


Let P and Q be points in a flat surface. Different curves between the points have
different lengths. But surface dwellers single out for special study the length ∆s
of the straight line between P and Q. They call ∆s the proper distance between
the points.
The proper distance ∆s between two points is defined geometrically, inde-
pendently of any planar frame. But there is a simple formula for ∆s in terms
of the coordinate differences between the points in a planar frame:
The Metric Postulate for a Flat Surface
Let ∆s be the proper distance between points P and Q. Let P and
Q have coordinate differences (∆x, ∆y) in a planar frame. Then
∆s2 = ∆x2 + ∆y 2 . (1.9)

The coordinate differences ∆x and ∆y be-


tween P and Q are different in different planar
frames. See Fig. 1.7. However, the particular
combination of the differences in Eq. (1.9) will
always produce ∆s. Neither ∆x nor ∆y has a
geometric significance independent of the par-
ticular planar frame chosen. The two of them
together do: they determine ∆s, which has a di-
rect geometric significance, independent of any
coordinate system.
2 2 2
Let E and F be events in a flat spacetime. Fig. 1.7: ∆s = ∆x +∆y
in both planar frames.
There is a distance-like quantity ∆s between
them. It is called the (spacetime) interval between the events. The definition
of ∆s in a flat spacetime is more complicated than in a flat surface, as there are
three ways in which events can be, we say, separated :
• If E and F can be on the worldline of a pulse of light, they are lightlike
separated. Then define ∆s = 0.
• If E and F can be on the worldline of an inertial clock, they are timelike
separated. Then define ∆s to be the time between the events measured
by the clock. This is the proper time between the events. Other clocks
moving between the events will measure different times. But we single out
for special study the proper time ∆s .
• If E and F can be simultaneously at the ends of an inertial rigid rod –
simultaneously in the sense that light flashes emitted at E and F reach
the center of the rod simultaneously, or equivalently, that E and F are
simultaneous in the rest frame of the rod – they are spacelike separated.
Then define |∆s | to be the length the rod. (The reason for the absolute
value will become clear later.) This is the proper distance between the
events.

18
1.3 The Metric Postulate

The spacetime interval ∆s between two events is defined physically, inde-


pendently of any inertial frame. But there is a simple formula for ∆s in terms
of the coordinate differences between the events in an inertial frame:

The Metric Postulate for a Flat Spacetime


Let ∆s be the interval between events E and F . Let the events have
coordinate differences (∆t, ∆x, ∆y, ∆z) in an inertial frame. Then

∆s2 = ∆t2 − ∆x2 − ∆y 2 − ∆z 2 . (1.10)

The coordinate differences between E and F , including the time coordinate


difference, are different in different inertial frames. For example, suppose that
an inertial clock measures a proper time ∆s between two events. In an inertial
frame in which the clock is at rest, ∆t = ∆s and ∆x = ∆y = ∆z = 0. This will
not be the case in an inertial frame in which the clock is moving. However, the
particular combination of the differences in Eq. (1.10) will always produce ∆s.
No one of the coordinate differences has a physical significance independent of
the particular inertial frame chosen. The four of them together do: they deter-
mine ∆s, which has a direct physical significance, independent of any inertial
frame.
This shows that the joining of space and time into spacetime is not an
artificial technical trick. Rather, in the words of Hermann Minkowski, who
introduced the spacetime concept in 1908, “Space by itself, and time by itself,
are doomed to fade away into mere shadows, and only a kind of union of the
two will preserve an independent reality.”
Physical Meaning. We now describe the physical meaning of the metric
postulate for lightlike, timelike, and spacelike separated events. We do not need
the y- and z-coordinates for the discussion, and so we use Eq. (1.10) in the form

∆s2 = ∆t2 − ∆x2 . (1.11)

Lightlike separated events. By definition, a pulse of light can move


between lightlike separated events, and ∆s = 0 for the events. From Eq. (1.11)
the speed of the pulse is |∆x|/∆t = 1 . The metric postulate asserts that the
speed c of light has always the same value c = 1 in all inertial frames. The
important thing is that the speed is the same in all inertial frames; the actual
value c = 1 is then a convention: Choose the distance light travels in one second
– about 3 × 1010 cm – as the unit of distance. Call this one (light) second of
distance. (You are probably familiar with a similar unit of distance – the light
year.) Then 1 cm = 3.3 × 10−11 sec. With this convention c = 1, and all other
speeds are expressed as a fraction of the speed of light. Ordinarily the fractions
are very small.

19
1.3 The Metric Postulate

Timelike separated events. By definition, an inertial clock can move be-


tween timelike separated events, and ∆s is the (proper) time the clock measures
between the events. The speed of the clock is v = |∆x|/∆t . Then from Eq.
(1.11), the proper time is
1 1 1
∆s = (∆t2 − ∆x2 ) 2 = [1 − (∆x/∆t)2 ] 2 ∆t = (1 − v 2 ) 2 ∆t. (1.12)

The metric postulate asserts that the proper time between two events is less than
the time determined by the synchronized clocks of an inertial frame: ∆s < ∆t.
Informally, “moving clocks run slowly”. This is called time dilation. According
1
to Eq. (1.12) the time dilation factor is (1 − v 2 ) 2 . For normal speeds, v is very
2
small, v is even smaller, and so from Eq. (1.12), ∆s ≈ ∆t, as expected. But
1
as v → 1, ∆t/∆s = (1 − v 2 )− 2 → ∞. Fig. 1.8 shows the graph of ∆t/∆s vs. v.
Exercise 1.6. Investigate the
Doppler redshift. Let a source of light
pulses move with velocity v directly
away from an observer at rest in an iner-
tial frame. Let ∆te be the time between
the emission of pulses, and ∆to be the
time between the reception of pulses by
the observer.
a. Show that ∆to = ∆te + v∆te /c .
b. Ignore time dilation in Eq. (1.6)
by setting ∆s = ∆t. Show that z = v/c
in this approximation.
c. Show that the exact formula is
1
z = [(1 + v)/(1 − v)] 2 − 1 (with c = 1). 1
Fig. 1.8: ∆t/∆s = (1 − v 2 )− 2 .
Use the result of part a.

Spacelike separated events. By definition, the ends of an inertial rigid


rod can be simultaneously present at the events, and |∆s | is the length of the
rod. Appendix 4 shows that the speed of the rod in I is v = ∆t/|∆x| . (That is
not a typo.) A calculation similar to Eq. (1.12) gives
1
|∆s | = (1 − v 2 ) 2 |∆x|. (1.13)

The metric postulate asserts that the proper distance between spacelike separated
events is less than an inertial frame distance: |∆s | < |∆x|. (This is not length
contraction, which we discuss in Appendix 5.)
Connections. We have just seen that the physical meaning of the metric
postulate is different for lightlike, timelike, and spacelike separated events. How-
ever, the meanings are connected: the physical meaning for lightlike separated
events (a universal light speed) implies the physical meanings for timelike and
spacelike separated events. We now prove this for the timelike case. The proof
is quite instructive. The spacelike case is less so; it is relegated to Appendix 4.

20
1.3 The Metric Postulate

By definition, an inertial
clock C can move between time-
like separated events E and F ,
and ∆s is the time C measures
between the events. Let C carry
a rod R perpendicular to its di-
rection of motion. Let R have
a mirror M on its end. At E a Fig. 1.9: ∆s2 = ∆t2 − ∆x2 for timelike
light pulse is sent along R from separated events. e and f are the points at
C. The length of R is chosen so which events E and F occur.
that the pulse is reflected by M
back to F . Fig. 1.9 shows the path of the light in I, together with C, R, and M
as the light reflects off M .
Refer to the rightmost triangle in Fig. 1.9. In I, the distance between E and
F is ∆x. This gives the labeling of the base of the triangle. In I, the light takes
the time ∆t from E to M to F . Since c = 1 in I, the light travels a distance
∆t in I. This gives the labeling of the hypotenuse. C is at rest in some inertial
frame I 0 . In I 0 , the light travels the length of the rod twice in the proper time
∆s between E and F measured by C. Since c = 1 in I 0 , the length of the rod
is 21 ∆s in I 0 . This gives the labeling of the altitude of the triangle. (There is a
tacit assumption here that the length of R is the same in I and I 0 . Appendix 5
discusses this.) Applying the Pythagorean theorem to the triangle shows that
Eq. (1.11) is satisfied for timelike separated events.
In short, since the light travels farther in I than in I 0 (the hypotenuse twice
vs. the altitude twice) and the speed c = 1 is the same in I and I 0 , the time (=
distance/speed) between E and F is longer in I than for C. This shows, in a
most graphic way, that accepting a universal light speed forces us to abandon a
universal time.
The argument shows how it is possible for a single pulse of light to have the
same speed in inertial frames moving with respect to each other: the speed (=
distance/time) can be the same because the distance and the time are different
in the two frames.
Exercise 1.7. Criticize the following argument. We have just seen that
the time between two events is greater in I than in I 0 . But exactly the same
argument carried out in I 0 will show that the time between the events is greater
in I 0 than in I. This is a contradiction.
Local Forms. The metric postulate for a planar frame, Eq. (1.9), gives only
the distance along a straight line between two points. The differential version
of Eq. (1.9) gives the distance ds between neighboring points along any curve:

The Metric Postulate for a Flat Surface, Local Form


Let P and Q be neighboring points. Let ds be the distance between
them. Let the points have coordinate differences (dx, dy) in a planar
frame. Then
ds2 = dx2 + dy 2 .

21
1.3 The Metric Postulate

Thus, if a curve is parameterized (x(p), y(p)), a ≤ p ≤ b, then


"   2 #
2
dx dy
ds2 = + dp2 ,
dp dp

and the length of the curve is


Z b Z b
" 2  2 # 12
dx dy
s= ds = + dp .
p=a p=a dp dp

Of course, different curves between two points can have different lengths.
Exercise 1.8. Calculate the circumference of the circle x = r cos θ, y =
r sin θ, 0 ≤ θ ≤ 2π.

The metric postulate for an inertial frame Eq. (1.10) is concerned only with
times measured by inertial clocks. The differential version of Eq. (1.10) gives
the time ds measured by any clock between neighboring events on its worldline:

The Metric Postulate for a Flat Spacetime, Local Form


Let E and F be neighboring events. If E and F are lightlike sep-
arated, let ds = 0. If the events are timelike separated, let ds be
the time between them as measured by any (inertial or noninertial)
clock. Let the events have coordinate differences (dt, dx, dy, dz) in
an inertial frame. Then

ds2 = dt2 − dx2 − dy 2 − dz 2 . (1.14)

From Eq. (1.14), if the worldline of a clock is parameterized

(t(p), x(p), y(p), z(p)), a ≤ p ≤ b,

then the time s to traverse the worldline, as measured by the clock, is


Z b Z b
" 2  2  2  2 # 21
dt dx dy dz
s= ds = − − − dp.
p=a p=a dp dp dp dp

In general, clocks on different worldlines between two events will measure dif-
ferent times between the events.
Exercise 1.9. Let a clock move between two events with a time difference
∆t. Let v be the small constant speed of the clock. Show that ∆t−∆s ≈ 21 v 2 ∆t.
Exercise 1.10. Consider a simplified Hafele-Keating experiment. One
clock remains on the ground and the other circles the equator in an airplane to
the west – opposite to the Earth’s rotation. Assume that the Earth is spinning
on its axis at one revolution per 24 hours in an inertial frame. (Thus the clock
on the ground is not at rest.) Notation: ∆t is the duration of the trip in the

22
1.3 The Metric Postulate

inertial frame, vr is the velocity of the clock remaining on the ground, and ∆sr
is the time it measures for the trip. The quantities va and ∆sa are defined
similarly for the airplane.
Use Exercise 1.9 for each clock to show that the difference between the
clocks due to time dilation is ∆sa − ∆sr = 21 (va2 − vr2 )∆t. Suppose that ∆t = 40
hours and the speed of the airplane with respect to the ground is 1000 km/hr.
Substitute values to obtain ∆sa − ∆sr = 1.4 × 10−7 s.

