You are on page 1of 291

THE DYNAMIC THEORY

A New View of Space-Time-Matter


by

Pharis E. Williams

Dedication
I dedicate this work to my family; to my father and uncle
who encouraged my thinking and individualism, to my
mother for her steady love, to my brothers and sister for
their confidence in my ability, to my children for growing
up with 'Dad's theory', and to my wife Jeri for she bore the
brunt of my mental absence.

Copyright

1993 by Pharis E. Williams

PREFACE
Present books, such as "The Arrow of Time" by Roger Highfield and
Peter Coveney and "The Big Bang Never Happened" by Eric Lerner, talk of a
new revolution is science. The first points to work by Ilya Prigogine and
others with regard to the flow of time and the dichotomy between the time
flow in the universe and physical theories wherein time may flow forward
and backward. The "Unended Quest" in " The Arrow of Time" is to find
how a foundation of science might be laid that describes dynamic systems
showing this one-way aspect in time. In "The Big Bang Never Happened"
Lerner also points out the need to find physical theories which correspond
to the directivity of nature's time. The main discussion though concerns
explanations of cosmological phenomena in terms of plasmas and
Maxwellian electromagnetic concepts.
I am in agreement with the authors of both these books with regard
to the majority of their points. I disagree with Highfield and Coveney in
that a foundation for physical theories restricted by a flow of time has been
found and reported starting in 1976. My disagreement with Lerner is very
limited, but may point out an important difference in our thinking. Let me
quote from Lerner's introduction where he states; "Today we again hear
renowned scientists, such as Stephen Hawking, claiming that a 'Theory of
Everything' is within our grasp, that they have almost arrived at a single
set of equations that will explain all the phenomena of nature --gravitation,
electricity and magnetism, radioactivity, and nuclear energy --from the
realm of the atoms to the realm of the galaxies and from the beginning of
the universe to the end of time. And once again they are wrong. For
quietly, without much fanfare, a new revolution is beginning which is likely
to overthrow many of the dominant ideas of today's science, while
incorporating what is valid into a new and wider synthesis." I believe
Lerner is correct. But only in the sense that I do not believe it possible to
know all of the phenomena of nature "from the beginning of the universe to
the end of time." What I put forth in this book is my research which shows
that one can start with a small, simple set of equations and derive the
basis for the currently accepted branches of physics by imposing restrictive
assumptions.
The search for a unifying field theory began in the early 1800's when
scientists began searching for a way of unifying the electromagnetic and
gravitation fields. When the proton-proton scattering results showed a
deviation from Coulombic scattering, once again scientists began trying to
find a way of unifying the fields, or forces, of nature. This was done
immediately upon the heels of assuming that the deviation from Coulombic
scattering must come, not from changes in Maxwellian electromagnetism,
but from an independent strong nuclear force. It has always appeared to
me that one should go back and address this assumption of independence
before seeking a means of unification.
i

One doesn't need to read too much of the scientific literature from
the 1930's to the present to see how much has been devoted to the notion
of unifying the forces, and/or fields, of nature. Within this body of work
lies the basis for Hawking's "Theory of Everything." I believe this work
misses the point of unification.
For instance, if we wish to approach a unification, what should we
unify? Should we unify the fields, or should we unify the various branches
of physics? It seems rather difficult to believe that nature is divided into
the different branches of physics, such as thermodynamics, Newtonian
mechanics, relativistic mechanics, and quantum mechanics, just because
we learned how to formulate the basis for each branch at different times in
our scientific advancement. Further, given a variational principle and a
metric we know how to derive field equations and force laws. Therefore,
shouldn't we be seeking to unify the various branches of physics and
deriving the necessary fields from that unification rather than trying to
unify the fields and not reconciling the difference between the foundations
of the different branches?
In my research I chose to seek a way of unifying the various
branches of physics. This entailed seeking a simple set of physical laws
from which one may derive the foundations of the different accepted
branches of physics as subsets of this more general set of laws. What has
emerged from this work is that there is a logical necessity for the branches
of physics that comes from the imposition of different restrictive
assumptions. The type of geometry need not be assumed as Newton and
Einstein did, but is dictated by the fundamental laws. The laws produce,
not one, but two variational principles from which we may derive the field
equations and force laws.
What resulted from the attempt to unify the branches of physics
produced not only the desired result, but, also that of unifying the fields
and forces of nature also. The fundamental laws, which could be written
on a T-shirt, produce field equations and force laws which accurately
describe phenomena intended to be included in Hawking's "Theory of
Everything." It does not, however, allow for the existence of a Big Bang or
beginning or end of time. Furthermore, since the fundamental laws are
based upon generalizations of classical thermodynamics, the equations of
motion derived from them come complete with an Arrow of Time built in. I
first reported this predicted flow of time in 1981.
If I were asked to explain why the research reported in this book has
not gained any wider distribution than it currently enjoys, I would have to
offer up our system of refereed journals as the most important reason. But
hand-in-hand with this must go the notion that "everyone knows that one
may derive classical thermodynamics from any number of different force
laws by using statistical mechanics." This notion was refuted by Peter G.
Bergmann in 1979, yet it persists today.
ii

