Sunteți pe pagina 1din 41

EINSTEINS

MASSENERGY
EQUATION,
VOLUME I
EINSTEINS
MASSENERGY
EQUATION,
VOLUME I
EARLY HISTORY AND
PHILOSOPHICAL
FOUNDATIONS

FRANCISCO FERNFLORES
Einsteins MassEnergy Equation, Volume I: Early History and
Philosophical Foundations
Copyright Momentum Press , LLC, 2018.

All rights reserved. No part of this publication may be reproduced, stored


in a retrieval system, or transmitted in any form or by any means
electronic, mechanical, photocopy, recording, or any otherexcept for
brief quotations, not to exceed 250 words, without the prior permission of
the publisher.

First published in 2018 by


Momentum Press , LLC
222 East 46th Street, New York, NY 10017
www.momentumpress.net

ISBN-13: 978-1-60650-857-2 (paperback)


ISBN-13: 978-1-60650-858-9 (e-book)

Momentum Press Energy Physics and Engineering Collection

DOI: 10.5643/9781606508589

Cover and interior design by S4Carlisle Publishing Service Private Ltd.


Chennai, India

First edition: 2018

10 9 8 7 6 5 4 3 2 1

Printed in the United States of America


ABSTRACT

In this volume, we examine the history and philosophical significance of


several demonstrations Einstein published for his massenergy relation,
which is often expressed by the iconic equation E = mc2 . Our goal is
to illustrate how these demonstrations display a clear shift away from
a reliance on electromagnetic phenomena culminating in Einsteins 1934
purely dynamic demonstration. Philosophically, this trend signals the
importance of recognizing special relativity as what Einstein called a prin-
ciple theory. Volume two of this work examines the role that Einsteins
massenergy relation played in the development of quantum mechanics
and general relativity. We also discuss the first empirical confirmation of
E = mc2 and some contemporary debates concerning the philosophical
interpretation of this important result.

Keywords: Einstein, equivalence of mass and energy, history of special rela-


tivity, mass-energy equivalence, mass-energy, philosophical foundations of
special relativity, special relativity.
CONTENTS

PREFACE ix
ACKNOWLEDGMENTS xi
1. Introduction and Conceptual Context 1
1.1 Einsteins 1905 Discovery of the MassEnergy Relation 1
1.2 Mass: From Newton to Einstein 3
1.2.1 Newton and Mass 4
1.2.2 Euler and Mass 7
1.2.3 Maxwell and Mass 13
1.2.4 Mach and Mass 15
1.3 Energy and Its Conservation in the 19th Century 18
1.4 Philosophical Reflections 23
2. Early Days of the MassEnergy Relation 25
2.1 Historical Review of Special Relativity 26
2.2 The First Demonstration 31
2.3 A Second Demonstration 42
2.4 A Third Demonstration 46
2.5 Philosophical Reflections on a Third Demonstration 51
2.6 Early Demonstrations, Philosophically Speaking 53
2.7 Composite Systems and the Inertia of Energy 57
2.8 Composite Systems, Philosophically Speaking 59
2.9 Einsteins Principle versus Constructive Theories 61
3. Einstein and Dynamic Demonstrations of E = mc2 65
3.1 Historical Review of Minkowski Spacetime 65
3.2 Einsteins 1935 Dynamic Demonstration 69
3.3 Einsteins Dynamic Demonstration, Philosophically
Speaking 79
3.4 Lewis and Tolmans 1909 Dynamic Demonstration 80
3.5 Feigenbaum and Mermins Dynamic Demonstration 91
3.6 Feigenbaum and Mermin, Philosophically Speaking 95
BIBLIOGRAPHY 99
INDEX 103
PREFACE

A historical and philosophical analysis of the development and interpreta-


tion of Einsteins famous equation E = mc2 has the potential to improve
our understanding of the subtleties associated with this central result in con-
temporary physics. The purpose of this book is to guide the reader through
a historical tour of efforts by physicists and philosophers to understand both
the epistemological foundations of Einsteins massenergy relation and the
theoretical relationship between this result and the foundations of quantum
mechanics and general relativity.
Einsteins own attempts to grasp the logical relationship between his
massenergy relation and the foundational principles of special relativity
plays a central role in our study. Specifically, in Volume I, we examine
in some detail several of Einsteins attempts to demonstrate, using the
analysis of certain physical configurations that can be understood as thought-
experiments, his massenergy relation. Philosophically, we emphasize a
trajectory that took Einstein from his early demonstrations beginning in
1905, which used electromagnetic phenomena and hence Maxwells the-
ory, to his entirely dynamic demonstration of the massenergy relation
published in 1935. This trajectory, we believe, displays how Einsteins
philosophical view that special relativity was a principle, and not a con-
structive, theory directly affected his scientific practice. We have selected
the demonstrations that most clearly highlight this aspect of Einsteins work.

***

Writing about historical sources in physics forces unique choices about


notation. Our guiding principle when reporting from a historical source has
been to use the notation used by the original author except when changing the
notation leads to a more consistent and simpler presentation. For example,
we have uniformly used c for the speed of light in quotations from Einsteins
papers even though in the first couple of years after his publication of special
x PREFACE

relativity he often used V . On the other hand, in our discussion of de Broglies


phase wave hypothesis in Volume 2, we have retained all of de Broglies
original notation as trying to make it match the notation we use elsewhere
in the text would seem to reduce the clarity of our exposition.
Regarding Einsteins result, part of the controversy among physicists
and philosophers, which we discuss in Volume 2, has been whether we should
say that mass can be converted into energy (or vice versa) and whether
we should say that mass and energy are equivalent. So as not to prejudice
our reader, we have avoided the phrase massenergy equivalence, though
we believe no harm is done by using it so long as we all understand it as
Einstein seemed to do in 1907, i.e., as stating the direct proportionality of the
values of two physical quantities. Consequently, we have throughout used
the phrase massenergy relation as a generic label for the several ways in
which Einsteins result can be expressed correctly both mathematically and
in the language of physics.
Similarly, there has been some controversy among physicists about how
best to express Einsteins result in symbolic notation. Without wading into
the details of this debate, let us simply note that there are two candidates for
how to denote the energy in the massenergy equation: E and Eo . Similarly,
there are two candidates for how to denote the mass: m and mo . There are,
then, four possible ways to write down the massenergy equation.
Our selection of symbols for mass and energy is partially influenced by
the ongoing, though nowadays waning, discussion about the practical value
of the terms rest-energy or rest-mass. We have simply chosen to use m
and E and the corresponding terms mass and energy to refer to the mass
and energy of an object as reckoned in its rest-frame. Our sense is that this
use is becoming sedimented in contemporary physics textbooks that, more
often than not, eschew the terms relativistic mass (there is only one mass,
what used to be called rest-mass) and relativistic energy, which is now
frequently called total energy.
ACKNOWLEDGMENTS

I wish to thank Tal Scriven (Chair, Philosophy Department) and Douglas


Epperson (Dean, College of Liberal Arts) at California Polytechnic State
University, San Luis Obispo, CA, for supporting my research. Dean Epper-
son was ultimately responsible for awarding me, on behalf of the College
of Liberal Arts, a research stipend during the summer of 2016 that was
invaluable. Two of my students made valuable contributions to this work.
Sam Castenholz carefully read and copy-edited about half of the manuscript;
Garrett Goff not only produced the diagrams but verified their accuracy and
appropriateness. Their work was funded by the Philosophy Department at
Cal Poly. I also wish to thank my colleague Karl Saunders (Physics Depart-
ment, California Polytechnic State University, San Luis Obispo, CA) for
insightful conversations about composite systems and levels of granularity
the impact of which is in between the lines in a variety of places and is
screaming for more attention. My correspondence with Roberto Torretti on
Einsteins massenergy relation several years ago was also, in retrospect,
extremely important to my developing understanding of material for this
book. Although I have never met Prof. Torretti, his work has inspired and
influenced me since I first encountered it as a graduate student. Finally, there
is no philosopher who has inspired me more than my spouse Rachel Fern-
flores, who, along with my daughters Olivia and Phaedra, provided me with
unending love and support during this project.
CHAPTER 1