Experimental Evidence. Because general relativistic effects play a part in


the Hafele-Keating experiment (see Exercise 2.1), and because the uncertainty
of the experiment is large (±10%), this experiment is not a precision test of time
dilation for clocks. Much better evidence comes from observations of subatomic
particles called muons. When at rest the average lifetime of a muon is 3 × 10−6
sec. According to the differential version of Eq. (1.12), if the muon is moving
in a circle with constant speed v, then its average life, as measured in the
1
laboratory, should be larger by a factor (1 − v 2 )− 2 . An experiment performed
in 1977 showed this within an experimental error of .2%. In the experiment
1
v = .9994, giving a time dilation factor ∆t/∆s = (1 − v 2 )− 2 = 29! The circular
motion had an acceleration of 1021 cm/sec2 , and so this is a test of the local
form Eq. (1.14) of the metric postulate as well as the original form Eq. (1.10).
There is excellent evidence for a universal light speed. First of all, realize
that if clocks at P and Q are synchronized according to the definition Eq.
(1.2), then the speed of light from P to Q is equal to the speed from Q to P .
We emphasize that with our definition of synchronized clocks this equality is a
matter of definition which can be neither confirmed nor refuted by experiment.
The speed c of light can be measured by sending a pulse of light from a point
P to a mirror at a point Q at distance D and measuring the elapsed time 2T
for it to return. Then c = 2D/2T ; c is a two way speed, measured with a single
clock. Equation 1.5 shows that this two way speed is equal to the one way speed
from P to Q. Thus the one way speed of light can be measured by measuring
the two way speed.
In a famous experiment performed in 1887, A. A. Michelson and E. W.
Morley compared the two way speed of light in perpendicular directions from a
given point. Their experiment has been repeated many times, most accurately
by G. Joos in 1930, who found that any difference in the two way speeds is
less than six parts in 1012 . The Michelson-Morley experiment is described in
Appendix 6. A modern version of the experiment using lasers was performed
in 1979 by A. Brillit and J. L. Hall. They found that any difference in the two
way speed of light in perpendicular directions is less than four parts in 1015 .
See Appendix 6.
Another experiment, performed by R. J. Kennedy and E. M. Thorndike in
1932, found the two way speed of light to be the same, within six parts in 109 ,
on days six months apart, during which time the Earth moved to the opposite
side of its orbit. See Appendix 6. Inertial frames in which the Earth is at rest
on days six months apart move with a relative speed of 60 km/sec (twice the
Earth’s orbital speed). A more recent experiment by D. Hils and Hall improved

23
1.3 The Metric Postulate

the result by over two orders of magnitude.


These experiments provide good evidence that the two way speed of light
is the same in different directions, places, and inertial frames and at different
times. They thus provide strong motivation for our definition of synchronized
clocks: If the two way speed of light has always the same value, what could be
more natural than to define synchronized clocks by requiring that the one way
speed have this value?
In all of the above experiments, the source of the light is at rest in the
inertial frame in which the light speed is measured. If light were like baseballs,
then the speed of a moving source would be imparted to the speed of light it
emits. Strong evidence that this is not so comes from observations of certain
neutron stars which are members of a binary star system and which emit X-ray
pulses at regular intervals. These systems are described in Sec. 3.1. If the
speed of the star were imparted to the speed of the X-rays, then various strange
effects would be observed. For example, X-rays emitted when the neutron star
is moving toward the Earth could catch up with those emitted earlier when it
was moving away from the Earth, and the neutron star could be seen coming
and going at the same time! See Fig. 1.10.
This does not happen; an
analysis of the arrival times
of the pulses at Earth made
in 1977 by K. Brecher shows
that no more than two parts
in 109 of the speed of the
source is added to the speed
of the X-rays. (It is not
possible to “see” the neutron
star in orbit around its com- Fig. 1.10: The speed of light is independent
panion directly. The speed of the speed of its source.
of the neutron star toward or
away from the Earth can be determined from the Doppler redshift of the time
between pulses. See Exercise 1.6.)
Finally, recall from above that the universal light speed statement of the
metric postulate implies the statements about timelike and spacelike separated
events. Thus the evidence for a universal light speed is also evidence for the
other two statements.

The universal nature of the speed of light makes possible the modern defini-
tion of the unit of length: “The meter is the length of the path traveled by light
during the time interval of 1/299,792,458 of a second.” Thus, by definition, the
speed of light is 299,792,458 m/sec.

24
1.4 The Geodesic Postulate

1.4 The Geodesic Postulate


We will find it convenient to use superscripts to distinguish coordinates. Thus
we use (x1 , x2 ) instead of (x, y) for planar frame coordinates.

Fig. 1.11: A geodesic in a planar frame.

The line in Fig. 1.11 can be parameterized by the (proper) distance s from
(b1 , b2 ) to (x1 , x2 ): x1 (s) = (cos θ)s+b1 , x2 (s) = (sin θ)s+b2 . Differentiate twice
with respect to s to obtain

The Geodesic Postulate for a Flat Surface


Parameterize a straight line with arclength s . Then in every planar
frame
ẍi = 0, i = 1, 2. (1.15)

The straight lines are called geodesics.


Not all parameterizations of a straight line satisfy the geodesic differential
equations Eq. (1.15). Example: xi (p) = ai p3 + bi .

25
1.4 The Geodesic Postulate

We will find it convenient to use (x0 , x1 , x2 , x3 ) instead of (t, x, y, z) for


inertial frame coordinates. Our third postulate for special relativity says that
inertial particles and light pulses move in a straight line at constant speed in an
inertial frame, i.e., their equations of motion are

xi = ai x0 + bi , i = 1, 2, 3. (1.16)

(Differentiate to give dxi /dx0 = ai ; the velocity is constant.) For inertial parti-
cles this is called Newton’s first law.
Set x0 = p, a parameter; a0 = 1; and b0 = 0, and find that worldlines of
inertial particles and light can be parameterized

xi (p) = ai p + bi , i = 0, 1, 2, 3 (1.17)

in an inertial frame. Eq. (1.17), unlike Eq. (1.16), is symmetric in all four
coordinates of the inertial frame. Also, Eq. (1.17) shows that the worldline is
a straight line in the spacetime. Thus “straight in spacetime” includes both
“straight in space” and “straight in time” (constant speed). See Exercise 1.1.
The worldlines are called geodesics.
Exercise 1.11. In Eq. (1.17) the parameter p = x0 . Show that the
worldline of an inertial particle can also parameterized with s, the proper time
along the worldline.

The Geodesic Postulate for a Flat Spacetime


Worldlines of inertial particles and pulses of light can be parameter-
ized with a parameter p so that in every inertial frame

ẍi = 0, i = 0, 1, 2, 3. (1.18)

For inertial particles we may take p = s.

The geodesic postulate is a mathematical expression of our physical assertion


that in an inertial frame inertial particles and light move in a straight line at
constant speed.
Exercise 1.12. Make as long a list as you can of analogous properties of
flat surfaces and flat spacetimes.

26
Chapter 2

Curved Spacetimes

2.1 History of Theories of Gravity


Recall the analogy from Chapter 1: A curved spacetime is to a flat spacetime
as a curved surface is to a flat surface. We explored the analogy between a flat
surface and a flat spacetime in Chapter 1. In this chapter we generalize from flat
surfaces and flat spacetimes (spacetimes without significant gravity) to curved
surfaces and curved spacetimes (spacetimes with significant gravity). General
relativity interprets gravity as a curvature of spacetime.
Before embarking on a study of grav-
ity in general relativity let us review, very
briefly, the history of theories of gravity.
These theories played a central role in the
rise of science. Theories of gravity have their
roots in attempts of the ancients to predict
the motion of the planets. The position of
a planet with respect to the stars changes
from night to night, sometimes exhibiting Fig. 2.1: The position of a
a “loop” motion, as in Fig. 2.1. The word planet ( ◦ ) with respect to the
“planet” is from the Greek “planetai”: wan- stars changes nightly.
derers.
In the second century, Claudius Ptolemy devised a scheme to explain these
motions. Ptolemy placed the Earth near the center of the universe with a planet
moving on a small circle, called an epicycle, while the center of the epicycle
moves (not at a uniform speed) on a larger circle, the deferent, around the
Earth. See Fig. 2.2. By appropriately choosing the radii of the epicycle and
deferent, as well as the speeds involved, Ptolemy was able to reproduce, with
fair accuracy, the motions of the planets. This remarkable but cumbersome
theory was accepted for over 1000 years.

27
2.1 History of Theories of Gravity

In 1543 Nicholas Copernicus published a theory which was to revolutionize


science and our perception of our place in the universe: he placed the Sun near
the center of our planetary system, with the planets orbiting the Sun in circles
at uniform speed.
In 1609 Johannes Kepler published a
theory which is more accurate than Coper-
nicus’: the path of a planet is an ellipse
with the Sun at one focus. At about the
same time Galileo Galilei was investigating
the acceleration of objects near the Earth’s
surface. He found two things of interest for
us: the acceleration is constant in time and
independent of the mass and composition of
the falling object.
Fig. 2.2: Ptolemy’s theory of
In 1687 Isaac Newton published a the-
planetary motion.
ory of gravity which explained Kepler’s as-
tronomical and Galileo’s terrestrial findings as manifestations of the same phe-
nomenon: gravity. To understand how orbital motion is related to falling mo-
tion, refer to Fig. 2.3. The curves A, B, C are the paths of objects leaving
the top of a tower with greater and greater horizontal velocities. They hit the
ground farther and farther from the bottom of the tower until C when the object
goes into orbit.
Mathematically, Newton’s theory says that a planet
in the Sun’s gravity or an apple in the Earth’s gravity is
pulled instantaneously by the central body (do not ask
how!), causing an acceleration

κM
a=− , (2.1)
r2

where κ is the Newtonian gravitational constant, M is


the mass of the central body, and r is the distance to the
center of the central body. Eq. (2.1) implies that the Fig. 2.3: Falling
planets orbit the Sun in ellipses, in accord with Kepler’s and orbital mo-
findings. See Appendix 7. By taking the distance r tion are the same.
to the Earth’s center to be sensibly constant near the
Earth’s surface, we see that Eq. (2.1) is also in accord with Galileo’s findings:
a is constant in time and is independent of the mass and composition of the
falling object.
Newton’s theory has enjoyed enormous success. A spectacular example oc-
curred in 1846. Observations of the position of the planet Uranus disagreed with
the predictions of Newton’s theory of gravity, even after taking into account the
gravitational effects of the other known planets. The discrepancy was about 4
arcminutes – 1/8th of the angular diameter of the moon. U. Le Verrier, a French
mathematician, calculated that a new planet, beyond Uranus, could account for
the discrepancy. He wrote J. Galle, an astronomer at the Berlin observatory,

28
2.1 History of Theories of Gravity

telling him where the new planet should be – and Neptune was discovered! It
was within 1 arcdegree of Le Verrier’s prediction.
Even today, calculations of spacecraft trajectories are made using Newton’s
theory. The incredible accuracy of his theory will be examined further in Sec.
3.3.
Nevertheless, Einstein rejected Newton’s theory because it is based on pre-
relativity ideas about time and space which, as we have seen, are not correct.
For example, the acceleration in Eq. (2.1) is instantaneous with respect to a
universal time.

29
2.2 The Key to General Relativity

2.2 The Key to General Relativity


A curved surface is different from a flat surface. However, a simple observation
by the nineteenth century mathematician Karl Friedrich Gauss provides the key
to the construction of the theory of surfaces: a small region of a curved surface
is very much like a small region of a flat surface. This is familiar: a small
region of a (perfectly) spherical Earth appears flat. This is why it took so long
to discover that it is not flat. On an apple, a smaller region must be chosen
before it appears flat.
In the next three sections we shall formalize Gauss’ observation by taking
the three postulates for flat surfaces from Chapter 1, restricting them to small
regions, and then using them as postulates for curved surfaces.
We shall see that a curved spacetime is different from a flat spacetime.
However, a simple observation of Einstein provides the key to the construction
of general relativity: a small region of a curved spacetime is very much like a
small region of a flat spacetime. To understand this, we must extend the concept
of an inertial object to curved spacetimes.
Passengers in an airplane at rest on the ground or flying in a straight line
at constant speed feel gravity, which feels very much like an inertial force. Ac-
celerometers respond to the gravity. On the other hand, astronauts in orbit or
falling radially toward the Earth feel no inertial forces, even though they are
not moving in a straight line at constant speed with respect to the Earth. And
an accelerometer carried by the astronauts will register zero. We shall include
gravity as an inertial force and, as in special relativity, call an object inertial
if it experiences no inertial forces. An inertial object in gravity is in free fall.
Forces other than gravity do not act on it.
We now rephrase key Einstein’s observation: as viewed by inertial observers,
a small region of a curved spacetime is very much like a small region of a flat
spacetime. We see this vividly in motion pictures of astronauts in orbit. No
gravity is apparent in their cabin; objects suspended at rest remain at rest.
Inertial objects in the cabin move in a straight line at constant speed, just as
in a flat spacetime. The Newtonian theory predicts this: according to Eq. (2.1)
an inertial object and the cabin accelerate the same with respect to the Earth
and so they do not accelerate with respect to each other.
In the next three sections we shall formalize Einstein’s observation by tak-
ing our three postulates for flat spacetimes, restricting them to small spacetime
regions, and then using them as our first three (of four) postulates for curved
spacetimes. The local inertial frame postulate asserts the existence of small in-
ertial cubical lattices with synchronized clocks to serve as coordinate systems
in small regions of a curved spacetime. The metric postulate asserts a universal
light speed and a slowing of moving clocks in local inertial frames. The geodesic
postulate asserts that inertial particles and light move in a straight line at con-
stant speed in local inertial frames. We first discuss experimental evidence for
the three postulates.