On the other hand, if one were to accept the potential of having


equations of motion derived from generalizations of classical
thermodynamics, then it is not difficult to imagine an Arrow of Time
accompanying them. But this is small incentive to a referee. Neither is the
ability to derive the field equations and the force laws for the different
branches of physics much more incentive for the referee to give a thumbsup for such a theory which 'everyone' knows is doomed before it gets
started.
The many attempts to get portions, or all, of this research published
in the refereed journals have produced many interesting comments. These
comments are interesting from the point of view that they expose the
human side of referees, not that they are based upon scientific evaluation.
Let me offer three excerpts as examples: from the physics department of a
name university, "While the equations you've derived are not wrong, we
somehow like it better the old way," from a scientist at a government
laboratory, "If you ask me to shoot you down, I can't. If you ask me to help
you, I won't. I suggest that you learn to play the game and then someone
may listen to you," and from a journal dedicated to speculation, "We no
longer have the time to consider articles which look into the foundations of
physics."
What I sought to do was to answer some personal questions about
science using all of the rigor contained in the logic of mathematics. What I
found was a methodology by which we may see how the various physical
phenomena from the nuclear realm to the cosmos come from a single,
simple set of three fundamental assumptions.
Many current
interpretations concerning fundamental aspects of several existing theories
are shown to be wrong, misleading, or too restrictive. Notice that I said
many current interpretations are wrong, not many current theories are
wrong. What I found is that there is a much more general theory available
in which the current theories are subsets or first, or second, order
approximations. That doesn't mean these theories are wrong any more
than the validity of the Special Theory of Relativity means that Newton's
equations of motion are wrong. It only means that Newton's dynamics
applies only to a limited range of velocities . If we then use Newton's
equations of motion for velocities approaching the speed of light our
interpretations will of necessity be wrong. However, we didn't know these
interpretations were in error until Einstein put forth his more general
theory. The same is found to be true of many interpretations based upon
the current theories which the Dynamic Theory shows to be wrong when
viewed in its more general light. Also, the reported research shows how the
various branches of physics fit together into a unified picture of a nature
built upon the dimensions of space, time, and mass.

iii

TABLE OF CONTENTS
Preface

Chapter 1

Overview
1.1
1.2
1.3

Questions concerning the theoretical basis


Possible new theoretical approach
A new view of space, time, and matter

Chapter 2
New Theoretical Fundamentals
A. General Laws
2.1 First Law
2.2 Second Law
2.3 Absolute Velocity and Einstein's postulate
2.4 The concept of Entropy
2.5 Third Law
B. General Relations
2.6 Energy and Maxwell's Relations
2.7 Equilibrium conditions
2.8 Stability conditions
C. Geometry
2.9 Geometry required by fundamental laws
D. Mechanical systems near equilibrium
2.10 Special relativistic and classical mechanics
2.11 Energy concepts
2.12 Non-isolated systems
E. Quantum mechanics
2.13 Quantum Mechanics derived
2.14 On the derivation of thermodynamics from
statistical mechanics
F. Summary
2.15 Summary of new theoretical fundamentals

1
2
6
36

36
39
45
49
51
52
55
56
58
71
77
81
82
85
87

Chapter 3

90

Five-Dimensional Systems
A. Systems near an equilibrium state
3.1 Equations of motion
3.2 Energy equation
B. Systems with non-Euclidean manifold
3.3 General variational principle

91
95
97

3.4
3.5
3.6
3.7
3.8
3.9

Gauge function field equations


Energy-momentum tensor
100
Force density vector
Equation of energy flow
Momentum conservation
Gauge field pressure

Chapter 4
Five-Dimensional Quantization
A. Quantization in five dimensions
4.1 Quantization
4.2 Five-dimensional Hamiltonian
4.3 Five-dimensional Dirac equation
4.4 "Lorentz" covarience
4.5 Spin
4.6 Dirac equation with fields
4.7 Allowed fundamental spin states
B. Quantized fields
4.8 Quantum condition applied to particles
4.9 Radial field dependence
4.10 Self-energy of charged particles
4.11 Nuclear phenomena
4.12 Hiesenberg's Uncertainty Principle and geometry
4.13 Nuclear masses
Chapter 5
Five-Dimensional Gravitation
5.1 Charge-to-Mass ratio and magnetic moments
5.2 Perihelion advance
5.3 Redshifts
5.4 "Fifth" force
5.5 Inertial and Gravitational mass equivalence
5.6 Cosmology
Chapter 6
Electromagnetogravitic Waves
6.1 Wave equations
6.2 Wave solutions
6.3 Non-thermal transmission
6.4 Wave boundary conditions
6.5 Reflection and refraction
6.6 Complex refraction angles