INTRODUCTION AND
CONCEPTUAL CONTEXT

1.1 EINSTEINS 1905 DISCOVERY OF THE


MASSENERGY RELATION

Einstein first expressed the relativistic relationship between mass and energy,
which we now routinely denote E = mc2 , in a 1905 article titled Does the
Inertia of a Body Depend upon Its Energy-Content? The implicit answer to
the question in the title already says a lot. Einstein is claiming that the extent
to which an object resists a change in its state of motion, be it by an impulsive
force or a continuous force, depends on the energy-content of the body.
The energy-content of a body is today sometimes called its internal energy.
Given the general trend among physicists in the early 20th century to regard
energy as a property of physical systems and not as a material entity, it would
no doubt have seemed quite surprising in 1905 that, e.g., upon losing thermal
energy a body would offer less resistance to changes in its state of motion.
The first conclusion Einstein states in his 1905 article is this: If a body
releases the energy L in the form of radiation, its mass decreases by L/V 2
[1, p. 174], where Einstein is using V to designate the speed of light in
vacuo, for which we will hereafter consistently use the more familiar c.
Given the physical configuration Einstein uses to arrive at this result, the
term radiation in this conclusion refers to the electromagnetic radiation.
However, Einstein quickly points out that the form of energy the body gives
off seems to be irrelevant to the result, assuming, as he seems to do, that any
one kind of energy can be transformed into another. Thus, he says, we are
led to the more general conclusion that

the mass of a body is a measure of its energy content; if the energy changes
by L, the mass changes in the same sense by L/9 1020 , if the energy is
measured in ergs and the mass in grams [1, p. 174].
2 EINSTEINS MASSENERGY EQUATION, VOLUME I

It has become commonplace to note that in this first demonstration of the


relativistic relationship between inertial mass and energy, Einstein nowhere
writes down any version of his iconic equation E = mc2 .
For Einstein, a good candidate for testing his second, more general con-
clusion empirically would be to measure the changing mass in bodies whose
energy content is variable to a high degree (e.g., salts of radium) [1, p. 174].
The more energy is released by a sample of radium, the more its mass should
diminish. Given that the change in mass is equal to the change in energy
divided by the speed of light squared, the change in mass would be very
small. Still, Einsteins claim is that a sample of radium should resist changes
to its state of motion less after it has given off energy in the form of radiation.
Finally, Einstein states a closely related third conclusion: If the theory
agrees with the facts, then radiation transmits inertia between the emitting
and absorbing bodies [1, p. 174]. The change in mass that correlates with a
change in the internal energy of a body occurs, as Einstein says, in the same
sense as the change in energy. If a body absorbs energy, e.g., in the form of
radiation, its inertial mass will increase; if it radiates energy, its inertial mass
will decrease. So, if we consider a pair of bodies such that the first emits
energy and the second absorbs it, some of the inertia from the first body will
be transmitted to the second body.
Einsteins third conclusion must have seemed rather surprising to New-
tonian eyes. For example, suppose a sample of radium is situated inside a
box of finite mass with perfectly absorbing walls. If in a given period of time
the radium sample emits energy E, the mass of the radium sample decreases
by an amount E/c2 . However, since the walls of the box are absorbing an
amount of energy E, the mass of the box increases by an amount E/c2 . Con-
sequently, if after this given period of time we magically remove the radium
sample from the box, we shall find that the radium sample offers less resis-
tance to changes in its state of motion, while the box offers more resistance to
changes in its state of motion. The surprising conclusion, from a Newtonian
perspective, is that inertia has been transmitted from the emitting to the
absorbing body.
Attending carefully to the evolving and increasingly rigorous meaning
of assertions made by physical theory will be one of our primary tasks in
this study. For example, Einstein nowhere states that matter is converted
into energy. Similarly, Einstein remains entirely silent about whether the
entire inertial mass of an object could be radiated away. Details of this
sort will be important to us as we try to bridge the sometimes yawning gap
between popular presentations of Einsteins famous equation and its rigorous
application by physicists and engineers.
INTRODUCTION AND CONCEPTUAL CONTEXT 3

Notice also that Einsteins 1905 result about the relationship between
mass and energy is conceptually distinct from considerations in special rel-
ativity about how measurements of inertial mass are related for a pair of
observers who are in a state of relative inertial motion. The issue here is
not about what has sometimes been called relativistic mass. Also, because
some contemporary treatments can leave the reader with the impression that
the relationship between energy and mass in special relativity issues directly
from the definition of the four-momentum, it is also important to note that
Einsteins result, as we shall see in later chapters, depends on accepting
at least the principle of relativity, the light principle (which roughly states
that light travels at the same velocity regardless of the state of motion of
the observer or emitter), and the principle of conservation of energy. It is
from these physical hypotheses, and not merely from definitions of physi-
cal quantities, that Einstein derives an empirically testable prediction. This
prediction, when verified, shows that one property of physical objects that
Newton believed to be unchangeable except by physically separating a part
of the object, viz., its inertial mass, can vary depending on the energy-
content of that body. For example, having radiated some of its thermal
energy, a rock that has cooled resists changes to its state of motion a tiny bit
less. Quantitatively, of course, the change to the inertial mass for a typical
rock is minute. Still, from a Newtonian perspective, the temperature of the
rock, or the amount of thermal energy it has released, is entirely irrelevant
to any consideration of its inertial mass.
With this rough understanding of Einsteins 1905 massenergy result,
we shall devote the remainder of this chapter to setting Einsteins result in its
historical and philosophical context. Our focus will be to step gingerly across
some of the stones in the river that separates Newtonian physics and Ein-
steins special relativity. Specifically, we wish to illustrate the accepted and
common usage of the terms mass and energy in physics near the begin-
ning of the 20th century so that we may better understand how Einstein might
have thought about his result and how others would have understood it.1

1.2 MASS: FROM NEWTON TO EINSTEIN

It is sometimes erroneously said that Einsteins famous equation entails


that matter can be converted into energy. Curiously, though perhaps

1 Readers who are familiar with this history, and especially with the distinction between inertial
and gravitational mass, may safely skip either to 1.4 or directly to the next chapter.
4 EINSTEINS MASSENERGY EQUATION, VOLUME I

understandably given the history of nuclear energy, seldom is it said that


energy can be converted into matter. Yet, the correlation between an objects
energy and its inertial mass in special relativity is clearly symmetric. Further-
more, there is ample evidence that, at least around 1905 and in the following
years that led to the development of general relativity in 1916, Einstein did
not interpret his result as best expressed by talk of converting matter into
energy. One of the main reasons for this is that Einstein focused on how
a change to an objects inertial mass correlates to a change in that objects
internal energy, where what is meant by inertial mass is not, as New-
ton would have said, a measure of the quantity of matter. We can further
appreciate this point by visiting both Newtons original definition of mass
and Machs famous criticism of it.

1.2.1 NEWTON AND MASS

In Definition 1 of Newtons Principia, he states, Quantity of matter is a


measure of matter that arises from its density and volume jointly [2, p. 403].
In the explanation of the definition, Newton states, I mean this quantity
whenever I use the term bodyor massin the following pages [2, p. 404].2
If we interpret this definition using our contemporary understanding of the
mentioned quantities and without attending to the context of the definition,
it is clear that Newton is defining the mass m as the product of the density
times the volume V . It is as if Newton is defining mass using the following
equation:

m = V . (1.1)

The circularity of this definition is patent, since density is routinely defined


as mass per unit volume, i.e., = m/V . One wonders how Newton himself
could have missed this circularity and how it could continue to be missed in
reformulations of Newtonian physics such as Eulers [3] and Maxwells [4].
It was Mach, in 1883, who famously pointed out the circularity when
he said:

With regard to the concept of mass, it is to be observed that the formu-


lation of Newton, which defines mass to be the quantity of matter of a
body as measured by the product of its volume and density, is unfortunate.