30
2.2 The Key to General Relativity

Experiments of R. Dicke and of V. B. Braginsky, performed in the 1960’s,


verify to extraordinary accuracy Galileo’s finding incorporated into Newton’s
theory: the acceleration of a free falling object in gravity is independent of its
mass and composition. (See Sec. 2.1.) We may reformulate this in the language
of spacetimes: the worldline of an inertial object in a curved spacetime is inde-
pendent of its mass and composition. The geodesic postulate will incorporate
this by not referring to the mass or composition of the inertial objects whose
worldlines it describes.
Dicke and Braginsky used the Sun’s
gravity. We can understand the principle
of their experiments from the simplified di-
agram in Fig. 2.4. The weights A and B,
supported by a quartz fiber, are, with the
Earth, in free fall around the Sun. Dicke
and Braginsky used various substances with
various properties for the weights. Any dif-
ference in their acceleration toward the Sun
would cause a twisting of the fiber. Due to Fig. 2.4: Masses A and B ac-
the Earth’s rotation, the twisting would be celerate the same toward the
in the opposite direction twelve hours later. Sun.
The apparatus had a resonant period of os-
cillation of 24 hours so that oscillations could build up. In Braginsky’s exper-
iment the difference in the acceleration of the weights toward the Sun was no
more than one part in 1012 of their mutual acceleration toward the Sun. A
planned satellite test (STEP) will test the equality of accelerations to one part
in 1018 .
A related experiment shows that the Earth and the Moon, despite the huge
difference in their masses, accelerate the same in the Sun’s gravity. If this were
not so, then there would be unexpected changes in the Earth-Moon distance.
Changes in this distance can be measured within 2 cm (!) by timing the return
of a laser pulse sent from Earth to mirrors on the Moon left by astronauts.
This is part of the lunar laser experiment. The measurements show that the
relative acceleration of the Earth and Moon is no more than a part in 1013 of
their mutual acceleration toward the Sun. The experiment also shows that the
Newtonian gravitational constant κ in Eq. (2.1) does not change by more than
1 part in 1012 per year. The constant also appears in Einstein’s field equation,
Eq. (2.27).
There is another difference between the Earth and the Moon which might
cause a difference in their acceleration toward the Sun. Imagine disassembling
to Earth into small pieces and separating the pieces far apart. The separation
requires energy input to counteract the gravitational attraction of the pieces.
This energy is called the Earth’s gravitational binding energy . By Einstein’s
principle of the equivalence of mass and energy (E = mc2 ), this energy is
equivalent to mass. Thus the separated pieces have more total mass than the
Earth. The difference is small – 4.6 parts in 1010 . But it is 23 times smaller
for the Moon. One can wonder whether this difference between the Earth and

31
2.2 The Key to General Relativity

Moon causes a difference in their acceleration toward the Sun. The lunar laser
experiment shows that this does not happen. This is something that the Dicke
and Braginsky experiments cannot test.
The last experiment we shall consider as evidence for the three postulates is
the terrestrial redshift experiment. It was first performed by R. V. Pound and G.
A. Rebka in 1960 and then more accurately by Pound and J. L. Snider in 1964.
The experimenters put a source of gamma radiation at the bottom of a tower.
Radiation received at the top of the tower was redshifted: z = 2.5 × 10−15 ,
within an experimental error of about 1%. This is a gravitational redshift.
According to the discussion following Eq. (1.6), an observer at the top of
the tower would see a clock at the bottom run slowly. Clocks at rest at different
heights in the Earth’s gravity run at different rates! Part of the result of the
Hafele-Keating experiment is due to this. See Exercise 2.1.
We showed in Sec. 1.3 that the assumption Eq. (1.4), necessary for synchro-
nizing clocks at rest in the coordinate lattice of an inertial frame, is equivalent
to a zero redshift between the clocks. This assumption fails for clocks at the
top and bottom of the tower. Thus clocks at rest in a small coordinate lattice
on the ground cannot be (exactly) synchronized.
We now show that the experiment provides evidence that clocks at rest in
a small inertial lattice can be synchronized. In the experiment, the tower has
(upward) acceleration g, the acceleration of Earth’s gravity, in a small inertial
lattice falling radially toward Earth. We will show shortly that the same redshift
would be observed with a tower having acceleration g in an inertial frame in
a flat spacetime. This is another example of small regions of flat and curved
spacetimes being alike. Thus it is reasonable to assume that there would be
no redshift with a tower at rest in a small inertial lattice in gravity, just as
with a tower at rest in an inertial frame. (It is desirable to test this directly
by performing the experiment in orbit.) In this way, the experiment provides
evidence that the condition Eq. (1.4), necessary for clock synchronization, is
valid for clocks at rest in a small inertial lattice. Loosely speaking, we may say
that since light behaves “properly” in a small inertial lattice, light accelerates
the same as matter in gravity.
We now calculate the Doppler redshift for a tower with acceleration g in an
inertial frame. Suppose the tower is momentarily at rest when gamma radiation
is emitted. The radiation travels a distance h, the height of the tower, in the
inertial frame. (We ignore the small distance the tower moves during the flight
of the radiation. We shall also ignore the time dilation of clocks in the moving
tower and the length contraction – see Appendix 5 – of the tower. These effects
are far too small to be detected by the experiment.) Thus the radiation takes
time t = h/c to reach the top of the tower. (For clarity we do not take c = 1.)
In this time the tower acquires a speed v = gt = gh/c in the inertial frame.
From Exercise 1.6, this speed causes a Doppler redshift
v gh
z= = 2. (2.2)
c c
In the experiment, h = 2250 cm. Substituting numerical values in Eq.

32
2.2 The Key to General Relativity

(2.2) gives the value of z measured in the terrestrial redshift experiment; the
gravitational redshift for towers accelerating in inertial frames is the same as
the Doppler redshift for towers accelerating in small inertial lattices near Earth.
Exercise 3.4 shows that a rigorous calculation in general relativity also gives Eq.
(2.2).
Exercise 2.1. Let h be the height at which the airplane flies in the simplified
Hafele-Keating experiment of Exercise 1.10. Show that the difference between
the clocks due to the gravitational redshift is

∆sa − ∆sr = gh∆t.

Suppose that h = 10 km. Substitute values to obtain ∆sa − ∆sr = 1.6 × 10−7
sec.
Adding this to the time dilation difference of Exercise 1.10 gives a difference
of 3.0 × 10−7 sec. Exercise 3.3 shows that a rigorous calculation in general
relativity gives the same result.

The Global Positioning System (GPS) of satellites must adjust its clocks for
time dilation and gravitational redshifts to function properly. In fact, the effects
are 10,000 times too large to be ignored.

33
2.3 The Local Inertial Frame Postulate

2.3 The Local Inertial Frame Postulate


Suppose that curved surface dwellers attempt to
construct a square coordinate grid using rigid
rods constrained (of course) to their surface. If
the rods are short enough, then at first they will
fit together well. But, owing to the curvature of
the surface, as the grid gets larger the rods must
be forced a bit to connect them. This will cause
stresses in the lattice and it will not be quite
square. Surface dwellers call a small (nearly)
square coordinate grid a local planar frame at P , Fig. 2.5: A local planar
where P is the point at the origin of the grid. In frame.
smaller regions around P , the grid must become
more square. See Fig. 2.5.

The Local Planar Frame Postulate for a Curved Surface


A local planar frame can be constructed at any point P of a curved
surface with any orientation.
We shall see that local planar frames
at P provide surface dwellers with an in-
tuitive description of properties of a curved
surface at P . However, in order to study
the surface as a whole, they need global co-
ordinates, defined over the entire surface.
There are, in general, no natural global co-
ordinate systems to single out in a curved
surface as planar frames are singled out in a
flat surface. Thus they attach global coordi-
Fig. 2.6: Spherical coordi- nates (y 1 , y 2 ) in an arbitrary manner. The
nates (φ, θ) on a sphere. only restrictions are that different points
must have different coordinates and nearby
points must receive nearby coordinates. In general, the coordinates will not
have a geometric meaning; they merely serve to label the points of the surface.
One common way for us (but not surface dwellers) to attach global coordi-
nates to a curved surface is to parameterize it in three dimensional space:

x = x(y 1, y 2 ), y = y(y 1, y 2 ), z = z(y 1, y 2 ) . (2.3)

As (y 1, y 2 ) varies, (x, y, z) varies on the surface. Assign coordinates (y 1, y 2 ) to


the point (x, y, z) on the surface given by Eq. (2.3). For example, Fig. 2.6 shows
spherical coordinates (y 1, y 2 ) = (φ, θ) on a sphere of radius R. The coordinates
are assigned by the usual parameterization

x = R sin φ cos θ, y = R sin φ sin θ, z = R cos φ. (2.4)

34
2.3 The Local Inertial Frame Postulate

The Global Coordinate Postulate for a Curved Surface


The points of a curved surface can be labeled with coordinates
(y 1 , y 2 ).

(Technically, the postulate should state that a curved surface is a two dimen-
sional manifold . The statement given will suffice for us.)
In the last section we saw that inertial objects in an astronaut’s cabin behave
as if no gravity were present. Actually, they will not behave ex:actly as if no
gravity were present. To see this, assume for simplicity that their cabin is falling
radially toward Earth. Inertial objects in the cabin do not accelerate exactly the
same with respect to the Earth because they are at slightly different distances
and directions from the Earth’s center. See Fig. 2.7. Thus, an object initially
at rest near the top of the cabin will slowly separate from one initially at rest
near the bottom. In addition, two objects initially at rest at the same height
will slowly move toward each other as they both fall toward the center of the
Earth. These changes in velocity are called tidal accelerations. (Why?) They
are caused by small differences in the Earth’s gravity at different places in the
cabin. They become smaller in smaller regions of space and time, i.e., in smaller
regions of spacetime.
Suppose that astronauts in a curved spacetime at-
tempt to construct an inertial cubical lattice using rigid
rods. If the rods are short enough, then at first they will
fit together well. But as the grid gets larger, the lattice
will have to resist tidal accelerations, and the rods can-
not all be inertial. This will cause stresses in the lattice
and it will not be quite cubical.
In the last section we saw that the terrestrial redshift
experiment provides evidence that clocks in a small iner-
Fig. 2.7: Tidal
tial lattice can be synchronized. Actually, due to small
accelerations in
differences in the gravity at different places in the lattice,
a radially free
an attempt to synchronize the clocks with the one at the falling cabin.
origin with the procedure of Sec. 1.3 will not quite work.
However, we can hope that the procedure will work with as small an error as
desired by restricting the lattice to a small enough region of a spacetime.
A small (nearly) cubical lattice with (nearly) synchronized clocks is called a
local inertial frame at E, where E is the event at the origin of the lattice when
the clock there reads zero. In smaller regions around E, the lattice is more
cubical and the clocks are more nearly synchronized. Local inertial frames are
in free fall.

The Local Inertial Frame Postulate for a Curved Spacetime


A local inertial frame can be constructed at any event E of a curved
spacetime, with any orientation, and with any inertial object at E
instantaneously at rest in it.

35
2.3 The Local Inertial Frame Postulate

We shall find that local inertial frames at E provide an intuitive description


of properties of a curved spacetime at E. However, in order to study a curved
spacetime as a whole, we need global coordinates, defined over the entire space-
time. There are, in general, no natural global coordinates to single out in a
curved spacetime, as inertial frames were singled out in a flat spacetime. Thus
we attach global coordinates in an arbitrary manner. The only restrictions are
that different events must receive different coordinates and nearby events must
receive nearby coordinates. In general, the coordinates will not have a physical
meaning; they merely serve to label the events of the spacetime.
Often one of the coordinates is a “time” coordinate and the other three are
“space” coordinates, but this is not necessary. For example, in a flat spacetime
it is sometimes useful to replace the coordinates (t, x) with (t + x, t − x).

The Global Coordinate Postulate for a Curved Spacetime


The events of a curved spacetime can be labeled with coordinates
(y 0 , y 1 , y 2 , y 3 ).

In the next two sections we give the metric and geodesic postulates of gen-
eral relativity. We first express the postulates in local inertial frames. This
local form of the postulates gives them the same physical meaning as in special
relativity. We then translate the postulates to global coordinates. This global
form of the postulates is unintuitive and complicated but is necessary to carry
out calculations in the theory.
We can use arbitrary global coordinates in flat as well as curved spacetimes.
We can then put the metric and geodesic postulates of special relativity in the
same global form that we shall obtain for these postulates for curved space-
times. We do not usually use arbitrary coordinates in flat spacetimes because
inertial frames are so much easier to use. We do not have this luxury in curved
spacetimes.
It is remarkable that we shall be able to describe curved spacetimes intrinsi-
cally, i.e., without describing them as curved in a higher dimensional flat space.
Gauss created the mathematics necessary to describe curved surfaces intrinsi-
cally in 1827. G. B. Riemann generalized Gauss’ mathematics to curved spaces
of higher dimension in 1854. His work was extended by several mathematicians.
Thus the mathematics necessary to describe curved spacetimes intrinsically was
waiting for Einstein when he needed it.

36
2.4 The Metric Postulate

2.4 The Metric Postulate


Fig. 2.5 shows that local planar frames provide curved surface dwellers with a
convenient way to express infinitesimal distances on a curved surface.

The Metric Postulate for a Curved Surface, Local Form


Let point Q have coordinates (dx1 , dx2 ) in a local planar frame at
P . Let ds be the distance between the points. Then

ds2 = (dx1 )2 + (dx2 )2 . (2.5)

Even though the local planar frame extends a finite distance from P , Eq. (2.5)
holds only for infinitesimal distances from P .
We now express Eq. (2.5) in terms of global coordinates. Set the matrix
 
◦ ◦ 1 0
f = (fmn ) = .
0 1

Then Eq. (2.5) can be written


2
X
2 ◦
ds = fmn dxm dxn . (2.6)
m,n=1

Henceforth we use the Einstein summation convention by which an index which


appears twice in a term is summed without using a Σ. Thus Eq. (2.6) becomes

ds2 = fmn dxm dxn . (2.7)

As another example of the summation convention, consider a function f (y 1 , y 2 ).