99
104
106
106
108
110

110
111
112
113
114
115
116
119
121
127
132
138
144
157
157
167
171
183
192
194
198
198
198
205
208
217
224

6.7

Assumptions and wave solutions

Chapter 7

239

Hydrodynamic Systems
7.1 First fundamental quadratic form
7.2 Second fundamental quadratic form
7.3 Tensor derivatives
7.4 Relativistic hydrodynamics
7.5 Classical hydrodynamics
7.6 Shock waves
7.7 Mass conservative electrodynamics
Chapter 8

241
245
249
256
257
259
262
267

Experimental Tests
8.1 Speed-of-light
8.2 Index of refraction
8.3 Neutron interferometer
8.4 Nuclear masses
8.5 Gravitational rotor
8.6 Nuclear Lamb shift
Chapter 9
Epilogue
9.1
9.2
9.3
9.4
9.5
9.6
9.7
9.8

227

268
269
270
271
271
276
277

Only three basic assumptions


Geometry is specified
The Arrow of Time
Mass as a coordinate
Non-singular gauge potential
Unification of the branches of physics
The pedagogical aspect of the Dynamic Theory
Where to from here?

277
279
279
280
280
281
281
282

CHAPTER 1 OVERVIEW
1.1

Questions concerning the theoretical basis

It seems that throughout my working career I have been a troubleshooter. This started when I entered the Navy as an Electrician's Mate
working on the power electrical equipment on Navy ships. Troubleshooting
was the main job, whether it was finding some electrical malfunction or the
presence of saltwater in an electrical box. Later, as a Naval Officer with an
Electrical Engineering degree, I was constantly required to ferret out some sort
of trouble. This at times would involve missile systems, gun systems, radars,
sonar systems, boilers, or other systems. It seemed only natural then to employ
this same procedure to investigate what appeared to me as problems in the
foundations of physics.
Though I had often asked "Why?" when confronted with some new
assumption or adopted postulate, the first really puzzling facet of current
physics I encountered was the concept of relativistic kinetic energy from
Einstein's Special Theory of Relativity. The puzzling part was that it depended
upon the speed of light independent of the mechanism by which this energy
might be transferred. To better illustrate what puzzled me, consider the
transfer of energy between two charged particles on collision courses. If the
particles have near-miss trajectories, then the energy is primarily transferred
by the electrical forces between the charges. From the view of retarded
potentials, or the concept of a limiting speed of electromagnetic signal
transmission, it is rather easy to accept the energy transferred being dependent
upon this limiting velocity. But suppose the particles are uncharged and the
interaction is strictly a gravitational one. Again the concept of a limiting signal
speed would imply that the energy exchanged between the particles depend
upon this limiting velocity. But is it the same as the limiting signal velocity for
the electromagnetic case? Do gravitational waves travel at the same speed as
electromagnetic waves?
Einstein, in the Special Theory of Relativity, adopted the position that
the constancy of the speed of light forces a modification of Newton's dynamic
law. This modification implies that all forces have the same limiting velocity,
namely, the speed of light. There exists an abundance of theoretical and
experimental evidence that the speed of light becomes the limiting velocity
whenever electromagnetic forces are involved. The point that bothered me was
whether other forces, such as gravitational, should also have the same limiting
velocity. Though we have had reports of the detection of gravitational waves,
we have no experimental determination of the speed of a gravitational wave.
Therefore, I object to the viewpoint that the modification to Newton's law
should be applied to all forces without some additional justification.
Let me describe an analogy which may not hold in the strictest sense yet
may serve to illustrate my point of view. A river, flowing toward the sea,

carries energy with it. The speed with which this energy can move from one
point to another is the velocity of the river's current. The river produces a force
on a boat tied up to a pier on the river. When the boat is set adrift, this force
accelerates the boat. However, the maximum velocity to which the river can
accelerate the boat is the current velocity; this is the velocity with which the
energy of the river can propagate.
From this point of view the speed of light, being the propagation velocity
of electromagnetic energy must be the limiting velocity associated with
electromagnetic forces. Certainly nature would be much simpler if all forces
have the same limiting velocity. Yet without some experimental evidence of the
propagation of gravitational energy, I find it difficult to feel comfortable with
Einstein's modification of Newton's law justified by electromagnetic
experimental evidence and arguments of simplicity.
The fundamental philosophical viewpoint that the force depends upon
velocity and vanishes as the velocity approaches the limiting velocity raises
another question concerning Einstein's modification of classical mechanics.
Under Einstein's modification Hamilton's principle is written with a relativistic
mass which depends upon the velocity and a velocity independent force. Does
this represent a different philosophy or are both views equivalent?
More
specifically, are the "real" concepts to be taken as a mass independent of
velocity together with a velocity dependent force or should we associate the
velocity dependent relativistic mass and velocity independent forces with "real"
world? Or does it make any difference which we chose?
At this point I faced the first major decision. If I adopted Einstein's
postulates, then it appeared that I would be required to change my intuitive
beliefs concerning certain physical phenomena. I found this extremely difficult
to do. On the other hand, if I did not embrace these postulates, I would have to
replace them with something that would say essentially the same thing in all
cases where the Special Theory of Relativity has been found to be very accurate.
Not only this but if a new point of view were adopted, then virtually the entire
sphere of physics may need to be reviewed in order to ensure that the new point
of view did not conflict with currently used theories where they have
experimental verification.
1.2 Possible new theoretical approach
History records the advancements in physics which came from the efforts
of people new to the field. Therefore my lack of training in physics might be
turned into an advantage if I sought to determine a philosophical basis
unhampered by the directed philosophy that comes from a study of physics as
currently taught. This is in contradistinction with current practices and
procedures of academicism where mastery of current theories generally
precedes the development of a new one. To deliberately choose this deviation
risks accusations of arrogance and naivete. On the other hand such a choice