2 Note that this expression sediments the use of the term mass as a noun, such as when we
say a mass suspended from a string . . . .
INTRODUCTION AND CONCEPTUAL CONTEXT 5

As we can only define density as the mass of unit volume, the circle is
manifest [5, p. 194].

There are, then, two questions we wish to address: (1) Why would Newton
define mass in an apparently circular way? (2) How do Machs criticism and
his own attempts to define mass help us clarify the meaning of the concept
of mass? Let us address question (1) first.
In A Guide to Newtons Principia, I. B. Cohen gives a compelling
account of why the charge of circularity against Newtons definition of mass
is misplaced [6]. The first important lesson to glean from Cohens work is
that in an earlier draft of Book I of Principia, Newton clearly illustrates the
need to introduce a term other than weight to discuss a bodys resistance
to changes in its state of motion. According to Cohen, Newton was aware
that weight varies at different locations on earth. For Newton, weight
is due to the force of gravity acting on an object, and it is this weight that
Newton claims to have found, by experiment, to be proportional to a bodys
quantity of matter [6, p. 87]. This is significant because it signals that
prior to Newton there was not a single, widely used term with a prescribed
meaning to refer to what we now call inertial mass. In the scholium to the
definitions, Newton states Thus far it has seemed best to explain the senses
in which the less familiar words are to be taken in this treatise [2, p. 408].
At the time Newton wrote Principia, the term mass was in this category
of less familiar words whose meaning is in need of explanation.
When we turn to the explanation that directly follows Definition 1 in
Principia, we find three important aspects of the explication of the term
mass that merit our attention. First, Newton lists a variety of examples of
materials whose density is variable and for which the product of density times
volume determines the amount of stuff present in a container. Newton
explains:

If the density of air is doubled in a space that is also doubled, there is four
times as much air, and there is six times as much if the space is tripled.
The case is the same for snow and powders condensed by compression or
liquefaction, and also for all bodies that are condensed in various ways
by any causes whatsoever [2, p. 403].

The role of the term density in this explanation is significant. The examples
Newton provides involve cases in which one can reasonably imagine that
the matter in question is composed of atomic (or at least individualizable)
particles. This suggests a rather intuitive interpretation of the term density
whose meaning is roughly the number of bits of matter in a given space. With
6 EINSTEINS MASSENERGY EQUATION, VOLUME I

this interpretation, it seems reasonable to conclude that if we fill a chamber


with air so that it now has twice as many particles as before and then imagine
two such spaces filled with this much air, we would have four times as much
air. While this might only seem to apply to gases, Newton also means it to
apply to solids. This is evident when he adds:

For the present, I am not taking into account any medium, if there should
be any, freely pervading the interstices between the parts of bodies [2,
pp. 403404].

Newton seems to be treating all bodies as having parts and denying any-
thing like a Cartesian aether that fills the spaces between these parts. As
Cohen points out, Newton was a convinced atomist, and it is precisely this
atomism that gives Newton a way to conceive of density that is distinct from
our familiar definition of mass per unit volume [6, p. 95].
A closely related observation from Cohen about Newtons use of the
term density is that in Newtons time, densities were typically expressed
as ratios and not using some unit of density [6, p. 90]. The significance of
this is that not only can one state a ratio of two densities when considering
materials of the same physical composition, e.g., the density of snow in this
container as compared to the density of snow in that container, but densities
were treated, in effect, as specific gravities [6, p. 94]. One could, even in
Newtons time, use an experimental procedure to determine the density of
snow, e.g., relative to the density of water.
The significance of these observations about the term density
increases when we continue reading Newtons explanation to Definition 1.
Newton says:

It [the quantity mass] can always be known from a bodys weight, for
by making very accurate experiments with pendulumsI have found it
to be proportional to the weight, as will be shown below [2, p. 404].

In Book II, Proposition 24, Corollary 2, Newton states that if two simple
pendulums have the same period of oscillation, the masses of the bobs will
be directly proportional to their weights. Thus, if we can determine the
weight of a body, either by direct measurement or by calculation, we thereby
determine its mass especially if we are, as Newton is, concerned only with
relative weights and masses, which frees us from having to determine the
constant of proportionality between mass and weight.
As Cohen points out, Newton states only one way for determining the
mass of an object in Principia, by determining the bodys weight [6, p. 88].
INTRODUCTION AND CONCEPTUAL CONTEXT 7

The most important example of this in Principia is the determination of the


masses of the planets, which Newton achieves in Corollary 2 to Proposition 8
in Book III. Significantly, as Cohen points out more generally, the masses of
the planets are stated as ratios or proportions relative to the mass of the Sun.
Newton gives the masses of the Sun, Jupiter, Saturn, and Earth as 1, 1/1,067,
1/3,021, and 1/169,282. Newton arrives at this relative determination of the
masses of the planets by first determining the relative weights of the planets.
We can now appreciate why Cohens final conclusion is that the charge
of circularity in the definition of mass leveled against Newton is somewhat
misplaced. In providing his definition of the term mass, Newtons primary
goal was to express the meaning of an unfamiliar term using more commonly
used terms. The definition Newton gives, then, is not so much a prescrip-
tion for how to obtain an empirical measure of mass but rather, as Cohen
argues, how to understand the newly introduced term mass by appealing to
more familiar, intuitive, and undefined terms, viz., density and volume
[6, p. 92].
Despite Cohens compelling account of why the charge of circularity
in the definition of mass in Principia is misplaced, Newtons definition of
mass did not quite provided the type of conceptual clarity that came to be
expected in the development of physics particularly in the late 19th century.
However, those expectations were framed only after the careful development
of Newtonian physics. Two exemplary cases that speak directly to the devel-
opment of the concept of mass are Eulers Theoria Corporum Solidorum seu
Rigidorum (Theoria hereafter) and Maxwells Matter in Motion.3

1.2.2 EULER AND MASS

Eulers development of Newtonian physics in his Theoria, which contains a


conceptually more nuanced and developed exposition than his Mechanica,
is far closer to a contemporary presentation than Newtons own exposition
in Principia. For example, contemporary readers of Principia are familiar
with the observation that Newton does not state his second law of motion
as F = ma. Instead, in Definition 2, Newton defines quantity of motion
(which we now call the momentum and designate p) as the product of the
velocity and quantity of matter or mass. Newton simply assumes that the
meaning of the term velocity is well-enough understood. In his second
law of motion, Newton states:

3 For a more detailed description of the history of the concept of mass, see [7].
8 EINSTEINS MASSENERGY EQUATION, VOLUME I

A change in motion is proportional to the motive force impressed and


takes place along the straight line in which that force is impressed
[2, p. 416].

Two things leap to the contemporary eye about Newtons own statement
of his second law. First, Newton is effectively defining a vector quantity.
Second, and more importantly, Newton is stating that the temporal rate of
change of momentum is proportional to the impressed force, a term he defines
in Definition 4. In our contemporary notation, the second law states:

dp
F. (1.2)
dt
As we have noted, because Newton worked exclusively with ratios of quan-
tities of the same type, he does not need to concern himself with determining
the constant of proportionality.
At a more conceptual level, the main difference between Eulers pre-
sentation and Newtons is that Newton begins his physics with definitions
of dynamical quantities. For example, Newtons first five definitions define
mass, momentum, force of inertia, impressed force, and centripetal force.
In contrast to Newton, in Theoria Euler begins his development of mechan-
ics with kinematical definitions and theorems. For example, Eulers first
five definitions define, for a moving point construed abstractly without any
physical properties, rest and motion, relative rest and motion, rectilinear and
curvilinear motion, and uniform and nonuniform motion. Significantly, Def-
inition 6 defines velocity as the ratio of the distances traversed to the times
[3, p. 11]. In his explanation to this definition, Euler is careful to point out
the legitimacy of dividing a distance by a time, something which as we have
seen was quite foreign at the time of Newton.
Eulers focus in the second chapter of Theoria, titled the elements of
internal motion, is three axioms that constitute a more refined assertion
of Newtons first law of motion combined with the views Euler shares with
Newton about the absolute nature of space and time. Euler states:

Axiom 1: Every body, even without being relative to other bodies, either
remains at rest or is moving, that is, it is either at absolute rest or in
absolute motion [3, p. 54].