P2
Then we may write the differential df = i=1 (∂f /∂y i ) dy i = (∂f /∂y i ) dy i .
Let P and Q be neighboring points on a curved surface with coordinates
(y 1 , y 2 ) and (y 1 + dy 1 , y 2 + dy 2 ) in a global coordinate system. Let Q have
coordinates (dx1 , dx2 ) in a local planar frame at P . We may think of the (xi )
coordinates as functions of the (y j ) coordinates, just as cartesian coordinates in
the plane are functions of polar coordinates: x = r cos θ, y = r sin θ. This gives
meaning to the partial derivatives ∂xi/∂y j . From Eq. (2.7), the distance from
P to Q is


ds2 = fmn dxm dxn
 m  n 
◦ ∂x j ∂x k
= fmn dy dy (sum on m, n, j, k)
∂y j ∂y k
m n
 
◦ ∂x ∂x
= fmn dy j dy k
∂y j ∂y k
= gjk (y) dy j dy k , (2.8)

37
2.4 The Metric Postulate

where we have set


◦ ∂xm ∂xn
gjk (y) = fmn . (2.9)
∂y j ∂y k

Since (fmn ) is symmetric, so is (gjk ). Use a local planar frame at each point of
the surface in this manner to translate the local form of the metric postulate,
Eq. (2.5), to global coordinates:

The Metric Postulate for a Curved Surface, Global Form


Let (y i ) be global coordinates on the surface. Let ds be the distance
between neighboring points (y i ) and (y i + dy i ). Then there is a
symmetric matrix (gjk (y i )) such that

ds2 = gjk (y i ) dy j dy k . (2.10)

(gjk (y i )) is called metric of the surface with respect to the coordinates (y i ).


Exercise 2.2. Show that the metric for the (φ, θ) coordinates on the sphere
in Eq. (2.4) is ds2 = R2 dφ2 + R2 sin2 φ dθ2 , i.e.,
 2 
R 0
g(φ, θ) = . (2.11)
0 R2 sin2 φ

Do this in two ways:


a. By converting from the metric of a local planar frame. Show that for
a local planar frame whose x1 -axis coincides with a circle of latitude, dx1 =
R sinφ dθ and dx2 = Rdφ. See Fig. 2.10.
b. Use Eq. (2.4) to convert ds2 = dx2 + dy 2 + dz 2 to (φ, θ) coordinates.
1
Exercise 2.3. Consider the hemisphere z = (R2 − x2 − y 2 ) 2 . Assign
coordinates (x, y) to the point (x, y, z) on the hemisphere. Find the metric in this
coordinate system. Express your answer as a matrix. Hint: Use z 2 = R2 −x2 −y 2
to compute dz 2 .

We should not think of a vector as its components (vi ) , but as a single object
v which represents a magnitude and direction (an arrow). In a given coordinate
system the vector acquires components. The components will be different in
different coordinate systems.
Similarly, we should not think of the metric as its components (gjk ) , but as
a single object g which represents infinitesimal distances. In a given coordinate
system the metric acquires components. The components will be different in
different coordinate systems.
Exercise 2.4. Let (y i ) and (ȳ i ) be two coordinate systems on the same
surface, with metrics (gjk (y i )) and (ḡpq (ȳ i )). Show that

∂y j ∂y k
ḡpq = gjk . (2.12)
∂ ȳ p ∂ ȳ q
Hint: See Eq. (2.8).

38
2.4 The Metric Postulate

The Metric Postulate for a Curved Spacetime, Local Form


Let event F have coordinates (dxi ) in a local inertial frame at E.
If E and F are lightlike separated, let ds = 0. If the events are
timelike separated, let ds be the time between them as measured by
any (inertial or noninertial) clock. Then

ds2 = (dx0 )2 − (dx1 )2 − (dx2 )2 − (dx3 )2 . (2.13)

The metric postulate asserts a universal light speed and a formula for proper
time in local inertial frames. (See the discussion following Eq. (1.11).) This
is another instance of the key to general relativity: a small region of a curved
spacetime is very much like a small region of a flat spacetime.
We now translate the metric postulate to global coordinates. Eq. (2.13) can
be written

ds2 = fmn dxm dxn , (2.14)
where
 
1 0 0 0
 0 −1 0 0 
f ◦ = (fmn

)=  .
 0 0 −1 0 
0 0 0 −1

Using Eq. (2.14), the calculation Eq. (2.8), which produced the global form
of the metric postulate for curved surfaces, now produces the global form of the
metric postulate for curved spacetimes.

The Metric Postulate for a Curved Spacetime, Global Form


Let (y i ) be global coordinates on the spacetime. If neighboring
events (y i ) and (y i + dy i ) are lightlike separated, let ds = 0. If
the events are timelike separated, let ds be the time between them
as measured by a clock moving between them. Then there is a sym-
metric matrix (gjk (y i )) such that

ds2 = gjk (y i ) dy j dy k . (2.15)

(gjk (y i )) is called the metric of the spacetime with respect to (y i ).

39
2.5 The Geodesic Postulate

2.5 The Geodesic Postulate


Curved surface dwellers find that some curves in their surface are straight in
local planar frames. They call these curves geodesics. Fig. 2.8 shows that the
equator is a geodesic but the other circles of latitude are not. To traverse a
geodesic, a surface dweller need only always walk “straight ahead”. A geodesic
is as straight as possible, given that it is constrained to the surface.
As with the geodesic postulate for flat surfaces Eq. (1.15), we have

The Geodesic Postulate for a Curved Surface, Local Form


Parameterize a geodesic xi (s), where s is arclength. Let point P be
on the geodesic. Then in every local planar frame at P

ẍi (P ) = 0, i = 1, 2. (2.16)

We now translate Eq. (2.16) to global coordinates y to obtain the global


form of the geodesic equations. We first need to know that the metric g = (gij )
has an inverse g−1 = (g jk ).
Exercise 2.5. a. Let the matrix a = (∂xn /∂y k ).
Show that the inverse matrix a−1 = (∂y k /∂xj ).
b. Show that Eq. (2.9) can be written g = at f ◦ a,
where t means “transpose”.
c. Prove that g−1 = a−1 (f ◦ )−1 (a−1 )t .
Introduce the notation ∂k gim = ∂gim /∂y k . Define
the Christoffel symbols:
Fig. 2.8: The equa- Γijk = 1
g im [∂k gjm + ∂j gmk − ∂m gjk ] . (2.17)
2
tor is the only circle
of latitude which is Note that Γijk = Γikj . The Γijk , like the gjk , are
a geodesic.
functions of the coordinates.
Exercise 2.6. Show that for the metric Eq.
(2.11)
Γφθθ = − sinφ cos φ and Γθθφ = Γθφθ = cot φ.
The remaining Christoffel symbols are zero.
You should not try to assign a geometric meaning to the Christoffel symbols;
simply think of them as what appears when the geodesic equations are translated
from their local form Eq. (2.16) (which have an evident geometric meaning) to
their global form (which do not):

The Geodesic Postulate for a Curved Surface, Global Form


Parameterize a geodesic with arclength s. Then in every global
coordinate system

ÿ i + Γijk ẏ j ẏ k = 0, i = 1, 2. (2.18)

40
2.5 The Geodesic Postulate

Appendix 9 translates the local form of the postulate to the global form.
The translation requires an assumption. A local planar frame at P extends to a
finite region around P . Let f = (fmn (x)) represent the metric in this coordinate
system. According to Eq. (2.7), (fmn (P )) = f ◦ . Appendix 8 shows that there
are coordinates, called geodesic coordinates, satisfying this relationship and also

∂i fmn (P ) = 0 (2.19)

for all m, n, i. A function with a zero derivative at a point is not changing much
at the point. In this sense Eq. (2.19) states that f stays close to f ◦ near P .
Since a local planar frame at P is constructed to approximate a planar frame
as closely as possible near P , surface dwellers assume that the metric of a local
planar frame at P satisfies Eq. (2.19).
Exercise 2.7. Show that the metric of Exercise 2.3 satisfies Eq. (2.19) at
(x, y) = (0, 0) .
Exercise 2.8. Show that Eq. (2.18) reduces to Eq. (2.16) for local planar
frames.
Exercise 2.9. Show that the equator is the only circle of latitude which is
a geodesic. Of course, all great circles on a sphere are geodesics. Use the result
of Exercise 2.6. Before using the geodesic equations you must parameterize the
circles with s.

The Geodesic Postulate for a Curved Spacetime, Local Form


Worldlines of inertial particles and pulses of light can be parame-
terized so that if E is on the worldline, then in every local inertial
frame at E
ẍi (E) = 0, i = 0, 1, 2, 3. (2.20)
For inertial particles we may take the parameter to be s .

The worldlines are called geodesics.


The geodesic postulate asserts
that in a local inertial frame iner-
tial particles and light move in a
straight line at constant speed. (See
the remarks following Eq. (1.18).)
The worldline of an inertial particle
or pulse of light in a curved space- Fig. 2.9: The path of an inertial parti-
time is as straight as possible (in cle in a lattice stuck to the Earth and in
space and time – see the remarks an inertial lattice. The dots are at equal
following Eq. (1.17)), given that it time intervals.
is constrained to the spacetime. The geodesic is straight in local inertial frames,
but looks curved when viewed in an “inappropriate” coordinate system. See
Fig. 2.9.
Einstein’s “straightest worldline in a curved spacetime” description is very
different from Newton’s “curved path in a flat space” description.

41
2.5 The Geodesic Postulate

We now assume that the metric of a local inertial frame at E satisfies Eq.
(2.19) for the same reasons as given above for local planar frames. Then Ap-
pendix 9 translates the local form of the geodesic postulate for curved spacetimes
to the global form:

The Geodesic Postulate for a Curved Spacetime,


Global Form
Worldlines of inertial particles and pulses of light can be parameter-
ized so that in every global coordinate system

ÿ i + Γijk ẏ j ẏ k = 0, i = 0, 1, 2, 3. (2.21)

For inertial particles we may take the parameter to be s .

42
2.6 The Field Equation

2.6 The Field Equation


Previous sections of this chapter explored similarities between small regions of
flat and curved surfaces and between small regions of flat and curved spacetimes.
This section explores differences.
The local forms of our curved surface postulates show
that in many ways a small region of a curved surface is
like a small region of a flat surface. Surface dwellers
might suppose that all differences between the regions
vanish as the regions become smaller. This is not so.
To see this, pass geodesics through a point P in every
direction.
Exercise 2.10. Show that there is a unique geodesic
through every point of a curved surface in every direction.
Hint: Use a basic theorem on the existence and unique- Fig. 2.10:
2
ness of solutions of systems of differential equations. K = 1/R on a
sphere.
Connect all the points at distance r from P along the
geodesics, forming a “circle” C of radius r. Let C(r) be
the circumference of the circle. Define the curvature K of the surface at P :
3 2π r − C(r)
K= lim . (2.22)
π r→ 0 r3
Clearly, K = 0 for a flat surface. From Fig. 2.10, C(r) < 2πr for a sphere
and so K ≥ 0. From Fig. 2.10, we find
C(r) = 2πR sin φ = 2πR sin(r/R) = 2πR r/R − (r/R)3 /6 + . . . .
 

A quick calculation shows that K = 1/R2 . The cur-


vature is a difference between regions of a sphere and
a flat surface which does not vanish as the regions
become smaller.
The surface of revolution of Fig. 2.11 is a pseudo-
sphere. The horizontal “circles of latitude” are con-
cave inward and the vertical “lines of longitude” are
concave outward. Thus C(r) > 2πr and so K ≤ 0.
Exercise 2.14 shows that K = −1/R2 , where R is a
constant. Despite the examples, in general K varies
from point to point in a curved surface.
Fig. 2.11: K =
Distances and geodesics are involved in the defi-
2
−1/R on a pseudo- nition of K. But distances determine geodesics: dis-
sphere. tances determine the metric, which determines the
Christoffel symbols Eq. (2.17), which determine the
geodesics Eq. (2.18). Thus we learn an important fact: distances determine K.
Thus K is measurable by surface dwellers.
Exercise 2.11. Show that a map of a region of the Earth must distort
distances. Take the Earth to be perfectly spherical. Make no calculations.

43
2.6 The Field Equation

We can roll a flat piece of paper into a cylinder without distorting distances
on the paper and thus without changing K. Thus K = 0 for the cylinder.
Viewed from the outside, the cylinder is curved, and so K = 0 seems “wrong”.
However, viewed from within (and remember, we are describing curved surfaces
and spacetimes without reference to a higher dimensional space), the rolling
does not distort distances in the paper. Thus surface dwellers could not detect
the curvature seen from the outside. Thus K “should” be zero for the cylinder.
The formula expressing K in terms of distances was given by Gauss. If
g12 = 0, then
n  1 1
  1 1
o
−1 − −
∂i ≡ ∂/∂y i . (2.23)

K = − (g11 g22 ) 2 ∂1 g112 ∂1 g22 2
+ ∂2 g222 ∂2 g11
2

Exercise 2.12. Show that Eq. (2.23) gives K = 1/R2 for a sphere of radius
R. Use Eq. (2.11).
Exercise 2.13. Generate a surface of revolution by rotating the param-
eterized curve y = f (u), z = h(u) about the z-axis. Let (r, θ, z) be cylin-
drical coordinates and parameterize the surface with coordinates (u, θ). Use
ds2 = dr2 + r2 dθ2 + dz 2 to show that the metric is
 02
f + h02 0

.
0 f2
Ru 1
Exercise 2.14. If y = Re−u and z = R 0 (1 − e−2t ) 2 dt, then the surface
of revolution in Exercise 2.13 is the pseudosphere of Fig. 2.11. Show that
K = −1/R2 for the pseudosphere.
Exercise 2.15. If y = 1 and z = u, then the surface of revolution in
Exercise 2.13 is a cylinder. Show that K = 0 for a cylinder using Eq. (2.23).