seemed the best way of avoiding the danger of becoming so familiar with
current ways of thinking as to make it improbable of giving due attention to
other ways.
Having decided to look for a new foundation for physics I was faced with
the question of how to begin. I recalled some Ozark hill philosophy I overheard
as a youngster. A native Ozarkian was giving directions to a stranger who was
trying to find a certain fishing hole. The directions went something like this:
"See yonder road going down that holler? Well, go down thar 'bout five mile
and you'll come to a fork in the road. Take the right hand fork. Now that's the
wrong one but you take it anyways. After you've gone a piece, you'll come to a
log across the road. Now you know you're on the wrong road. So go back and
take the left hand fork. You can't miss it."
A quick review of physics reveals that there are different branches with
different sets of fundamental laws or postulates. Though it is easy to see how
the distinction between these branches came about, it was difficult for me to
believe that nature shared the same divisions. I felt that all natural
phenomena should be explained by a single set of fundamental laws. This
belief is somewhat like a grove of redwood trees or bamboo forest. Above the
ground each tree appears as a distinct plant. Yet we know that below the
ground they may be found to grow from the same root system. Thus, I felt that
a more fundamental approach might display the unity in nature and that prior
attempts at unification in the search for a unified field theory could be likened
to attempts to tie the trees together at the tree top level rather than down at
the root level.
Is nature symmetrical in time? Does everything run backward in time
as well as forward? Obviously, not every process in nature will run backwards,
yet the equations of motion in Newtonian and relativistic mechanics are time
symmetrical. I believe in an asymmetrical nature and this belief played a role
in the eventual selection of fundamental laws.
How then did I use this philosophy to determine a set of generalized laws
on which to base an attempt to construct a new approach to physics?
Before proceeding let me offer a word of caution.
During any
theorization the philosophy of the theorist plays such an important role that an
attempt to understand the theory is aided by a knowledge of this philosophy.
Therefore the following includes not only the philosophical basis upon which
the theory is based and the mathematical development but also ideas and
beliefs which played a part in the various decisions.
Because of the
individualistic nature of philosophy the following will deviate occasionally from
a strict third person presentation, risking a loss of professional appearance, to
the clearly personal first person.
Newtonian mechanics fails to describe events involving high velocities,
relativistic mechanics fails to describe the atom, and gravitational effects have
resisted quantization. If these are viewed as logs and the Ozarkian's directions
are followed, then we must retrace our steps and seek another approach rather

than attempting to chop up the log and continue to push forward up one of
these roads.
The branch of thermodynamics, however, does not appear to have a log
somewhere along the way. Here the classical thermodynamic laws are very
general, particularly Caratheodory's statement of the second law. Thus the
thermodynamic laws appeared to be the fork in the road where a new route
might be chosen.
However, in mechanics we talk of equations of motion, field equations,
and geometry while in thermodynamics we speak of equations of state and
equilibrium. If a generalization of the classical thermodynamic laws is adopted,
how might we obtain the equations with which we are familiar in mechanics?
More particularly, how could this type of general law yield geometry and a
variational principle? The second law of thermodynamics can produce a
variational principle through principles such as increasing entropy and
minimizing free energy, but can it also produce a geometry?
This seemed to be a crucial point. If the laws could not produce a
geometry, then a geometry would have to be assumed, thus necessitating an
additional assumption. The belief that a simple fundamental set of laws should
lead to the fundamental principles of the different branches of physics made the
thought of additional assumptions abhorrent. The notion that the adopted laws
should specify the type of geometry that must be used seemed very satisfying.
Newton found that the absolute nature of Euclidean geometry brought
undesirable features. Einstein, in his General Theory, displayed the benefits
that might be gained by going to a more general geometry. He showed that
physical phenomena might be displayed as elements determined by certain
physical laws. This is essentially the question here. Can a set of laws, which
are generalizations of the classical thermodynamic laws, determine the metric
elements and hence the geometry?
By appealing to the mathematics of functions of more than one variable
we find that a quadratic form becomes involved when a maximum or minimum
is sought. Further, this quadratic form generates a natural geometry for that
function.
In thermodynamics the stability conditions provide a similar
quadratic form and therefore the quadratic form which specifies the stability
conditions should form a natural geometry for a physical system governed by
laws such as the thermodynamic laws.
Thus the foundations of the theory have been outlined, namely the belief
that all physical phenomena should be derivable from a single set of physical
laws which are generalizations of the classical thermodynamic laws. Such a
theory should be capable of describing all the dynamic events in nature.
Therefore it seems appropriate to call it the "Dynamic Theory". Obviously, for
such a theory to be tenable it must reproduce, or be consistent with, the various
fundamental postulates and/or laws currently used in the various branches of
physics. Indeed it should do even more. It should also reduce the number of
necessary assumptions and provide an unprecedented unification of physics.