Axiom 2: A body, which is absolutely at rest, if subjected to no external


actions, will persist indefinitely in a state of rest [3, p. 56].
INTRODUCTION AND CONCEPTUAL CONTEXT 9

Axiom 3: A body, which is moving absolutely, if subjected to no exter-


nal actions, continues to progress uniformly along the same direction
[3, p. 57].

At this stage, unlike Newton, Euler is not using the term body as synony-
mous with mass. Euler treats the meaning of the term body as intuitively
understood and then goes on to define the term inertia in Definition 11.
Euler states:

Definition 11: That quality of bodies, the reason for persisting in the same
state present within a body itself, is called inertia, and also sometimes
the force of inertia [3, p. 59].

In the two corollaries to Definition 11, Euler explains that inertia is the cause
of a body remaining in its state of motion and another cause besides the
inertia of this body cannot be assigned [3, p. 60].
Philosophically, one feels the strength of Eulers convictions concerning
inertia as the singular cause of a body persisting in its state is unsupported by
the purported conceptual clarity actually delivered by the definition of inertia
and its attending corollaries and scholium. If, in a metaphysical mood, we
ask What is inertia?, Eulers definition leaves us wanting, for it only tells
us inertia is an unknown property of objects that results in them persisting in
a uniform state of motion (including the state of rest). Euler also tells us, it
is true, that inertia is the only cause of bodies persisting in a state of uniform
motion, though we are left wondering why this must be so.
However, there is a more refined way to approach Eulers definition of
inertia. The term inertia functions, in Eulers Theoria, as a placeholder
for whatever it is in physical bodies that leads them to exhibit the idealized
behavior described in Eulers three axioms. While we do observe objects at
rest remaining at rest unless acted upon by a net impressed force, we tend
not to witness directly terrestrial objects moving with a constant nonzero
velocity. If the term inertia merely functioned as a name for that unknown,
hypothetical property of objects, one could argue that merely having a name
for that unknown property does not increase our understanding very much.
Yet, there are two reasons why Eulers careful treatment of inertia goes
beyond merely the selection of a name.
First, Euler improves our conceptual clarity of the foundations of
mechanics by introducing the term inertia as the single name for the
unknown property of objects to persist in the same state of motion. Signifi-
cantly, in the scholium to Definition 11, Euler states that we should reserve
the term force for causes that change the state of a body. Consequently,
10 EINSTEINS MASSENERGY EQUATION, VOLUME I

since inertia is not what changes the state of a body but instead causes
it to persist in its state of uniform motion, Euler recommends that we stop
using the phrase force of inertia or vis inertiae as it engenders confusion
[8, p. 36].
Second, if Euler was merely adding to or modifying the vocabulary we
use to give qualitative descriptions of physical objects and processes, we
may have well chosen to ignore his suggestion. However, the significance
of the term inertia in a system of quantitative physics, and the importance
of not construing inertia as a force, is that inertia is a measurable quantity.
Precisely how we can measure inertia and how it is related to the notion of
a force is Eulers main focus in Chpt. 3 of Theoria.
For Euler, a force is the name given to that which prevails to change
the absolute state of bodies [3, p. 81]. If we interpret Euler as only picking
a name, then again it seems not much is gained. However, the measur-
able, quantitative relationships between instances of the term force and
especially the term inertia (or mass) solidify the meaning of the latter.
For example, in Theorem 3 of Chpt. 3, Euler proves the following:

If equal forces should act on unequal bodies at rest, the effect to be


produced in the same small time intervals are inversely proportional to
the inertias of the bodies [3, p. 94].

Euler proves this theorem constructively. First, he asks us to consider a


corpuscle A that begins at rest and is moved by a force F for a time dt.
Let us suppose that during this time, the object A is displaced a distance
ds. Now suppose we combine n such corpuscles into a compound object
nA. To displace this object the same distance ds in the same time dt would
require a force nF. If, however, we only apply a force F to the object nA in
the time dt, it will only move a distance ds/n. If we now consider an object
kA composed of k corpuscles each of which is identical to A, applying the
same force F to the object kA for a time dt would result in a displacement
ds/k. Consequently, if we apply the same force F for the same amount of
time dt to the two objects nA and kA, the ratio of their displacements is
inversely proportional to the ratio of their inertias, i.e., using sn and sk for
the displacements of nA and mA, respectively, we have:
sn k
= . (1.3)
sk n
Significantly, while the inertia of each body is originally given as a multiple
of the inertia of the corpuscle A, the latter does not appear in Eqn. (1.3).
INTRODUCTION AND CONCEPTUAL CONTEXT 11

In the explanation that follows the proof of Theorem 3, Euler sum-


marizes a collection of claims about the inertia of bodies that ultimately
determine the arithmetical rules by which instances of the term mass are
combined and characterize mass as a measure of a bodys resistance to
changes in state of motion. Anticipating Definition 15, let us designate the
inertia of a body by the letter m. The first implicit claim, which Euler seems
clearly to assume especially since he still treats inertia as a measure of the
quantity of matter, is that any material body has a positive nonzero inertia,
i.e., m (0, ).4 In the scholium to Definition 15, Euler clearly states that
inertia is a property of matter and so is connected with solidity or impene-
trability. Any solid object with some finite extension must have inertia. For
Euler, there cannot exist matter, i.e., anything with inertia, that does not have
extension, where by extension Euler is alluding to the Cartesian notion of
an object possessing spatial dimensions.
In the explanation to Theorem 3, Euler then suggests:

1. n bodies each of inertia m when combined have inertia n m, i.e.,


instances of inertia are additive; they follow the ordinary rules of
addition. Implied is also the view that if we started with a body that
consisted of n bodies each of inertia m and then removed k such
bodies, the resulting body would have an inertia (n k) m.
2. If twice as much force is required to move an object through a given
distance ds in a time dt as another object, then the former has twice
as much inertia as the latter. In effect, this establishes the direct
proportionality of force and inertia, i.e., F m.
3. By combining n and k multiples of a body A (with n  = k), we can
create two bodies whose inertia is in any ration n/k, where n and k
are positive integers.

From these assertions, it is clear, Euler points out, that the magnitude of
inertia is central to mechanics and merits its own name, which he then
specifies in Definition 15:

Definition 15: The mass of a body or the quantity of matter is called the
amount of the inertia which is present in that body, by which just as it
tries to continue in its own state so it tries to resist all changes [3, p. 95].

4 The precise mathematical details concerning instances of m are a bit vague in Euler. For
example, we have deliberately left the set of numbers to which m belongs unspecified. Because
Euler is committed to a form of atomism, he treats mass ratios as rational numbers. Any given
mass would then be a positive integer multiple of some unit mass. This would entail that an
objects mass could never be a nonrational, real number, such as .
12 EINSTEINS MASSENERGY EQUATION, VOLUME I