The local forms of our curved spacetime postulates show that in many re-
spects a small region of a curved spacetime is like a small region of a flat space-
time. We might suppose that all differences between the regions vanish as the
regions become smaller. This is not so.
To see this, refer to Fig. 2.7. Let ∆ r be the small distance between objects
at the top and bottom of the cabin and let ∆ a be the small tidal acceleration
between them. In the curved spacetime of the cabin ∆ a 6= 0 , which is different
from the flat spacetime value ∆ a = 0. But in the cabin ∆ a → 0 as ∆ r → 0 ;
this difference between a curved and flat spacetime does vanish as the regions
becomes smaller. But also in the cabin ∆ a/∆ r 6= 0 , again different from the flat
spacetime ∆ a/∆ r = 0 . This difference does not vanish as the regions become
smaller: in the cabin, using Eq. (2.1), ∆ a/∆ r → da/dr = 2κM/r3 6= 0.
The metric and geodesic postulates describe the behavior of clocks, light,
and inertial particles in a curved (or flat) spacetime. But to apply these pos-
tulates, we must know the metric of the spacetime. Our final postulate for
general relativity, the field equation, determines the metric. Loosely speaking,
the equation determines the “shape” of a spacetime, how it is “curved”.

44
2.6 The Field Equation

We constructed the metric in Sec. 2.4 using local inertial frames. There
is obviously a relationship between the motion of local inertial frames and the
distribution of mass in a curved spacetime. Thus, there is a relationship between
the metric of a spacetime and the distribution of mass in the spacetime. The
field equation gives this relationship. Schematically it reads
   
quantity determined quantity determined
= . (2.24)
by metric by mass/energy

To specify the two sides of this equation, we need several definitions. Define
the Ricci tensor

Rjk = Γptk Γtjp − Γptp Γtjk + ∂k Γpjp − ∂p Γpjk . (2.25)

Don’t panic over this convoluted definition: You need not have a physical un-
derstanding of the Ricci tensor. And while the Rjk are extremely tedious to
calculate by hand, computers can readily calculate them for us.
The Ricci tensor is entirely determined by the metric. We shall see that
it contains information about the curvature K of two dimensional surfaces in
four dimensional spacetime. Like K, it involves second derivatives of the gjk
(because the Γijk involve the first derivatives).
As with the metric g , we will use R to designate the Ricci tensor as a single
object, existing independently of any coordinate system, but which in a given
coordinate system acquires components Rjk .
Define the curvature scalar R = g jk Rjk .
We can now specify the left side of the schematic field equation Eq. (2.24):
the quantity determined by the metric is the Einstein tensor

G = R − 12 R g.

The right side of the field equation is given by the energy-momentum tensor
T. It represents the source of the gravitational field in general relativity. All
forms of matter and energy, including electromagnetic fields, and also pressures,
stresses, and viscosity contribute to T. But for our purposes we need to con-
sider only matter of a special form, called dust. In dust, matter interacts only
gravitationally; there are no pressures, stresses, or viscosity. Gas in interstellar
space which is thin enough so that particle collisions are infrequent is dust.
We now define T for dust. Choose an event E with coordinates (y i ). Let ρ be
the density of the dust at E as measured by an observer moving with the dust.
(Thus ρ is the same in all coordinate systems.) Let ds be the time measured by
the observer between E and a neighboring event on the dust’s worldline with
coordinates (y i + dy i ). Define

dy j dy k
T jk = ρ . (2.26)
ds ds

We can now state our final postulate for general relativity:

45
Exploring the Variety of Random
Documents with Different Content
LECTURE XXVI
GERMINAL SELECTION (continued)
Germinal selection, spontaneous and induced—Climatic forms of
Polyommatus phlæas—Deformities—Excessive augmentation of
variations—Can it lead to the elimination of a species?—Saltatory
variations, copper-beech, weeping trees—Origin of sexual
distinguishing characters—Formation of breeds among domesticated
animals—Degenerate jaws—Human teeth—Short-sightedness—Milk-
glands—Small hands and feet—Ascending variation—Talents, intellect—
Combination of mental endowments—The ultimate roots of heritable
variation—There are only plus- and minus-variations—Relations of the
determinants to their determinates—The play of forces in the
determinant system of the id—Germinal selection inhibited by personal
selection—Objection on the score of the minuteness of the substance
of the germ-plasm.