Further, there is the possibility that the theory might produce an


experimentally verifiable prediction.
The first requirement that should be placed upon the Dynamic Theory is
that it reproduce, or be consistent with, current theories. In order to show that
the Dynamic Theory satisfies this requirement, Section A of Chapter 2 states
the adopted laws and then sections of the remainder of the book show how
appropriate restrictions upon the system do yield the fundamental principles
for the various current theories.
Though a theory which has the capability of displaying a unification of
physical theories might have significant value based solely upon this
capability, it would become more attractive if it could explain phenomena for
which no explanation exists or make some new prediction which might lead to
an experimental test of the theory. Since restrictions were placed upon the
system in order to show how current theories may be obtained, the easiest way
to see the expanded coverage of the theory is to relax one or more of the
restrictions and consider a more general system. In Chapters 3, 4, 5, 6, and 7
some of the previously imposed restrictions are relaxed and the results are
worked out for several types of systems. Chapter 8 presents some experiments
which might test the Dynamic Theory.
A theory, such as the Dynamic Theory, immediately poses several
problems which are not associated with its validity or applicability. First, there
is a new point of view to be dealt with. Initially it would appear to be
inconsistent with all past concepts of system energy or relativistic concepts.
Yet in the end it is completely consistent with current theories and sheds an
entirely new light upon physical phenomena.
Another imposing difficulty with the Dynamic Theory stems from its
generality. The scope of the theory includes all physical phenomena while in
the past half century the vast amount of scientific knowledge that has been
accumulated has demanded specialists. Increasing expansion of mankind's
knowledge demands further specialization. Such a progression produces no
demand for a generalist. The result is that the greater portion of this theory
will be outside the field of many readers.
Closely associated with this problem is another. Throughout science
symbols and words are used to denote concepts and quantities. The limited
number of available symbols and words together with the expanded scope of
scientific knowledge requires duplication. For the specialists this duplication
can be somewhat minimized. However, in the case of a general theory touching
virtually all areas of specialization the problem becomes very significant. In
particular, if a certain symbol or set of words is used, a particular notion or
concept may be associated with them by the reader. This association will likely
depend upon the reader's specialty and therefore will vary with the reader.
Any attempt to choose symbology or word usage aimed at a particular specialty
risks increased confusion for readers in other fields. Therefore, the reader is
cautioned to keep in mind that conceptualizations and symbology familiar

because of its use in one branch of physics may now take on an entirely new
meaning.
1.3

A New View of Space-Time-Matter

The history of mankind's attempts to unify electromagnetic and


gravitational fields, or interactions, began when man begin to learn of electric
and magnetic fields. The first formal theory attempting to unify the two fields
of science was presented in 18361. However, no theory has yet been suggested
that has gained undeniable experimental verification. Theoretical physicists
are still at work trying to find a theory that will ultimately unify the forces of
nature. Such is the strength of the belief in the unity of nature.
The theory developed below adopts the premise that a description of
physical phenomena should be based upon a simple set of fundamental
postulates and that the current physical theories should be found to be subsets
of this more general theory by applying restrictive assumptions. The selection
of the three following fundamental laws reflect this premise.
Generalized Laws
In looking for a choice of fundamental basis for a theory to unify the
various branches of science, consider the following. Newtonian mechanics fails
to describe events involving high velocities, relativistic mechanics fails to
describe the atom, and gravitational effects have resisted quantization. On the
other hand, one finds that thermodynamics is the one branch of science which
has always been found to hold. Here one finds the classical thermodynamic
laws to be very general, particularly Caratheodory's statement of the Second
Law.
In mechanics the basic equations discussed are equations of motion, field
equations, and geometry, while in thermodynamics the basic equations are
equations of state and equilibrium. If a generalization of the classical
thermodynamic laws are adopted as a fundamental basis for a unifying theory,
how may the familiar equations from mechanics be obtained? The crucial point
is how to obtain geometry and a variational principle from these general laws.
Given a geometrical description and a variational principle, established
procedures may be used to obtain equations of motion and field equations.
Geometry may be obtained from a quadratic form. Therefore, stability
conditions should yield a natural geometry based upon laws generalized from
the classical thermodynamic laws. Further, in thermodynamics we find two
variational principles; one in the maximum entropy principle for isolated
systems, the other is the minimum free energy principle for non-isolated
systems.