In this definition, Euler is, on the one hand, introducing the term mass
exclusively as a measure of a bodys resistance to changes in state of motion.
On the other hand, Euler is using this definition as a bridge to the prior,
Newtonian notion of the mass of a body as related to the quantity of matter,
which is reasonable since for Euler also it would be difficult to imagine
anything contributing to the mass of an object other than what is intuitively
regarded as a material substance.
However, at a conceptual level, Euler has achieved something deeply
important to future physics: he has divorced the term mass from the intu-
itive notion of a measure of how much stuff there is in an object. We can
appreciate this by noticing how Eulers treatment of the term mass makes
it easy to operationalize the term. As Maxwell would later point out (see
1.2.3), assuming that we can find an empirical procedure for applying the
same force to two objects, if we apply the same force for the same amount of
time to two objects, we can determine the ratio of their masses by measur-
ing the ratio of the velocities they acquire. Consequently, the term mass
in Euler comes to designate an empirically measurable quantity because
of its proportionality to force, which follows familiar arithmetical rules for
combining instances of the term.
Finally, it is in Problem 9 that Euler comes close to writing F = ma.
Using contemporary notation, the problem is to determine the change in
motion for an object of mass m that is originally moving with a velocity
v when a force F acts in the same direction as v while the object moves
through a distance ds in a time dt. Euler clearly states that once the force F
2
begins to act on the body, its acceleration ddt 2s is a positive number that is
proportional to F/m. However, since Euler does not have a system of units
at his disposal, the most he could write is:

d 2s F
. (1.4)
dt 2 m
If we do assume, Euler suggests, a reference body of mass that serves as
our unit of mass, then we can write:

d 2s F
2
= , (1.5)
dt m/
or equivalently,

F = m a, (1.6)

d2s
where a = dt 2
and m = m/.
INTRODUCTION AND CONCEPTUAL CONTEXT 13

Thus, it is these considerations from Theoria that support Jammers


contention that Eulers development of Newtonian mechanics constitutes an
important transition in the history of the term mass. Euler treats mass, as
we now do, as a numerical coefficient, characteristic of the individual body
and determined by the ratio of force to acceleration [7, p. 87]. Although he
would not have anticipated it, given his own philosophical views about the
nature of matter, Eulers work makes it possible to treat the term mass phe-
nomenologically, i.e., as a term that designates an empirically determinable
quantity strongly correlated to individual bodies. The origins of mass,
from an Eulerian perspective, is something that need not detain us in our use
and development of a system of mechanics. We can safely leave any effort
to determine the origins of inertial mass to future scientific theorizing.

1.2.3 MAXWELL AND MASS

Eulers work draws our attention to what appeared to some, e.g., Mach [9],
to be a new circularity in the definition of mass. The mass m is defined as the
ratio of force F to acceleration a. However, to define a force F, one simply
writes F = ma (see, e.g., [9, p. 82]). Contemporary approaches in rational
mechanics circumvent this apparent circularity by noting, as Truesdell does,
that the term force is only implicitly defined, in the way that terms such
as point and line are in Euclidean geometry, by the particular collection
of axioms in which it appears [10, p. 129]. However, Maxwell attempted to
avoid the apparent circularity by assuming the notion of force as a primitive
term in his Matter and Motion [4]. Maxwells assumption is grounded on the
empirical possibility of applying the same force in different circumstances.
Furthermore, Maxwells work is of interest to us because his presentation
of Newtonian mechanics in Matter and Motion is deeply influenced by
developments in, and refinements to, the concept of energy in the mid to late
19th century.
After introducing Newtons second law of motion in 44 of Matter and
Motion, Maxwell immediately recognizes the need to define equal masses
and equal forces in 45 [4, p. 32]. Although he still talks about mass as the
quantity of matter, Maxwells definition of equal masses studiously avoids
this phrase. The only assumption we need to make to have such a definition,
Maxwell notes, is that objects have permanent properties.
Maxwell asks us to consider a thread of natural rubber (caoutchouc).
When stretched beyond a certain point, such a thread will exert a certain
force. If we stretch the thread to the same distance on two separate occasions,
14 EINSTEINS MASSENERGY EQUATION, VOLUME I

and assuming that its properties remain constant, it will exert the same force.
From these elementary observations, which Maxwell treats as primary and
not in need of further explanation at least for the purposes of developing
the science of abstract dynamics, the definition of equal masses quickly
follows.
Suppose we have two objects M and N , Maxwell suggests. We attach
M to our rubber thread and pull on the thread. Assuming that no other force
is acting on M , a specific force F will act on M and accelerate it so that after
one unit of time it will acquire a velocity v. Now suppose we perform the
same operation with N and that at the end of one unit of time N also acquires
a velocity v. We can then conclude that M and N have the same mass [4,
p. 33]. Although Maxwell does not state it explicitly, we could say that two
objects M and N have equal mass if, and only if, after being acted upon
by the same force F for one unit of time, they acquire the same velocity v.
Again, while this definition of equal masses clearly depends on the definition
of equal force, it is not circular because the latter is operationally defined.
For Maxwell, despite having this definition of equal masses in hand,
it is nevertheless important to establish clearly its connection to methods
commonly used to measure mass. Maxwell explains that if we are dealing
with bodies of exactly the same kind [4, p. 33], by which he means bodies
of the same chemical composition, it is clear that we can use several methods
for comparing masses. In general, Maxwell says,

If equal quantities of a the substance produce equal effects of any kind,


we may employ these effects as measures of the quantity of the substance
[4, p. 33].

For instance, Maxwell says, if we wish to compare different quantities of sul-


furic acid, we can measure the mass of each sample by weighing it, measuring
its volume, or even measuring how much potash it neutralizes. However, we
cannot employ these methods to compare the masses of objects of differ-
ent types unless we make additional assumptions, which may or may not
have empirical evidence to support them. This is brought to the foreground
most clearly perhaps with the commonly used method of ascertaining mass:
weighing (with a balance).
Without using the labels, Maxwell clearly articulates the difference
between inertial mass and what we today would call (passive) gravitational
mass. First, Maxwell points out that in abstract dynamics we are exclu-
sively concerned with mass as a measure of how a bodys motion changes
as the result of an applied force. When a body falls toward the surface of
the earth, it is attracted by the force of gravity. However, it is an empirical
INTRODUCTION AND CONCEPTUAL CONTEXT 15

fact, and not something derived strictly from the three laws of motion, that
objects of different chemical composition experience the same acceleration
toward the earth. Maxwells discussion is worth quoting in full:

It is an observed fact that bodies of equal mass, placed in the same posi-
tion relative to the earth, are attracted equally towards the earth, whatever
they are made of; but this is not a doctrine of abstract dynamics founded
on axiomatic principles, but a fact discovered by observation, and ver-
ified by the careful experiments of Newton, on the times of oscillation
of hollow wooden balls suspended by strings of the same length, and
containing gold, silver, lead, glass, sand, common salt, wood, water, and
wheat [4, p. 34].

Although Maxwell does not explicitly mention it, the key lies in recognizing
that although two objects M and N may have equal mass as determined by
the empirical procedure Maxwell outlines to define equal masses, it need not
be the case that M and N respond (or couple) equally to noncontact forces
such as the force of gravity or, say, the magnetic force. It is an empirically
well-confirmed observation that objects that have the same inertial mass,
i.e., equal mass as Maxwell defines it, can be accelerated quite differently
by the same magnetic force. Maxwells point is just that a similar thing
could have been observed about gravitation. As it happens, this does not
turn out to be the case. Instead, it is a fact that in the same geographical
position the weights of equal masses are equal [4, p. 34], and it is this
fact, Maxwell explains, that underwrites our use of weighing as a measure
of mass.

1.2.4 MACH AND MASS

Mach does not directly address Maxwells definition of mass, though in his
Science of Mechanics, he is famously critical of Newtons definition. As
early as 1867, 16 years before the publication of the first German edition
of the Science of Mechanics, Machs article On the definition of mass
already included his unique definition of mass, though that article, which
seems to have rather different philosophical motivations, does not include
the explicit criticisms of Newton. We will here briefly review both Machs
criticisms and his definition of mass not so much because it enhances the
conceptual clarity of the foundations of mechanics but because it reinforces
the important distinction between inertial and gravitational mass that was so
influential to Einstein in his quest for a relativistic theory of gravitation that
was sensitive to the massenergy relation.
16 EINSTEINS MASSENERGY EQUATION, VOLUME I

In his Science of Mechanics, Mach levels three main criticisms against


Newtons, or even just a Newtonian, definition of mass. First, if we look
at Newtons own definition, Mach says, we see that he defines mass as the
product of volume and density. However, Mach says, as we can only define
density as the mass of unit of volume, the circle is manifest [5, p. 194]. As
we have already discussed, this charge of circularity may not have the force
Mach hoped.
Machs second criticism is that the notion of mass as quantity of matter
is effectively meaningless. He says:

In the first place we do not find the expression quantity of matter


adapted to explain and elucidate the concept of mass, since that expression
itself is not possessed of the requisite clearness [5, p. 216].