Hitherto we have derived the variations of the determinants of the


germ-plasm, upon which we based the process of germinal
selection, from chance local fluctuations in nutrition, such as must
occur in an individual id, independently of the nutrition of the other
ids of the same germ-plasm. But there are doubtless also influences
which set up similar nutritive changes in all ids, and by which,
therefore, all homologous determinants, in as far as they are
sensitive to the nutritive change in question, are affected in the
same manner. To this category belong changes in the external
conditions of life, and particularly climatic changes. It is, then,
germinal selection alone which brings about the presence of a
majority of ids with determinants varying in the same direction, and
personal selection has no part in the transformation of the species.
Many years ago I instituted experiments with a small butterfly,
Pararga egeria, and these showed that a heightened temperature so
influenced the pupæ of this form that the butterflies emerged with a
different and deeper yellow ground-colour, similar to that of the
long-known southern variety Meione. More thoroughly decisive,
however, were the experiments on Polyommatus phlæas, the small
'fire-butterfly,' which were carried on in the eighties by Merrifield in
England and by myself almost at the same time. I shall discuss these
later in more detail, and will only say here that this butterfly, whose
range extends from Lapland to Sicily, occurs in two forms, the
southern distinguished by a 'dusting' of deep black from the
northern, in which the wing-surfaces are of a pure red-gold. The
experiments showed that the southern form can be artificially
produced by warmth, and the interpretation must be that the direct
influence of higher temperature affects the quality of the nutritive
fluids in the germ-plasm, and thereby at the same time the
determinants of one or more kinds of wing-scales are caused to vary
in all the ids in the same direction, in such a fashion that they give
rise to black scales instead of the former red-gold ones. It is thus
certain that there are external influences which cause particular
determinants to vary in a particular manner. I call this form of
germinal variation 'induced' germinal selection, and contrast it with
'spontaneous' selection, which is caused, not by extra-germinal
influences, but by the chances of the intra-germinal nutritive
conditions, and which will, therefore, not readily occur at the same
time in all the ids of a germ-plasm, and so will not give rise to
variation of the same kind in the homologous determinants of all the
ids.
The two processes must also be distinguished from each other in
their relation to personal selection, for induced germinal selection
will go on increasing until the maximum of variation corresponding
to the nature of the external influences and of the determinants
concerned is reached. Since all the ids are equally affected and
caused to vary in the same way, personal selection has nothing to
take hold of, and the variation might go on intensifying even if it
should become biologically prejudicial. But it is quite otherwise with
spontaneous germinal selection, which has its roots not in all, but
only in a majority of the ids. Here the variation may go on increasing
by germinal selection alone, but only until it acquires a positive or
negative biological value, that is, until it becomes advantageous or
prejudicial to the life of the individual; then personal selection
intervenes and decides whether it is to go on increasing or not.
Spontaneous germinal selection can therefore only lead to the
general variation of a whole species when it is supplemented by
some external factor such as, especially, the utility of the variation.
This does not imply, however, that indifferent variations of large
amount could not arise through spontaneous germinal selection, but
they would remain confined to a small number of individuals, and
would sooner or later disappear again. The congenital deformities of
Man may in part fall under this category. If, for instance, certain
determinants are, by reason of specially favourable local nutritive
conditions, maintained for a long time in progressive variation, they
will become so strong that the part which they determine will turn
out excessive, perhaps double. Hereditary polydactylism in Man may
perhaps be explained on this principle, and I had already referred it
to the more rapid growth and duplication of certain determinants of
the germ even before formulating the idea of germinal selection. In
this I was at one with the pathologist Ernst Ziegler, who had
designated polydactylism as a germ-variation, and in contrast to
others had not interpreted it in an atavistic sense, as a reversion to
unknown six-fingered ancestors. All excessive or defective hereditary
malformations may be referred to germinal selection alone, that is,
to the long-continued progressive or regressive variation of particular
determinant-groups in a majority of ids.
The fact that, as far as our experience goes, superfluous fingers are
never inherited for more than five generations may be simply
explained, for there has been no reason for the intervention of
personal selection, either in the negative sense, for the six-fingered
state does not threaten life, nor in the positive, since it is not of
advantage. The deformity depends on spontaneous germinal
variation, which must have taken place in a majority of ids or it
would not have become manifest. But such a majority of
'polydactylous' ids is liable to become scattered again in every new
descendant, and to be reduced again into a minority which can no
longer make itself felt by the chances of reducing division and the
admixture of normal ids in amphimixis. A polydactylous race of men
could only arise through the assistance of personal selection; in that
case there would doubtless be just as much chance of success in
breeding a six-fingered race as there was in breeding the crooked-
legged Ancon sheep from a single ram which was malformed in this
manner. Without a gradual setting aside of the germs with normal
ids, that is, without personal selection, such spontaneous
deformities, and indeed all spontaneous variations, must fail of
attaining to permanent mastery.
This must frequently be the case in free nature also, but we shall
have to investigate later on, in the section devoted to the formation
of species, whether external circumstances (inbreeding) may not
also occur which make it possible for spontaneous variations to
become constant breed-characters, even although they remain
neither good nor bad, and are thus not subject to the action of
personal selection.
In general, however, amphigony with its reduction of the ids and its
constant mingling of strange ids will form the corrective to the
deviations which may arise through the processes of selection within
the id, and which lead to excessive or superfluous development of
certain structures, to a complete disturbance of the harmony of the
parts, and ultimately to the elimination of the species.
It must be admitted, however, that Emery was probably right when
he directed attention to the possibility of a 'conflict between
germinal and personal selection.' It is quite conceivable that in cases
of useful variations, that is, of adaptations, the processes of
selection within the germ-plasm may lead to excessive
developments, which personal selection cannot control, because, on
account of their earlier usefulness, they have in the course of a
series of generations and species become fixed not only in a majority
of ids, but in almost all the ids of the collective germ-plasm of the
species. In this case a reversal must be difficult and slow, for the
gathering together of ids with relatively weaker determinants can
only take place slowly, and it is questionable whether the species
would survive long enough for the slow process to take effect. But,
apart from the question of time, such a reduction of an excessive
development would sometimes be quite impossible, for the simple
reason that there is nothing for personal selection to take hold of.
Döderlein has pointed out that many characters go on increasing
through whole series of extinct species, and ultimately grow to such
excess that they bring about the destruction of the species, as, for
instance, the antlers of the giant stag or the sabre-like teeth of
certain carnivores in the diluvial period. I shall have to discuss this in
more detail in speaking of the extinction of species; it is enough to
say here that such long-continued augmentations in the same
direction can never be referred solely to germinal selection, since it
is hardly conceivable that a species—much less a whole series of
species—should arise with injurious characters; they would have
become extinct while they were still in process of arising. Although
we see that the Irish stag, with his enormous antlers over ten feet
across from tip to tip, was heavily burdened, we are hardly justified
in concluding that the size and weight of the burden on his head
tended to his destruction from the first—for in that case the species
would never have developed at all—but it may well be that at some
time or other the life-conditions of the species altered in such a
manner that the heavy antlers became fatal to it. In this case the
variation-direction which had gained the mastery in all ids could no
longer be sufficiently held in check by personal selection, because
the variations in the contrary direction would be much too slight to
attain to selective value. Sudden, or at least rapidly occurring
changes in the conditions of life, such as the appearance of a
powerful enemy, exclude all chance of adaptation by the slow
operation of personal selection.
If we look into the matter more carefully, we see that it is not strictly
true to say that germinal selection alone brings about the extinction
of a species by cumulative augmentation of structures which are
already excessive; it is the incapacity of personal selection to keep
pace with the more rapid changes in the conditions of life and to
reduce excessive developments to any considerable extent in a short
time. This would always be possible in a long time, for the
determinants of the excessive organ E can never be equally strong in
all the ids; they always fluctuate about a mean, however high this
mean may be. Here again it must still be possible that reducing-
divisions and amphimixis may lead to the formation of majorities of
ids with weaker E-determinants, and if sufficient time be allowed,
artificial selection could, by consistently selecting the individuals
with, let us say, weaker antlers, give rise to a descending variation-
movement. There are no variation-movements which cannot be
checked; every direction can be reversed, but time and something to
take hold of must be granted. That was wanting in the case of the
giant stag, for it would not have been saved even if its antlers had at
once become a couple of feet shorter, and germinal selection can
hardly make so much difference as that.
Analogous to hereditary deformities, and of special interest in
connexion with the processes within the germ-plasm, are 'sports'
variations of considerable magnitude which suddenly appear without
our being able to see any definite external reason for them. I have
already discussed these in detail in my Germ-plasm, and have shown
how simply these apparently capricious phenomena of heredity can
be understood in principle from the standpoint of the germ-plasm
theory.
The chances of the transmission of the saltatory variation will be
greater or less according to whether the variation of the relevant
determinants involves a bare majority of ids or a large majority, for
the more ids that have varied, the greater is the probability that the
majority will be maintained throughout the course of ensuing
reducing divisions and amphimixis, that is, that the seeds of the
plant will reproduce the variation, and will not revert to the ancestral
form. Although one of the most satisfactory results of the id-theory
lies precisely in the interpretation of these conditions, I do not wish
to enter into the matter here, but will refer to the details in my
Germ-plasm, published in 1894, which I consider valid still. At that
time I had not formulated the idea of germinal selection, but the
explanation of the occurrence of such sport-variations which I gave
was based upon the assumption of nutritive fluctuations in the germ-
plasm, which gave rise to variations in certain determinants. There
was still lacking the recognition that the direction of variation once
taken must be adhered to until resistance was met with, and that
the determinants stand in nutritive correlation with one another, so
that changes in one determinant must re-act upon the neighbouring
ones, as I shall explain more fully afterwards. I also showed from
definite cases that such sports, though they are sudden
—'saltatory'—in their mode of occurrence, are long being prepared
for by intimate processes in the germ-plasm. This 'invisible prelude'
of variation depends on germinal selection. When a wild plant is
sown in garden-ground it does not require to vary at once; several,
even many, generations may succeed each other which show no
sports; suddenly, however, sports appear, at first singly, then,
perhaps, in considerable numbers. It is not, however, by any means
always the case that considerable numbers occur, for some varieties
of our garden flowers have arisen only once, and then have been
propagated by seed; and such saltatory sports in plants which are
raised from seed are usually constant in their seed, and if they are
fertilized with their own pollen they breed true—a proof that the
same variations must have taken place in the relevant determinants
in a large majority of ids.
In animals, it would appear, such saltatory variations occur much
more rarely than in plants; the case examined in detail by Darwin of
the 'black-shouldered peacock' which suddenly appeared in a
poultry-yard is an example of this kind. Much more numerous,
however, are the instances among plants, and especially among
plants which are under cultivation. This indicates that we have here
to do with the effect of external conditions, of nutritive influences
which cause the slow variation of certain determinants, sometimes
abetting and sometimes checking. As soon as a majority of ids
varied in this way comes to lie in a seed, a sport springs up suddenly
and apparently discontinuously—a plant with differently coloured or
shaped petals or leaves, with double flowers, with degenerate
stamens, or with some other distinguishing mark, and these new
characters persist if the variety is propagated without inter-crossing.
But it happens sometimes, though more rarely, that not the whole
plant but individual shoots may exhibit the variation. To this class
belong the 'bud-variations' of our forest trees, the copper-beeches,
copper-oaks, and copper-hazels, the various fasciated varieties of
oak, beech, maple, and birch, and the 'weeping trees'; also the
numerous varieties of potato, plantain, and sugar-cane. It seems
that only a few of these breed true when reproduced from seed, or
in other words, they usually exhibit reversions to the ancestral form:
on the other hand, in the weeping oak for instance, nearly all the
seedlings exhibit the character of the new variety, though 'in varying
degrees.' The records as to the transmissibility of bud-variations
through seed are probably not all to be relied upon, and new
investigations are much to be desired, but the fact that in many
cases they may be propagated not only by means of layers and
cuttings but by seed also, is most important in our present
discussion, for it proves that here too the varied determinants must
be contained in a majority of ids. As it is only a single shoot that
exhibits the saltatory variation, only the germ-plasm which was
contained in the cells of this one shoot can have varied, and it must
have done so in so many ids that the variation prevailed and found
expression. But that, in this case also, the variation does not appear
in all, but only in a small majority of ids, is proved by the frequent
reversion of bud-varieties to the ancestral form. I have already
reported a case of this kind shown to me by Professor Strasburger in
the Botanic Gardens in Bonn, where a hornbeam with deeply
indented 'oak-leaves' had one branch which bore quite normal
hornbeam leaves. In my own garden there is an oak shrub of the
'fern-leaved' variety, whose branches bear some leaves of the
ordinary form; variegated maples with almost white leaves often
exhibit in individual branches a reversion to the fresh green leaves of
the ancestral form. We see from this that what is so energetically
disputed by many must in reality occur—namely, differential or non-
equivalent nuclear division—for otherwise it would be unintelligible
how the ids of the new variety, if they once attain a majority in the
tree, could give place in an individual branch to a majority of the
ancestral ids. Only differential nuclear division, in the manner of a
reducing division, can be the cause of this. Of course this implies
only a dissimilar or differential distribution of the ids between the
two daughter-nuclei, not a splitting up of the individual ids into non-
equivalents.
That in free Nature bud-variations left to themselves can ever
become permanent varieties is probably an unlikely assumption,
because of the inconstancy of their seeds which only breed true in
rare cases; nor is it likely that such variations as the copper-beech,
the weeping ash, and so on could hold their own in the struggle for
existence with the older species; but there is certainly nothing to
prevent our assuming that, in certain circumstances, saltatory
variations, when they have a germinal origin, may become persistent
varieties and may even lead to a splitting of the species. This may
happen, for instance, when the variations remain outside the limits
of good and bad, and thus are neither of advantage to the existence
of the species nor a drawback thereto. In the next chapter we shall
discuss the influence of isolation upon the formation of species, and
it will be seen that in certain conditions even indifferent variations
may be preserved, and that saltatory variations, as for instance in
the evolution of species of land-snails or butterflies, may have
materially contributed to bring this about.
I should like to emphasize still more the part played by saltatory
variations arising from germinal selection in the origin of secondary
sexual characters. As soon as personal selection, whether sexual or
ordinary, prefers as useful in any sense a saltatory variation, it is not
only preserved and becomes a character of a variety, but it may
increase, and we have to ask whether such sudden variations are
frequently of a useful kind, especially when not individual characters
alone, but whole combinations of them are implicated. If we may
judge from the sports of the flowers and the leaves of plants,
transformations useful to the species as a whole rarely occur
suddenly, that is, they occur only in a few out of very numerous
sports; they are much more frequently indifferent, although quite
visible and often conspicuous variations.
For this reason I am disposed to attribute to saltatory variations a
considerable share in the production of distinctive sexual characters.
From saltatory variations in flowers, fruits, and leaves we know that
these may be conspicuous enough even on their first appearance,
and so we are justified in finding in such variations the first
beginnings of many of the decorative distinguishing characters which
occur in the males of so many animals, especially butterflies and
birds. As soon as it is admitted that variations of considerable
amount, which have been slowly prepared in the germ-plasm by
means of germinal selection, can suddenly attain to expression, one
of the objections against sexual selection is disposed of, for
conspicuous variations are necessary for the operation of this kind of
selection, since the changes in question must attract the attention of
the females if they are to be preferred. Without such preference,
even though it be not quite strict and consistent, a long-continued
augmentation of the decorative characters is inconceivable.
But as intra-germinal disturbances of the position of equilibrium in
the determinant system is at the root of the saltatory variations of
our cultivated plants, it must also have played a large share in the
evolution of breeds among our domesticated animals, which is
therefore by no means wholly due to artificial selection operating
upon the variation of individual characters. In all breeds in the
formation of which the production of more than a single definite
character was concerned, as, for instance, in the broad-nosed
breeds of dog—bull-dog and pug-dog—we may refer the peculiar
variation of many parts to disturbances of the equilibrium of the
determinant system, which bring to light, not suddenly as in the
case of saltatory variations, but gradually and increasingly, the
curious complex of characters. Darwin referred such transformations
of the whole animal facies, where a single varying character is
deliberately selected, to correlation, and by this he understood the
mutual influence of the parts of an animal upon one another. Such
correlation certainly exists, as we have already seen in discussing
histonal selection, but here we have rather to do with the correlation
of the parts of the germ-plasm, with the effects of germinal
selection, which, affected by the artificial selection of particular
characters, gradually brings about a more marked disturbance in the
whole determinant system.
In the evolution of our breeds of domesticated animals, germinal
selection in the negative sense must also have played a part—I
mean through the weakening and degeneration of individual
determinants. Only in this way, it seems to me, can we explain the
tameness of our domestic animals, dogs, cats, horses, &c., in which
all the instincts of wildness, fleeing from Man, the inclination to bite,
and to attack, have at least partly disappeared. It is, of course, very
difficult to estimate how much of this is to be ascribed to acquired
habitude during the individual lifetime. The case of the elephant
might be cited in evidence of tameness which arises in the individual
lifetime, for all tame elephants are caught wild, but it seems that
captured young beasts of prey, such as the fox, wolf, and wild cat,
not to speak of lions and tigers, never attain to the degree of
tameness exhibited by many of our domesticated dogs and cats. The
very considerable differences in the degree of tameness of dogs and
cats go to show that the case is one of instincts varying in different
degree.
If this be so, then the instinct of wildness, if I may express myself so
for the sake of brevity, has degenerated in consequence of its
superfluity, and through the process of germinal selection, which
allowed the determinants of the brain-parts concerned to set out on
a path of downward variation upon which they met with no
resistance on the part of personal selection.
Herbert Spencer adduced against my position the case of the
reduction in the size of the jaws in many breeds of dog, especially in
pugs and other lap-dogs, which he regarded as evidence of the
inheritance of acquired characters. But this and analogous cases of
the degeneration of an organ during a long period in which the
animal had been withdrawn from the conditions of natural life is
intelligible enough on the assumption of persistent germinal
selection aided by panmixia. The jaws and teeth in these spoilt pets
no longer require to be maintained at the level of strength and
sharpness essential to their ancestors which depended on these
characters, and so they fell below it, became smaller and weaker,
but could not disappear altogether, for the process of degeneration
was brought, or is being brought, to a standstill by the intervention
of personal selection.
Even the lower jaw in Man is declared by many authors to be
degenerate. Collins found that the lower jaw of the modern
Englishman was one-ninth smaller than that of the ancient Briton,
and one-half smaller than that of the Australians; Flower showed
that we are a microdont race like the Egyptians, while the Chinese,
Indians, Malays, and Negroes are mesodont, and the Andamanese,
Melanese, Australians, and Tasmanians are macrodont. This does not
of itself imply that we exhibit a degeneration of dentition, though
this conclusion is hinted at by other facts, such as the variability of
the wisdom-teeth. It need not surprise us, indeed, that a
retrogressive variation tendency should have started in this case, for,
with higher culture and more refined methods of eating, the claims
which personal selection was obliged to make on the dentition have
been greatly diminished, and germinal selection would thus
intervene.
Every one knows how the quality of human teeth has deteriorated
with culture, and this not in the higher classes only, but even among
the peasantry, as Ammon has observed. The time is past when raw
flesh was a dainty, and when bad teeth meant poor nutrition, if not
actual starvation. Even nowadays famine plays a terrible and
periodically recurrent rôle as an eliminator among some negroid
races.
Many other organs in man have been reduced from their former
pitch of perfection through culture, and some of them are still in
process of dwindling. When I formulated the idea of panmixia and
applied it to explain cases which had previously been referred to the
inheritance of the results of disuse, I regarded the short-sightedness
of civilized Man from this point of view. My opinion aroused lively
opposition at the time, especially on the part of oculists, who very
emphatically referred the phenomenon to the inheritance of acquired
shortsight, and indeed regarded it as a proof of the transmission of
functional modifications.
But, apart from the fact that the assumption of this mode of
inheritance must now be regarded not only as unproved, but as
contradicted by reliable data, panmixia, in conjunction with the
ceaseless fluctuations within the germ-plasm—germinal selection—
affords a better explanation than the other theory was ever in a
position to offer. At that time I pointed out that the survival of the
individual among civilized races had not for a very long time
depended on the perfection of his eyesight, as it does for instance in