First Law
The First Law is taken as the statement equating the energy exchanged
between the system and its surroundings to the change in the system energy
plus any work that the system does. The form for expressing this law is
j

_E = dU - f j d q ; (j = 1,...,n).

1
In Eqn. (1), dU represents the differential change in the system's energy, E
represents any and all energy exchanged between the system and its
surroundings that cannot be expressed by a work term.
There is no restriction in this law concerning the number of independent
variables. The dimensionality depends only upon the applicable, independent
work terms. However, in this presentation it is beneficial to initially place
some restrictions upon the type and number of allowed work terms. Therefore,
a system with only one work term which is the pdv expansion work of
thermodynamics will be called a "thermodynamic" system. A system with three
mechanical fdx work terms will be called a "mechanical" system.
An important aspect of this law is that, while the energy of the system is
a function that is independent of the path, both the energy exchanged with the
surroundings and the work done depend upon the path by which the system
goes from one state to another. The path dependence of these terms places
severe limitations upon the utility of this law and will become important when
viewing relativistic and Newtonian mechanics using the new theory.
Second Law
Caratheodory's statement of the Second Law of Thermodynamics is very
abstract and does not depend upon the type or number of variables used and,
therefore, is already in very general form. The law simply says that there exist
states to which the system may not go and then be able to return to its original
state. Though Caratheodory formed this statement in terms of neighborhoods,
it is known from thermodynamics that it contains the aspects of prohibiting
perpetual motion; to be exact, perpetual motion of the second kind. The point is
that this law seems intuitively to apply to mechanical systems as well as
thermodynamic systems.
The Second Law is stated as:
In the neighborhood (however close) of any state of a system of
any number of independent variables, there exist states that
cannot be reached by reversible E-conservative (E=0) processes.

Obviously, if attention is restricted to purely thermodynamic systems


with only a pdv work term, these laws produce classical thermodynamics.
Therefore, the important question is whether or not the laws contain the
existing mechanical theories when only mechanical fdx work terms are
considered.
In thermodynamics, Caratheodory used his statement of the Second Law
to show that the Second Law guarantees the existence of an integrating factor
for the First Law. One important feature of such a result is that the
integrating factor converts the path dependent First Law into a path
independent statement. Two other features resulting from Caratheodory's
work have increased significance when applied to a mechanical system.
Caratheodory showed, in classical thermodynamics, that the integrating factor
is a function of temperature only and that it is independent of the system.
When a mechanical system is considered, the integrating factor can be
2
shown not only to exist but also to be a function of the velocity only and
independent of the type of force considered. Since the integrating factor is
strictly a function of velocity, an absolute velocity may be defined as in
thermodynamics where an absolute zero temperature is defined. Thus, the
absolute velocity is defined as that constant velocity at which a system may
undergo a process from one solution curve to another without exchanging
energy with its surroundings.
Mechanical Entropy
The integrating factor may be used to define a mechanical entropy just
as we do for a thermodynamic system. Here the definition becomes
dS =

_E
,
(q&)

where S is the mechanical entropy and the process is a reversible one. Thus,
the path independent function obtained by using the mechanical integrating
factor is the function defined as the mechanical entropy.
The Second Law may be used, as done in Section 2.4, to show that an
isolated mechanical system, which cannot exchange energy with its
surroundings, undergoing a spontaneous, or irreversible, process must
experience an increase in its mechanical entropy.
Third Law
Just as in thermodynamics, where a Third Law was needed in order to
associate the entropy of one system to the entropy of another, so also a Third
Law is needed here. The Third Law may be stated:

The entropy of a system, when the integrating factor becomes


infinite, is a constant, and this constant may be taken to be zero.
Some of the immediate results of these adopted laws may now be
presented. In particular, the definition of the absolute velocity says the
integrating factor goes to zero for this unique velocity. The Third Law
combines with the Second to say that the absolute velocity may not be obtained
in a finite number of steps. Thus, the absolute velocity becomes a unique
limiting velocity. Also, the Second Law showed that the integrating factor was
independent of the type of force considered. Therefore, the limiting velocity
does not depend upon the force and, hence, must be the same regardless of the
type of force. Thus, not only must all forces have the same limiting velocity, but
since the absolute velocity is unique and the only velocity found in Nature that
exhibits this characteristic is the speed of light, then the speed of light must be
the absolute velocity. Further, since the definition of the absolute velocity is
made for a constant-velocity process, Einstein's assumption concerning the
constancy of the speed of light comes directly from the adopted laws (see
Section 2.3).
Geometry
In order to find the equations of motion for a mechanical system, the
geometry required by the adopted laws must first be determined. Since the
mechanical system was considered to have three fdx work terms, the energy of
the system becomes a function of four independent variables: three space
variables and the mechanical entropy. Thus, the quadratic form obtained from
the stability conditions may be expressed in terms of the variables of space and
mechanical entropy and is
2U
2 U
2 U
2

(dS
+
2
(dS)(
)
+
( dq )( dq ) > 0 ;
)
dq
S2
S q
q q
( , = 1,2,3).