Mach goes on to explain that about the only way one can imagine the phrase
quantity of matter to be meaningful is if one interpreted it to mean, roughly,
the number of atoms of a given substance in a given sample. In this descrip-
tion, by atoms one would have to mean this term in its strict philosophical
sense, i.e., the smallest, finite, indivisible amount of stuff there is. Famously,
Mach rejected any such notion because of his unique empiricist epistemol-
ogy, which was intimately wedded to a unique brand of sensationalism, i.e.,
very roughly, the view that the only things we can know directly are the
sensory contents of our own experience. However, Mach does not reject
the concept quantity of matter on these epistemological grounds. Instead,
Mach argues that if we lived in a world in which matter was homogeneous,
i.e., a world in which there was only one kind of stuff, then the phrase
quantity of matter would be meaningful indeed. Mach continues:

But the moment we suppose chemical heterogeneity, the assumption that


there is still something that is measurable by the same standard, which
something we call quantity of matter, may be suggested by mechanical
experiences, but is an assumption nevertheless that needs to be justified
[5, pp. 216217].

It is the empirical fact of the chemical heterogeneity of the physical world


that precludes a clear conception of quantity of matter when we attempt
to construe it as the number of atoms of stuff in a sample.
Machs third criticism of the Newtonian concept of mass is, like the
second, not pointed directly at Newton but rather at any approach that is
broadly speaking Newtonian. For suppose, Mach suggests, that one con-
siders defining the ratio of the masses of two objects m/m as equal to the
ration of the gravitational forces acting on these two bodies. In Newtons
INTRODUCTION AND CONCEPTUAL CONTEXT 17

physics, this is, of course, a consequence of the law of gravitation. However,


for Mach, if we define mass using the force of gravity, then it is not at all
clear how masses defined in this way will enter into the action-reaction law,
i.e., Newtons third law of motion.
This third criticism is perhaps the most compelling, and it is certainly
more compelling than Machs first criticism. As we have seen, the equality of
inertial and gravitational mass in Newtonian physics is based on a contingent
fact about gravitational interactions in our world. Just as we cannot assume
that two objects with the same inertial mass necessarily have the same grav-
itational mass, we cannot assume the converse either. Clearly, just as it is the
inertial mass that enters into the second law of motion, it is the very same
inertial mass that appears in the third law of motion. So, while it is true that
Mach may be demolishing a straw figure here, it is one worth demolishing.
Machs definition of mass is, implicitly at least, meant to overcome these
criticisms, which in a way it manages to do though at the risk of falling into
other traps. In his earlier article On the definition of mass Mach proposes
the following:

Definition.Bodies which communicate to each other equal and opposite


accelerations are said to be of equal mass. We get the mass-value of a
body by dividing the acceleration which it gives the body with which we
compare others, and choose as the unite, by the acceleration which it gets
itself [9, p. 85].

Imagine, then, that we have arranged things so that two interact bodies by
whatever means. For Mach, it is an empirical fact, which he calls a Theorem
of experience that,

Bodies placed opposite to one another communicate to each other accel-


erations in opposite senses in the direction of their line of junction
[9, p. 85].

So, if under these conditions, which are much more general than discussions
about gravitation as the interaction in question is completely unspecified,
impart an equal acceleration to each other, then they have the same mass.
This, in effect, is what Mach calls his definition of mass in the Science of
Mechanics, which he states as follows:

Definition of equal masses.5 All those bodies are bodies of equal mass,
which, mutually acting on each other, produce in each other equal and
opposite accelerations [5, p. 218].

5 This phrase appears as marginalia.


18 EINSTEINS MASSENERGY EQUATION, VOLUME I

Finally, if the accelerations of the two bodies are not equal, then their masses
are in the inverse proportion to their accelerations.
Machs definition is not without its challenges. For example, as Koslow
has reported, one of the criticisms leveled against Mach is that his defi-
nition of mass presupposes the definition of an inertial frame [11, p. 223
ff.]. However, to define an inertial frame one needs to appeal to the notions
of force and mass. Koslow concludes that this charge of circularity is mis-
placed, and he may be correct. If he is not, it may be that there is a cluster of
terms, including mass, force, inertia, and acceleration, at the foun-
dations of mechanics that constitute a definitional circle that commands that
we make a conventional decision about which one(s) to treat as primary.
Conventional, of course, does not mean arbitrary if, as we have seen in the
case of Maxwell, we have empirical procedures for identifying instances
of one of the terms by ostension. Still, the most important lesson to draw
from these discussions for our purposes is that these attempts to define mass
consistently improved our understanding of the distinction between inertial
mass and gravitational mass.

1.3 ENERGY AND ITS CONSERVATION IN THE


19TH CENTURY

The concept of heat, and the development of classical thermodynamics in the


19th century, unified a variety of different branches of physics and effected
a progressive and irreversible change in physics. The generalized concept
of energy, as opposed to the concept of mechanical energy implicitly con-
tained in, e.g., Leibnizs notion of vis viva or living force, grew out of the
gradual recognition, through the insights of several prominent physicists,
that thermal energy was interchangeable with mechanical, chemical, and
electromagnetic energy. Contributions, which would require more than one
excellent book such as Crosbie Smiths The Science of Energy to review,
came from at least the following natural philosophers: Sadie Carnot, Rudolf
Clausius, James P. Joule, William Thomson (Lord Kelvin), Macquorn Rank-
ine, James Clerk Maxwell, P. G. Tait, Hermann Helmholtz, Julius R. Mayer,
and Ernst Mach.
As Smith has observed, some of these developments in studies of heat
and energy clearly were intermingled with pragmatic interests related to the
development of efficient machines, specifically, the machines that drove the
industrial revolution particularly in northern England, where the engines for
steamers that would cross the seas were also being developed [12]. Even
INTRODUCTION AND CONCEPTUAL CONTEXT 19

a cursory glance at documents from this time also reveals other intrigues
in the development of both thermodynamics in general and the principle of
conservation of energy more specifically.
For example, in the preface to his Sketch of Thermodynamics, after
explaining that part of his goal is to correct the erroneous reports other have
given of the history of the subject, P. G. Tait says:

I cannot pretend to absolute accuracy, but I have taken every means of


ensuring it, to the best of my ability; though it is possible that circum-
stances may have led me to regard the question from a somewhat too
British point of view. I have tried to preserve a happy medium between
the old absurd British contempt for all things foreign, and the still more
absurd custom (of comparatively recent origin) which gives to foreigners
all that the writer is not in a position to claim for himself or for some of
his particular friends [13, p. viii].

Although these remarks may seem rather quaint to us today, Tait may have
had good reason for making them in the preface to his book. For as he himself
reports in the same preface, Clausius had accused Tait of misrepresenting
Clusius work in an earlier article. Tait then explains:

In the new edition of his Abhandlungen recently published, Professor


Clausius has modified this accusation to the charge that my book was
written with the view of claiming the Dynamical Theory of Heat as far
as possible for the British nation [13, p. xv].

Yet, despite all the nonscientific debates about priority and intellectual prop-
erty, tainted as they were with nationalist tendencies and proclivities, not
to mention the very unique, if not quite eccentric, characters of the scien-
tists involved, a gradual convergence to a single concept of energy and its
attendant conservation principle took place.
By the close of the 19th century, Ernst Mach could confidently assert that
to-day the law of conservation of energy is accepted as a fully established
truth and receives the widest applications in all domains of natural science
[14, p. 22]. In the same article from 1894, in which Mach attempts to describe
the genesis of both the concept of energy and its associated conservation
principle, Mach explicitly recognized that the principle of conservation of
energy was outside mechanics. Mach says:

The most multifarious physical changes, thermal, electrical, chemical


and so forth, may be brought about by mechanical work. When such
20 EINSTEINS MASSENERGY EQUATION, VOLUME I

alterations are reversed they yield anew the mechanical work in exactly
the quantity which was required for the production of the part reversed.
This is the principle of conservation of energy; energy being the term
which has gradually come into use for the indestructible something of
which the measure is mechanical work [14, p. 23].