the case of a hunting or warlike Indian, or of a beast of prey, or of a
herbivore persecuted by the beast of prey. And this is by no means
due solely to the invention of spectacles, but in a much greater
degree to the fact that every man no longer has to do everything, so
that numerous possibilities of gaining a livelihood remain open to the
less sharp-sighted; that is, the division of labour in human society
has made the survival of the short-sighted quite feasible. As soon as
this division of labour reached such a degree that the founding of a
family offered no greater difficulty to the short-sighted individual
than to one with normal sight, short-sightedness could no longer be
eliminated; and partly because of the mingling with normal sight,
but partly also because of the never-failing minus-fluctuations of the
germ-plasm determinants concerned, a variation in a downward
direction was bound to set in, and will continue until a limit is set to
it by personal selection. Meantime, we are obviously still in the midst
of the process of eye-deterioration; and the resistance to it is
somewhat inhibited in its operation, because although individuals
with extremely bad sight are for the most part hindered from gaining
an independent livelihood and having a family, this is certainly,
thanks to our mistaken humanity, not always the case. There are
even instances of marriage between two blind persons!
As yet, however, the deterioration of eyes has not advanced very far;
not nearly all families are affected by it, and even in Germany, the
land of the 'longest school form' and of the greatest number of
spectacle-wearers, short-sight is still usually acquired by individuals,
although there must frequently be a more or less marked
predisposition to it. It is a common objection to this view that in
England, France, and Italy the percentage of short-sighted
individuals is much lower, and, in point of fact, one sees far fewer
people wearing spectacles in those countries. This, however, does
not prove that a similar deterioration of eyes has not begun there
also, for how could the small inherited beginnings be detected if
they were not accentuated by the spoiling of the eyesight in the
lifetime of the individual by much reading of bad print, and by
writing with bent head, as is still too often the case in many German
schools.
That our interpretation, through panmixia on a basis of germinal
selection, is the correct one, we infer also from the fact that short-
sightedness has been proved to be a frequent character even among
our domesticated animals, such as the dog and the horse. These
animals receive protection and maintenance from Man, and their
survival and reproduction no longer depend on the acuteness of
their sight, and thus the eye has fallen from its original perfection,
just as in Man, although in this case reading and writing play no
part.
A whole series of similar slight deteriorations of individual organs
and systems of organs might be enumerated, all of which have
appeared in consequence of long and intensive culture in Man. All
these must depend upon germinal selection, on a gradually
progressive weakening of the determinant-groups concerned, under
the conditions of panmixia, that is, in the absence of positive
selection.
To these must be added the deterioration of the mammary-glands
and breasts, and the inability to suckle the offspring which results
chiefly from this. Here we have a variational tendency which could
not appear in a people at a lower stage of culture, and it has not
become general in the lower classes of society among ourselves.
The muscular weakness of the higher classes is another case in
point, and all gymnastics and sports will be of no avail as long as a
relative weakness of the muscles is not a hindrance to gaining a
livelihood, and having a family. Even universal conscription will do
nothing to check this falling off of the bodily strength. Certainly
military service strengthens thousands, and hundreds of thousands
of individuals, but it does not prevent the weaklings from
multiplying, and thus reproducing the race-deterioration. But it
would indeed be well if only those who had gone through a term of
military service were allowed to beget children.
It is only among the peasantry, inasmuch as they really work and do
not merely look on as proprietors of the ground, that such a
deterioration of the general muscular strength could not become the
permanent variational tendency of the determinants concerned,
because among genuine peasants bodily strength is a condition of
having and supporting a family—at least on an average.
The diminution in the firmness and thickness of the bones in the
higher classes, and many another mark of civilization, must be
looked at from the point of view of panmixia and germinal selection;
perhaps also the smaller hands and feet which frequently occur
along with a more graceful general build in the higher ranks of
European peoples. It would certainly not be surprising if in families
which usually intermarry, and which in no way depend for their
material subsistence on the possession of large and powerful hands
and feet or bones generally, a downward variation of the relevant
germ-determinants should have developed, but this could never
overstep a certain limit, because it would then be prejudicial even in
civilized life. That we must be very careful not to regard large hands
and feet as the direct result of hard physical toil was brought home
to me by an observation of Strasburger's. He was particularly struck
by the fact that the peasants of the high Tatra (Carpathians) were
distinguished by the smallness of their hands and feet.
But while civilization has excited numerous downward variations in
the germ, it has, on the other hand, been the cause of numerous
hereditary improvements—variations in an upward direction. This
opens up new ground, for hitherto we have been confronted with
the alternative of either accepting the inheritance of acquired
characters, and on this basis referring the talents and mental
endowments of civilized Man to exercise continued throughout many
generations, or of admitting an increase of mental powers only in as
far as they possess 'selection value,' that is, as they may be decisive
in the struggle for existence. To these mental qualities belong
cleverness and ingenuity in all directions, courage, endurance, power
of combination, inventive power, with its roots in imagination and
fertility of ideas, as well as desire for achievement, and industry.
Throughout the long history of human civilization these mental
qualities must have increased through the struggle for existence, but
how have the specific talents such as those exhibited in music,
painting, and mathematics come into existence? And how have the
moral virtues of civilized Man been evolved, and particularly
unselfishness? For it can hardly be maintained of any of these
endowments that they possess selection-value for the individual.
It is not my intention to discuss these questions in detail; they are
too many-sided and of too much importance to be treated of merely
in passing; moreover, I gave expression years ago to my views on
this subject by dealing with one example—the musical sense in Man.
I do not believe that the musical sense had its beginnings in Man, or
that it has materially increased since the days of primitive Man, but
in conjunction with the higher psychical life of civilized peoples its
expressions and applications have risen to a higher level. It is, so to
speak, an instrument which has been transmitted to us from our
animal ancestors, and on which we have learnt to play better the
more our mind has developed; it is an unintended 'accessory effect'
of the extremely fine and highly developed organs of hearing with
their nerve-centres which our animal ancestors acquired in the
struggle for existence, and which played a much more important rôle
in the preservation of life in their case than it does in ours. The
musical sense may be compared to the hand, which was developed
even among the apes, but which civilized Man in modern times no
longer uses merely to perform its original function, grasping, but
also for many other purposes, such as writing and playing the piano.
And just as the hand did not originate through the necessities of the
piano, neither did the extremely delicate sense of hearing of the
higher animals develop for the sake of music, but rather that they
might recognize their enemies, friends, and prey, in darkness and
mist, in the forest, on the heath, and at great distances.
The case is probably the same with the rest of the special psychical
endowments or talents. I do not of course maintain that they, like
the musical sense, did not at some time play a rôle in the struggle
for existence and survival, and therefore could not increase, but the
increase was certainly not continuous, but much interrupted, so that
it would extend only to small groups of descendants, and therefore
could only contribute very slowly to the elevation of the psychic
capacities of a whole people. But in certain individuals and families
such augmentations would certainly take place through germinal
selection, and it seems to me probable that these would never be
wholly lost again, even if they appeared to be so, but would be
handed on, in id-minorities, through the chain of generations, and
would slightly raise the average of the talent in question, and might
even, under favourable circumstances, combine in the development
of a genius. We know how strongly hereditary such specific talents
are; let us suppose that the determinants of, say, the musical sense
have, by the intra-germinal chances of nutrition, been started on a
path of ascending variation; they will continue in this path until a
halt is called from some quarter or other. This can only happen if, in
the reducing division, or in amphimixis, the highly developed musical
determinants are wholly or partly eliminated, or are reduced to a
minority. As long as this does not happen the ascending variation will
go on, and then we may have the birth of a Mozart or of a
Beethoven. Personal selection will not interfere either in a positive or
a negative sense, since high development of the musical sense has
no effect either in advancing or retarding the struggle for existence;
the increase will therefore go on until the large majority of highly
developed musical determinants, which we must assume in the case
of a musical genius, is reduced, or even transformed into a minority,
through unfavourable reducing divisions of the germ-cells, and by
association with the germ-cells of less musical mates.
The fact that highly developed specific talents have never been
known to be inherited through more than seven generations is quite
in keeping with this view. But even this persistence has been
observed only in the case of musical talent, and the long
continuance of the inherited talent may well be due, as Francis
Galton suggests in his famous statistical investigations into the
phenomena of inheritance, to the fact that musical men do not
readily choose wives who are absolutely lacking in this talent. It
would be easy to rear an exceedingly highly gifted musical group of
families within the German nation, if we could secure that only the
highly-gifted musically should unite in marriage—that is, if personal
selection could play its part. In another more general domain of
mental endowment a case of this kind has been recorded, for Galton
tells us of three highly gifted English families which intermarried for
ten generations, and in that time scarcely produced a descendant
who did not deserve to be called a distinguished man in some
direction or other.
Of course, such continued persistence, through a long series of
generations, of a high general mental level is more possible than the
transmission and increase of a specific talent, for in the former case
it is a question of a mixture of different high mental endowments, of
which not all need be developed in every individual, and yet the
individual need not fall to mediocrity if he possesses a combination
of other qualities. But in musical talent, on the other hand, the
falling from the height once attained takes place as soon as this one
character is no longer represented in a sufficiently strong majority of
determinants. Of course it would be a mistake to believe that the
talent of a Sebastian Bach or a Beethoven depended solely on the
highly developed musical sense; in them, as in all great artists, many
highly developed mental qualities must have combined with the
musical sense; a simpleton could never have written the Mass in B
minor or the Passion of St. Matthew even if he had possessed the
musical genius of Sebastian Bach. In this fact lies a further reason
why genius is seldom found at the same pitch in two successive
generations; the combination of mental characters always varies
from father to son, and slight displacements may give rise to very
great differences in relation to the manifestations of the specific
talent. Under certain circumstances, the weak development of a
single trait of character, as, for instance, power of action, or the
excessive development of another, such as indecision or
desultoriness, may so nullify the existing favourable combinations of
mental characters, such as, let us say, musical sense, inventive
talent, depth of feeling, &c., that they bear no fruit worth
mentioning. And since as we have already seen, the different mental
qualities of the parents are to a certain extent separately
transmitted, that is, since they may appear in the children in the
most diverse combinations, we should rather be surprised that
pronounced talent in a specific direction can persist in a family for
two and a half centuries than that it should do so very rarely. For
reducing division is always combining the existing mental qualities
anew, and amphimixis is adding fresh ones to them.
Thus germinal selection, that is, the free, spontaneous, but definitely
directed variation of individual groups of determinants, is at the root
of those striking individual peculiarities which we call specific talents;
but it can attain to the highest level only rarely and in isolated cases,
because these talents are not favoured by personal selection, and
therefore the excessively highly developed determinants upon which
they depend may be dispersed in the course of generations; they
may sink to smaller majorities, or even to minorities, in which case
they will no longer manifest themselves in visible mental qualities.
We deduced the process of germinal selection on the basis of the
assumption that the nutrition of all the parts and particles of the
body, therefore also of the determinants and biophors of the germ-
plasm, is subject to fluctuations. We regarded the resulting
variations of these last and smallest units of the germ-plasm as the
ultimate source of all hereditary variation, and therefore the basis of
all the transformations which the organic world has undergone in the
course of ages and is undergoing still.
We have still to inquire whether we can give any more precise
account of the nature of these units of the germ-plasm. If I mistake
not, we may say at least so much, that all variations are, in ultimate
instance, quantitative, and that they depend on the increase or
decrease of the vital particles, or their constituents, the molecules.
For this reason I have hitherto always spoken of only two directions
of variation—a plus or a minus direction from the average. What
appears to us a qualitative variation is, in reality, nothing more than
a greater or a less, a different mingling of the constituents which
make up a higher unit, an unequal increase or decrease of these
constituents, the lower units. We speak of the simple growth of a
cell when its mass increases without any alteration in its
composition, that is, when the proportion of the component parts
and chemical combinations remains unchanged; but the cell changes
its constitution when this proportion is disturbed, when, for instance,
the red pigment-granules which were formerly present but scarcely
visible increase so that the cell looks red. If there had previously
been no red granules present, they might have arisen through the
breaking up of certain other particles—of protoplasm, for instance, in
the course of metabolism, so that, among other substances, red
granules of uric acid or some other red stuff were produced. In this
case also the qualitative change would depend on an increase or
decrease of certain simpler molecules and atoms constituting the
protoplasm-molecule. Thus, in ultimate instance, all variations
depend upon quantitative changes of the constituents of which the
varying part is composed.
It might be objected to this argument that chemistry has made us
acquainted with isomeric combinations whose qualitative differences
do not depend upon a different number of the molecules composing
them, but upon their different arrangement; it might be supposed
that something similar would occur also in morphological relations.
And, in point of fact, this seems to be the case. We may, for
instance, imagine one hundred hairs as being at one time equally
distributed on the back of a beetle, and at another standing close
together and forming a kind of brush, but although this brush would
be a new character of the beetle, yet its development would depend
upon quantitative differences, namely, on the fact that the same
skin-area, which in the first case bore perhaps only one hair, had in
the second case a hundred. The quantity of hair cells has notably
increased upon this small area. In the same way the characteristic
striping of the zebra depends not on a qualitative change in the skin
as a whole, but upon an increased deposit of black pigment in
particular cells of the skin, therefore on a quantitative change. In
relation to the whole animal it is a qualitative variation, as
contrasted, for instance, with the horse, but in respect of the
constituent parts which give rise to the qualitative variation it is
purely quantitative. The character of the whole edifice is changed
when the proportion of the stones of which it consists are altered.
Thus the determinants of the germ may not only become larger or
smaller as a whole, but some kinds of the biophors of which they are
made up may increase more than others, under definite altered
conditions, and in that case the determinants themselves will vary
qualitatively, so that, from the changing numerical proportions of the
different kinds of biophors, a variation of the characters of the
determinants can arise, and consequently also qualitative variations
of the organs controlled by the determinants—the determinates. But,
since nothing living can be thought of as invariable, the biophors
themselves may, on account of nutritive fluctuations, grow
unequally, and thereby vary in their qualities. To follow this out in
greater detail and attempt to guess at the play of forces within the
minutest life-complexes would at present only be giving the rein to
imagination, but in principle no objection can be made to the
assumption that every element of life down to the very lowest and
smallest can, by reason of inequalities in its nutrition, be not only
started on an ascending or descending movement of uniform
growth, but can also be caused to vary qualitatively, that is, in its
characters, because its component parts change their proportions.
Of course we know nothing definite or precise with regard to the
units of the germ-plasm, and we cannot tell what is necessary in
order that a determinant shall determine a part of the developing
body in this way or in that; thus we have no definite idea of the
relations subsisting between the variations of the determinants and
those of their determinates, but we know at least so much, that
hereditary variation of a part is only possible when a corresponding
particle in the germ-plasm varies; and we may at least assume that
these correspond to each other so far, that a greater development of
the one implies a greater development of the other, and that a
reversal of these relations is impossible. If the determinant X
disappears from the germ-plasm the determinate X´ disappears
from the soma. It is therefore justifiable to infer from the degree of
development of an organ the strength of its determinant, and to
assume that plus- and minus-variations in both are correspondingly
large.
But in addition to the fluctuations in the equilibrium of the germ-
plasm which lie at the root of all hereditary variation, we have to
take into account something which we have already touched upon
briefly—the correlation of the determinants, the influencing of one
determinant by those round about it. I have spoken for the sake of
brevity of 'the determinant' of a part, although all the large and
more important parts must certainly be thought of as represented by
several or many, if not, indeed, by whole groups of determinants.
Although it is quite out of our power to follow the complex processes
of the mutual influences of the determinants upon each other, we
can say this at least, that these influences must exist, and we have
here a faint indication of what must occur in the case of
spontaneous variations within the germ-plasm. We must, in the first
place, think of the individual determinants as arranged in groups, so
that, for instance, the determinants of the right and left half of the
body lie together, and therefore are frequently affected together by
influences which cause variation, so that both vary in the same
direction at the same time. In point of fact, analogous deformities,
such as polydactylism of both right and left hands, and even of
hands and feet at once, do actually occur. That the right and left
hands, the fore- and hind-limbs, are represented in the germ by
particular determinants, may be inferred from their frequently
different phyletic evolution into different forms of hand and foot, e.g.
into flipper and rudimentary hind-leg in the whale, as well as from
the cases of particulate inheritance, which are rare, but which
undoubtedly do occur, such as when, in Man, there is a maternal
blue eye on one side of the head and a paternal brown eye on the
other. But almost more striking than the differences between these
homologous or homotypic parts are their points of resemblance, and
these may probably be in part referred to their disposition side by
side and common history in the germ-substance, although a far
larger proportion of them are probably due to their adaptation to
similar functions, and are therefore to be regarded as a phenomenon
of convergence within the same organism.
We have already seen that the first increase in the growth of one
determinant means a withdrawal of nourishment, however slight,
from its neighbours; this can, of course, be equalized again if the
claims on the common nutritive stream from another quarter are at
the same time diminished; but it is possible that the claims from
another quarter may also be increased, and the withdrawal will then
be more marked, and the determinants being thus injured from two
directions at once will sink downwards with greater rapidity. But it is
also conceivable that the majority of determinants of a part may
vary upwards, and, by their combined increased power of
assimilation, direct towards themselves such a greatly increased
stream of nourishment that the whole organ—for instance, a
particular feather in a bird—varies in an upward direction, and
becomes larger and larger, as we see in the case of many decorative
feathers; or that certain determinants vary only as far as some of
their biophors are concerned, and similarly for their determinates, as
when a group of scales on a butterfly's wing that had previously
been black turn out a brilliant blue. It can probably also happen that
such variations within the determinants are transmitted to
neighbouring determinants because the nutritive conditions which
caused the first to vary have extended to those about them. The
increase of brightly coloured spots in birds and butterflies gives us
ground for concluding that there are processes of this kind within
the germ-plasm.
I will refrain from following this idea into greater detail, and
translating the observable relations and variations of the fully-formed
parts of the body into the language of the germ-plasm; but so much
may be taken as certain, that multitudinous inter-relations and
influences exist between the elements of the germ-plasm, and that
one variation brings another in its train, so that—usually at a very
slow rate, that is, in the course of generations and of species-
forming, definite variations occur from purely intra-germinal reasons
—variations which as far as they remain outside the limits of good or
bad may of themselves change the character of a species, but which
when they are seized upon by personal selection may, by sifting and
combination of the ids, be led on to still higher development.
If we consider further that the variation of a part must depend not
only on the quality of the external stimulus but also upon the
constitution, the reacting power of the part, we shall understand
that similar nutritive variations may cause two different determinants
to vary in different ways, and when we reflect that every nutritive
change must extend from the point from which it started with
diminishing strength in a particular direction, we have a further
factor in the variation of determinants and one which influences
even similar determinants differently.
Finally, if we remember that determinants of different constitution
will also extract different ingredients from the nutritive stream and
thus set up in it different kinds of chemical change, thus causing an
altered supply of nutritive substances to flow to the neighbour
determinants, we get some insight into a very complex and delicate
but perfectly definite set of processes, into a mechanism which we
can certainly only guess at, but whose results lie plainly before us in
the spontaneous variations of the organism. We understand in
principle the possibility of saltatory variation, as a more or less
widespread, more or less marked disturbance of the species-type in
this or that group of characters, and we may acknowledge that
those 'kaleidoscopic variations' which Eimer supposed to be the sole
basis of the transformation of species, and which have been brought
to the foreground again quite recently by De Vries[20], are probably
factors in transmutation operative within a limited sphere.
[20] See end of chap. xxxiii.