(2)
Adopting this quadratic form as the metric of a general system whose
thermodynamic variables are held fixed, the metric may be written as
(ds )2 = hij dqi dq j ; (i, j = 0,1,2,3)

(3)
where the summation convention is used and
hij =

2
U
qi q j

where q0 is the entropy. Thus, the stability conditions provide a metric in the
four-dimensional manifold of space-entropy.
The arc length s, in the space-entropy manifold, may be parameterized
by choosing
ds = q& 0 dt = cdt,

where c is the unique velocity appearing in the integrating factor of the Second
Law. The metric may now be written as
2
i
j
c 2 (dt ) = hij dq dq ; (i, j = 0,1,2,3).

(4)
But Einstein's relativistic theories are in space-time manifolds. In order to
show that the proposed theory contains Einstein's theories, a space-time
manifold must come from the adopted laws. This is indeed the case if the
mechanical system is restricted by requiring that it be isolated (E=0). This
restriction establishes the condition necessary for the principle of increasing
mechanical entropy which becomes a variational principle
(d S )2 = ( dq0 )2 = 0.

(5)
In order to use this variational principle, Eqn. (4) may be expanded, solved for
dq0 and squared to arrive at the quadratic form
1
( dq0 )2 = [ c 2 (dt )2 + 2 h0 Adtdq - h dq dq ],
h00

(6)

where

A=

h0 q&
+_
h00


2
c 2 h q& q&
h ( q& )
+ 0
h00
h00
h00

with q& n = dq n /dt. 11


By defining x0 = ct and xn = qn then Eqn. (6) may be written as

10

( dq0 )2 =

1
g dxi dx j ; (i, j = 0,1,2,3)
f ij

(7)
where f = h00. This metric obviously reduces, in the Euclidean limit of constant
coefficients, to the metric of Minkowski's space-time manifold of Special
Relativity. Thus, the stability conditions and the principle of increasing
entropy combine to require that the equations of motion for an isolated system
be the equations of geodesics in a space-time manifold. However, this manifold,
whose arc length is the entropy, is related to another space-time manifold by a
gauge function so that a discussion of geometry involves two space-time
manifolds. Recalling Eqn.(7) we have
( dq0 )2 =

1
1
2
g ij dxi dx j = (d ) = g ij dxi dx j .
f
f

(8)
The path independence of the entropy fully specifies the geometry of both
manifolds3.
For the entropy manifold, the geometry is required to be
Riemannian with a vector curvature. The other manifold, which may be called
the "sigma" manifold, is required to have a Weyl geometry with both a vector
curvature and a distance curvature. The distance curvature refers to the
changes of the length of a vector under parallel displacements in the sigma
manifold and is found to be given by
1 log f
dl =
l = ( i dxi )l.
2 xi

(9)
The requirement of two manifolds for an isolated system and the fact
that the adopted laws fully determine the geometry of each are two of the most
significant aspects of the proposed theory. The requirement that there be two
manifolds coupled by a gauge function gives rise ultimately to Maxwell's
electromagnetic theory as well as quantization. The fact that the laws specify
the geometry removes the necessity of assuming a particular geometry and
leads to the removal of objections to Weyl's unified field theory of 19184 and to
London's quantization of Weyl's work in 19275. It is these aspects of the theory
which allow the unification of the different branches of physics.
In 1918, the German mathematician Weyl proposed a unified field
theory based upon his extension of geometry. However, this theory has not
gained acceptance, partly because his theory produced only Einstein's General
Theory of Relativity and Maxwell's Electromagnetism. Weyl's theory said
nothing in addition to these theories. Another reason Weyl's theory failed to