Despite Machs unique philosophical position, which leads him to remain


somewhat ambivalent about what energy is, though he leans toward a sub-
stantial conception of energy (see, e.g., [14]), Machs view concerning the
generality of the principle of conservation of energy were already widely
shared. For example, as early as 1863, in the introduction to a series of lec-
tures at Carlsruhe (now Karlsruhe) in Germany, which was published as an
English translation with the title On the conservation of force, Helmholtz
announces what he calls a new universal law of all natural phenomena by
stating:6

The law in question asserts, that the quantity of force which can be brought
into action in the whole of Nature is unchangeable, and can neither be
increased nor diminished [15, par. 8].

Similarly, in 1877, in his Sketch of Thermodynamics, P. G. Tait says:

The principle of Conservation of Energy asserts that the whole amount


of energy in the universe, or in any limited system which does not receive
energy from without, or part with it to external matter, is invariable
[13, p. 70].

Thus, by the time Einstein was preparing his work on Special Relativity, it
was justifiable for him to regard the principle of conservation of energy as an
uncontroversial and well-established physical principle, i.e., a principle that
was empirically well-confirmed in a variety of different types of physical
scenarios and could be treated as if it was true of all physical systems.7
However, before turning to Einsteins discovery of the massenergy
relation in special relativity, there are three philosophical issues surrounding

6 The word force in both the title of Helmholtz article and throughout his text is an English
translation of the German word Kraft. Helmholtz explicitly states that the quantity of
Kraft is the amount of work. Consequently, from our contemporary perspective, the word
Force is not the best translation of Helmholtzs Kraft as he is clearly articulating a version
of the principle of conservation of energy.
7 This description of treating a statement as if it is true is consistent both with Einsteins
description of physical science in his letter to Solovine from 1952 [16] and with how he
speaks of taking an assertion and raising it to the status of a postulate in his 1905 paper on
relativity [17].
INTRODUCTION AND CONCEPTUAL CONTEXT 21

the principle of conservation of energy, and energy itself, that merit our
attention. These issues will figure, albeit only implicitly for the most part,
in our discussions in later chapters.
First, a significant feature of the principle of conservation of energy,
which was clearly recognized by natural philosophers in the 19th century, is
that it is a principle that is above or outside mechanics, though it clearly
applies to mechanics. Similarly, energy, unlike the inertial mass, was also
clearly recognized as a concept that is applicable not only to mechanics, but
to description of a wide range of physical processes. Consequently, the grad-
ual acceptance of the principle of conservation of energy, along with other
concomitant factors such as the development of Lagrangian and Hamilto-
nian mechanics, began to shift the attention of some physicists away from a
Newtonian program toward a new way to do physics.8
In the Preface to Principia, Newton explains that the aim of his book,
which sets out the mathematical principles of natural philosophy, is to
discover the forces of nature from the phenomena of motions and then to
demonstrate the other phenomena from these forces [2, p. 382]. By phe-
nomena Newton here means roughly first-level generalizations of empirical
data. For example, the Keplerian orbits of, say, the moons of Jupiter are a
good example of Newtons phenomena, because they are generalizations
induced from a finite set of observations (see [18]). The study of heat, how-
ever, resisted any explanation of phenomena by appeal to forces. Even when
heat became identified with motion, the task of finding some specific force
law to explain a phenomenon related to heat became otiose. Furthermore, the
identification of heat with work combined with the realization that one could
be certain that the quantity of work was conserved in any kind of physical
interaction regardless of whether one knew the nature of the specific forces
at work again shifted attention away from the Newtonian program of find-
ing specific force laws. A physics that was founded on physical principles,
which in a sense were more general than Newtons three laws of motion,
was quickly developing.
A second notable aspect of the analysis of physical systems from the
point of view of energy is the introduction of new conceptual tools, evident
in the language used by physicists, that were deployed in the description of
physical systems. There are two key notions that, with the hindsight of his-
tory, we can see had a deep influence in how the massenergy relation was
discussed in the early 20th century: (1) the notion that one type (or form)

8 Koslow makes a similar claim though he tends to emphasize less the shift in certain areas of
physics toward what Einstein would later call principle theories [11]. See 2.9.
22 EINSTEINS MASSENERGY EQUATION, VOLUME I

of energy can be transformed into another type of energy (2) the notion
that heat is equivalent to work. Energy, it was realized, could be manifest,
as the capacity to do work, in a variety of forms. Chemical energy, e.g.,
could be transformed into thermal energy which, in turn, could be trans-
formed into mechanical energy. Energy in one form, e.g., thermal energy,
could disappear only to be replaced by a different form of energy that
appeared. While this language was rather imprecise in meaning, it nev-
ertheless continued to influence not only discussions about energy, but, not
surprisingly, discussions about Einsteins massenergy relation. Similarly,
though without further research this is a bit more speculative, talk about the
equivalence of different types of energy may have influenced Einsteins
own way of talking about the massenergy relation especially during the
early days of his discovery.
Third, it bears mentioning that, although the issue had been largely
settled by Einsteins time, the epistemological standing of the princi-
ple of conservation of energy had been rather controversial. As Smith
reports, soon after Rankine introduced the phrase potential energy, the
astronomer royal Sir John Herschel argued that the principle of conserva-
tion of energy was a tautology [12, p. 6]. Rankine responded that while the
terms actual energy and potential energy made the principle of con-
servation of energy appear to be a tautology, it was not. The key lies in
noticing, Rankine suggested, that we can measure both the actual and poten-
tial energy in a physical system independently and verify empirically that
the law of conservation of energy is indeed confirmed. Thus, the law of
conservation of energy is not merely a bookkeeping device but a genuine
empirical law.
A separate epistemological question, which Einstein often mentioned in
a variety of his popular writings and letters, concerned whether the principle
of conservation of energy should be treated as an a priori principle, in the
philosophical sense of the term. There were, as expected, proponents on all
sides of this question. For example, J. R. Mayer held that the principle of
conservation of energy was ultimately a direct consequence of the meta-
physical principle that effects are always equal to their causes. The latter,
philosophers such as Mayer supposed, is something we can know about the
physical world simply by using our intellect and thinking clearly; it is an
a priori truth. On the other hand, natural philosophers such as P. G. Tait
were staunch empiricists who shared the view of the 18th-century philoso-
pher David Hume, who revered Newtons work, that any statement about
the physical world must receive its evidence from the physical world and
hence cannot be a priori. Finally, Mach, whose views we know influenced
INTRODUCTION AND CONCEPTUAL CONTEXT 23

Einsteins, held a sort of post-Kantian empiricism. For Mach, Mayer is cor-


rect to regard the principle of conservation of energy as based on more
general principles, such as the ex nihilo nihil fit principle, which have the
appearance of being metaphysical principles. However, Mach argues, such
principles are merely an expression of the formal need of a simple, clear,
and living grasp of the facts, which receives its development in practical
and technical life, and which carry over, as best we can, into the province of
science [14, p. 53]. Thus, such principles, according to Mach, are merely
ways of ensuring the economy of our thought as we try to find the patterns
in our sensations that eventually we call scientific laws.