But we must think of all these struggles and mutual influencings as


taking place on the smallest possible scale, so that it is only by long
summation that they can produce any visible effect, and we must
never forget the essential significance of the plurality of ids, for
these 'spontaneous' variations may take place in a different and
quite independent manner in each individual id. If this were not so
no intervention of personal selection would be possible, natural
selection would not exist, and the adaptation of the organism from
the single cell up to the whole would remain wholly unexplained.
The whole crop of spontaneous germ-variations, whenever it ceases
to be 'indifferent,' and becomes either 'good' or 'bad,' comes under
the shears of personal selection and under its almost sovereign
sway.
On the other hand, the sudden first appearance of a saltatory
variation takes place quite independently of personal selection,
depending on similar variations in a number of ids, which remain
latent until they have by the process of reducing division which
precedes amphimixis, chanced to attain a majority. In sudden bud-
variations we may perhaps suppose that reducing division occurring
in some still unverified abnormal manner is the reason why the
germinal variation suddenly makes itself visible—a supposition
previously suggested as the explanation of the reversion of these
sports.
The rarity of bud-variation is thus explained, while the greater
frequency of saltatory variations in plants propagated by seed may
be accounted for by the regular occurrence of reducing division in
sexual reproduction. But that the same or similar variations may
occur in several, it may be in many, ids at the same time must
depend upon similar general influences which affect the plant as a
whole, as happens through cultivation, manuring, and so on. I shall
return to this when discussing the influence of the environment.
In some quarters this whole conception of germinal selection has
been characterized as the merest figment of imagination,
condemned on this ground alone, that it is based on the differences
in nutrition between such extremely minute quantities of substance
as the chromosomes of nuclear substance within the germ-cell. The
quantity of substance is certainly minute, but it needs nutriment
none the less, and can we believe that the stream of nourishment
for all the invisibly minute vital elements is exactly alike? It may be
admitted that the nourishment outside the ids is usually abundant,
although undoubtedly fluctuations occur in it also, but it certainly
does not follow from this that every vital unit within the id is
similarly disposed in relation to the nutritive supply, or has food in
equal quantities at its command, or even that each has as much as it
can ever need. To make an assertion like this seems to me much the
same as if an inhabitant of the moon, looking at this earth through
an excellent telescope and clearly descrying the city of Berlin with its
thronging crowds and its railways bringing in the necessaries of life
from every side, should conclude from this abundant provision that
the greatest superfluity prevailed within the town, and that every
one of its inhabitants had as much to live upon as he could possibly
require.
We certainly ought not to conclude from the fact that we cannot see
into the structure and requirements and methods of nutrition of a
very minute mass of substance that its nutrition cannot be unequal,
and that it cannot, by its inequalities, give rise to very material
differences, especially when we are dealing with a substance to
which we must attribute an extraordinarily complex organization
built up of enormous numbers of extremely minute particles. That
this complexity is undeniable is now admitted by many who formerly
thought it possible to believe in the simple structure of the germ-
substance. How complex not only the germ-substance but every cell
of a higher organism is in its structure, and how far below the limits
of visibility its differentiations and arrangements reach, is pressed
upon our attention by the most recent histological researches, such
as those we owe to Heidenhain, Boveri, and many others. The whole
scientific world was amazed when it came to know the mysterious
nuclear spindle in the seventies, and since then this has been quite
thrown into the shade by the discovery of the centrosphere, the
centrosome, and more recently even the centriole, and now we
believe that these marvellous centres of force may, or must, possess
their own dividing apparatus! In the face of discoveries like these no
one is likely to be able to persist in recognizing as existing only what
is disclosed or even hinted at by the most powerful lenses; no one
can any longer doubt that far below the limit of visibility organization
is still at the basis of life, and that it is dominated by orderly forces.
To me, at least, it seems more cogent to argue from the phenomena
of heredity and variation to an enormous mass of minute vital units
crowded together in the narrow space of the id, than to argue from
the calculated size of atoms and molecules to the number which we
are justified in assuming to be present in an id. In my book on the
germ-plasm I made a calculation of this kind, and I arrived at figures
which seemed rather too small for the requirements of the germ-
plasm theory. This has been regarded as a proof that I disregard the
facts for the sake of my theory, but it should rather be asked
whether the size of the atoms and molecules is a fact, and not rather
the very questionable result of an uncertain method of calculation.
Undoubtedly modern chemistry has established the relative weight-
proportions of the atoms and molecules with admirable precision,
but it can make only very uncertain statements in regard to the
absolute size of the ultimate particles. It is therefore admissible to
assume that these have a still greater degree of minuteness when
the facts in another domain of science require this.
We must assume determinants, and consequently the germ-plasm
must have room for these; the variations of species can only be
explained through variations of the germ-plasm, for these alone give
rise to hereditary variation. It is upon this foundation that my
germinal selection is built up; whether I have in the main reached
the truth the future will show: but that I have not exhausted this
new domain, but only opened it up, I am very well aware.
LECTURE XXVII
THE BIOGENETIC LAW
Fritz Müller's ideas—Development of the Crustaceans—Of the
Daphnidæ—Of Sacculina—Of parasitic Copepods—Larvæ of the higher
Crustaceans—Change of phyletic stages in Ontogeny—Haeckel's
Fundamental Biogenetic Law—Palingenesis and Cœnogenesis—
Variation of phyletic forms by interpolation in a lengthened Ontogeny—
Justification of deductions from Ontogeny to Phylogeny—
Würtemberger's series of Ammonites—Phylogeny of the markings in
the caterpillars of the Sphingidæ—Condensation of Phylogeny in
Ontogeny—Example from the Crustaceans—Disappearance of useless
parts—The variation of homologous parts, according to Emery—Germ-
plasmic correlations—Harmony with the theory of determinants—
Multiplication of the determinants in the course of the phylogeny.

What I propose to discuss in this lecture should have been


considered at an earlier stage, if we had pledged ourselves to
adhere strictly to the historical sequence of scientific discovery, for
the phenomena which we are about to deal with attained recognition
shortly after the revival of the evolution idea, and indeed they
formed the first important discovery which was made on the basis of
the Darwinian Doctrine of Descent. I have introduced them at this
stage because they have to do with phenomena of inheritance and
modifications of these, the understanding of which—in as far as we
can as yet speak of understanding at all—is only possible on the
basis of a theory of inheritance. Therefore, in order to examine
these phenomena and their causes, it was necessary first to submit
a theory of heredity, as I have done in the germ-plasm theory. We
have to treat of the connexion between the development of many-
celled individuals and the evolution of the species, between germinal
history and racial history, or, as we say with Haeckel, between
ontogeny and phylogeny.
Long before Darwin's day individual naturalists had observed that
certain stages in the development of the higher vertebrates, such as
birds and mammals, showed a likeness to fishes, and they had
spoken of a fish-like stage of the bird-embryo. The 'Natural
Philosophers' of the beginning of the nineteenth century, Oken,
Treviranus, Meckel, and others, had, on the basis of the
transmutation theory of the time, gone much further, and had
professed to recognize in the embryonic history of Man, for example,
a repetition of the different animal stages, from polyp and worm up
to insect and mollusc. But von Baer afterwards showed that such
resemblances are never between different types, but only between
representatives of the same general type, e.g. that of Vertebrata;
and Johannes Müller maintained, from the standpoint of the old
Creation theory, that an 'expression of the most general and simple
plan of the Vertebrates' recurred in the development of higher
Vertebrates, giving as an instance that, at a certain stage of
embryogenesis, even in Man, gill-arches were laid down and were
subsequently absorbed. But why this 'plan' should have been carried
out where it was afterwards to be departed from remained quite
unintelligible.
An answer to this question only became possible with the revival of
the Theory of Descent, and the first to throw light in this direction
was Fritz Müller, who, in his work Für Darwin, published in 1864,
interpreted the developmental history of the individual, 'the
ontogeny,' as a shortened and simplified repetition, a recapitulation,
so to speak, of the racial history of the species, the 'phylogeny.' But
at the same time he recognized quite clearly—what indeed was plain
to all eyes—that the 'racial history' cannot be simply read out of the
'germinal history,' but that the phylogeny is often 'blurred,' on the
one hand by the fusing and shortening of its stages, since
development is always 'striking out' a more direct course from the
egg to the perfect animal, while, on the other hand, it is frequently
'falsified' by the struggle for existence which the free-living larvæ
have to maintain.

You might also like