1.4 PHILOSOPHICAL REFLECTIONS

The philosophical significance of our discussions about mass and energy in


the 19th century is threefold. First, Einstein found himself in a tradition in
natural philosophy that had begun to outgrow the Newtonian program of
finding specific force laws to describe all known physical interactions. The
novel concept of energy, and its attendant conservation principle, held the
promise of unifying even more diverse areas of physics than Newton would
have probably imagined. In practice, this unification was brought about, as,
e.g., in the case of classical thermodynamics, by subsuming a wide range of
phenomena under a small collection of physical principles. As we will see,
Einstein was impressed by this aspect of classical thermodynamics, which
inspired him to frame his distinction between principle and constructive
theories (see 2.9).
Second, Einstein found himself at the height of a long tradition, cul-
minating in the philosophical work of Ernst Mach, of dedicated efforts to
fortify the foundations of physics by attending to the meanings of its cen-
tral terms. Curiously, as far as we have been able to determine, Einstein
himself nowhere addresses the definition of mass, though he clearly and
famously addresses the definitions of terms at the foundations of kinematics
(see 2.1). For Einstein, it seems, the distinction between inertial mass and
gravitational mass, combined with the separation of the notion of quantity
of matter from the notion of inertial mass, and the rejection of the former,
seemed to be enough.
Third, by the end of the 19th century, heat was no longer regarded as a
substance especially because of what was then often called the equivalence
between mechanical work and heat [19, p. 9]. Instead, heat had been iden-
tified as just one form of energy that could be transformed into a different
24 EINSTEINS MASSENERGY EQUATION, VOLUME I

form of energy. For example, Tait states:

The Transformation of Energy is the enunciation of the experimental


fact, that in general any one form of energy may by a suitable process
be transformed, wholly or in part, to an equivalent amount in any other
given form [13, p. 70].

Energy, then, was construed as a property of physical systems and not as


a thing. Thus, as with inertial mass, the connection between energy and
metaphysically suspect entities, such as Blacks caloric, was severed in no
small measure because of efforts to understand the meaning of the central
terms in natural philosophy.
INDEX

A E
Abstract dynamics, 14 E = mc2 . See Massenergy relation
Act of emission, 3336, 46, 5051, Einstein, A., 6979
91, 93 Electromagnetic phenomena,
Atoms, 16, 58 2627, 66
Electron, 47, 50
B kinetic energy of, 51
Blacks caloric, 24 Elements, 63
Energy. See also Massenergy
C relation
Cause and effect, 96 and its conservation in 19th
Center-of mass theorem, 44 century, 1823
Chemical energy, 22 philosophical reflections, 2324
Classical thermodynamics, 18, 62 transformation, 22
Composite systems Equivalence of mass and energy, 59
and inertia of energy, 5758 Euler, L., 713
philosophically speaking, 5961
Conservation laws, 80 F
Constructive theories, 61 Feigenbaum, M., 9197
Einsteins principle vs., 6164 First demonstration, of
massenergy relation, 3142
D Force, 10
Density, defined, 45 Force of inertia. See Inertia
Dynamic demonstration of
massenergy relation
Einstein, A., 6979 G
philosophically speaking, 79 Gas container, analysis of, 6061
Feigenbaum, M. and Mermin, Gravitational field equation, 66
N. D., 9195
philosophically speaking, 9597 H
Lewis, G. N. and Tolman, R. C., Heat, concept of, 18
8091 Hume, David, 96
104 INDEX

I Feigenbaum, M. and Mermin,


Inelastic collision, 77 N. D., 9197
Inertia, 910 Lewis, G. N. and Tolman, R. C.,
and energy, 55 8091
Inertial mass, 35, 14, 17, 3536, early demonstrations, 25
4142, 51, 5361, 73, 77, 92 Einsteins principle vs.
Inertia of energy, 42, 46, 50, 5154, constructive theories, 6164
5657, 73, 78 first demonstration, 3142
composite systems and, 5758 philosophically speaking,
Internal energy, 1, 4, 32, 58, 5357
60, 73 second demonstration, 4246
special relativity, historical
K review of, 2631
Kant, Immanuel, 6263 third demonstration, 4653
Kinetic energy, 72, 86, 96 Maxwell, J. C., 1315
Maxwells theory, 56, 65, 68, 79
L Mayer, J. R., 63
Law of conservation of Mermin, N. D., 9197
energy, 19 Minkowski spacetime (MST),
Lewis, G. N., 8091 6569, 96
Light principle, 27, 42, 45, 53, 56, Momentum, 70
57, 63, 80 MST. See Minkowski spacetime
London Times, 61, 62
LorentzFitzGerald hypothesis, 29
Lorentz transformations, 31, 46, 67, N
69, 70, 72, 77, 79 Natural philosophy, 2324
Newton, I., 47
M Newtonian kinetic energy, 49
Mach, E., 1518 Newtonian limit, passage to, 3639,
Magnitude of inertia, 1112 5253
Mass, 1213. See also Newtonian theorem, 42
Massenergy relation Newtons second law of motion, 7,
Euler L., 713 13, 17
Mach, E., 1518 Non-Newtonian Mechanics, 81, 84
Maxwell, J. C., 1315
Newton, I., 47
philosophical reflections, 2324 P
Mass dilation, 84, 90 Particle-pair, 74
Massenergy relation PCE. See Principle of conservation
discovery of, 13 of energy (PCE)
dynamic demonstration PCM. See Principle of conservation
Einstein, A., 6979 of impulse (PCM)
INDEX 105

Philosophical reflection S
on third demonstration of Second demonstration, of
massenergy relation, 5153 massenergy relation, 4246
of mass and energy in 19th Second-order effect, 90
century, 2324 Simultaneity, 2931
Post-Kantian empiricism, 23 Space diagram, 82
Potential energy, 22 Spacetime, Minkowski, 6569
Pre-Minkowskian period of Special relativity, 34, 2021, 40,
relativity, 57 50, 5455, 57, 61, 66, 69, 80,
Principle of Causality, 97 84, 97
Principle of conservation of energy historical review of, 2631
(PCE), 2021, 63, 70, 7273, as principle theory, 64
93, 96 Synchrony convention, 29
Principle of conservation of impulse
(PCM), 70, 7273, 92, 96
Principle of equivalence of mass T
and energy, 59 Taylor series expansions, 35
Principle of relativity, 15, 27, 42, Theorem of experience, 17
45, 53, 56, 57, 63, 64, 80 Theory of relativity, 45, 55, 57
Principle theories, 6164 Thermal energy, 22
Thermodynamics, 62
Q Third demonstration, of
Quantity of matter, 4, 5, 7, 1113, massenergy relation, 4651
16, 23 philosophical reflections on,
5153
R Thought-experiments analysis, 53
Radium, 41 Time dilation, 25, 39, 57, 81,
Relativistic kinetic energy, 7275 8385, 88, 90
Relativistic mass, 3, 8891, 89 Tolman, R. C., 8091
Relativistic momentum, 72, 91, Total energy
92, 96 of electron, 5152
Relativistic three momentum, of mass point, 58
7276
Relativity principle, 80
Rest-energy, 34, 5253, 5859, 69, W
70, 72, 73, 7778, 8687 Weight, 57
FORTHCOMING TITLES IN OUR ENERGY
PHYSICS AND ENGINEERING COLLECTION
Jonathan Swingler, Editor
Einsteins Mass-Energy Equation, Volume II: Quantum Mechanics and
Gravitation, Empirical Tests, and Philosophical Debates by Francisco
Fernores
Utilization of Variable Energy Sources by Wolf-GerritFrh
Electrical Power and Energy: Transformers, Synchronous Machines, and
Power Networks by Siong Lee Koh
Energy in Electrical Fields by Paul Weaver

Momentum Press is one of the leading book publishers in the eld of


engineering, mathematics, health, and applied sciences. Momentum Press
offers over 30 collections, including Aerospace, Biomedical, Civil,
Environmental, Nanomaterials, Geotechnical, and many others.

Momentum Press is actively seeking collection editors as well as authors.


For more information about becoming an MP author or collection editor,
please visit http://www.momentumpress.net/contact

Announcing Digital Content Crafted by Librarians


Momentum Press offers digital content as authoritative treatments of advanced
engineering topics by leaders in their eld. Hosted on ebrary, MP provides
practitioners, researchers, faculty, and students in engineering, science, and
industry with innovative electronic content in sensors and controls engineering,
advanced energy engineering, manufacturing, and materials science.

Momentum Press offers library-friendly terms:

perpetual access for a one-time fee


no subscriptions or access fees required
unlimited concurrent usage permitted
downloadable PDFs provided
free MARC records included
free trials

The Momentum Press digital library is very affordable, with no obligation to


buy in future years.

For more information, please visit www.momentumpress.net/library or to


set up a trial in the US, please contact mpsales@globalepress.com.

S-ar putea să vă placă și