11

gain acceptance is that Einstein produced an argument that, using Weyl's


theory, the spectral lines produced from an atom must be dependent upon the
history of the atom, which contradicts experience.
The proposed theory removes both these objections. The first objection is
quickly removed by the fact that, from the point of view of this theory, the
system has been restricted to be, first, a strictly mechanical system, and
secondly, an isolated system. The removal of either of these restrictions allows
the theory to discuss events that cannot be addressed by Einstein's General
Theory of Relativity or Maxwell's Electromagnetism as will later be shown.
The second objection is removed by the fact that the Second Law requires that
the entropy and, in addition, the change in entropy to be independent of the
path and hence, Einstein's argument of path dependence is nullified. Later it
will be shown how the theory arrives at the predictions of phenomena predicted
by Einstein's General Theory and Maxwell's Electromagnetic Theory from a
different approach in which the geometry is required to be that of Weyl.
SPACE-TIME-MASS
In 1918 Weyl published his book titled "Space-Time-Matter" in which he
discussed Einstein's theories and his own unified theory of gravitation and
electromagnetism. Yet, within the text, matter did not share the same role as
space and time, though an equivalence was implied by the title. Space and
time were coordinates in the relativistic manifold while matter was not.
The theory proposed here does for mass what Einstein's Special Theory
of Relativity did for time; mass also becomes a coordinate on an equal footing
with space and time. Just as relativity created considerable conceptual
difficulty, so also can the proposed theory be expected to create conceptual
difficulty. However, if the unification provided by the theory is considered as
justification for attempting to see what the theory might produce, then a fifth
dimension with physical interpretation follows rather quickly.
The first restriction placed upon any system discussed thus far has been
that of restricting the system to be either a thermodynamic system, with only a
pdv work term, or a mechanical system with three fdx work terms. By
removing this restriction, a system must be considered, which may be called a
thermo-mechanical system, that experiences four types of work. Thus, the first
law includes four work terms and, therefore, involves five dimensions. Since the
specific volume is the reciprocal of the mass density, the First Law may also be
written in terms of the mass density. For such a system the coordinates
become the specific entropy, mass density, and the three space variables.
If five dimensions, which include the mechanical and thermodynamic
variables, seem odd consider how thermodynamics is taught. The First Law of
Thermodynamics is written on the blackboard, equating the differential heat
exchanged between the system and its surroundings with the differential
change in the internal energy plus the differential work terms. In the work

12

terms are the three mechanical work terms in addition to the thermodynamic
work. It is then pointed out to the students that the right hand side of the
equation has five independent variables and it is stated that five equations are
needed which relate these variables in order to have a solvable system. Usually
the first statement made at this point is that conservation of mass guarantees
that the mass density may be written as a function of space and time and,
therefore, only four additional equations are needed, which are stated as being
the Equation of State and the three mechanical equations of motion, such as
Newton's.
But what about the case when mass isn't conserved? Can mass density
be written as a function of space and time for this case also? If it may not then
the fundamental dimensionality of nature must be five dimensions. Where
does this lead? It has already been shown that the stability conditions lead to
metrics upon which the Entropy Principle works to provide equations of motion
when the metric coefficients are assumed known and field equations for these
coefficients when they are not known. Thus, it is necessary only to work out
what the implications of the five fundamental dimensions would be, compare
them to the existing theories in those regions of physical phenomena where the
existing theories are known to work and see if there is some predictable critical
experiment that may be conducted to test the new theory.
To begin the investigation of the implications of this five-dimensional
system first consider the system to be isolated. The principle of increasing
entropy becomes effective and the equations of motion are the equations of
geodesics in a five-dimensional manifold of space, time, and mass density.
The First Law for five dimensions may be written as
~
~ P
_E = dU - 2 d - F dq , ( = 1,2,3)

(10)
where the tilde denotes specific quantities. The entropy variational principle,
as stated in Eqn. (5), becomes
(dS )2 = ( dq0 )2 = ( dq0 )2 = 0

(11)
where now q0 is the specific entropy.
The system's specific energy is now given in terms of the five variables
specific entropy, space, and mass density. The stability condition, and hence
the metric, is then stated in terms of these same variables. The stability
condition is stated as

13

hii dq dq =

2 ~
U
i
i
dq dq > 0 ; (i = 0,1,2,3,4)
qi qi

(12)
where q4 = /a0. The metric may then be written as Eqn. (3) with the indices
running from 0 through 4. Eqn.s (7) and (8) give the five-dimensional geometry
when the indices also take on the value 4.
Equations of Motion
The equations of motion are obtained when it is assumed that the
coefficients of the metric are given and one looks at the Euler equations giving
the variations of the coordinates which satisfy the variational condition. By
using the variational principle from Eqn. (11) one finds the force densities to be
given by
i
i
F = f

(13)
with
i
i
du i
f = 0 + u l u k
dq lk

where ui are the components of the five-dimensional velocity vector, the


i

lk

are the Christoffel symbols, and fi are the components of the five-dimensional
acceleration vector. Obviously, if the mass density is considered to be conserved
such that u4 = 0 and the system is near equilibrium so that a flat metric makes
a good approximation, then the volume integral of Eqn. (13) becomes the
force-mass-acceleration relation of Special Relativity. Therefore, Einstein's
Special Theory is obtained within this theory by employing the restrictions of
an isolated system near equilibrium, with conservation of mass.
It is interesting to note that the inertial mass density comes from the
fact that the stability conditions are given in terms of specific quantities while
the Entropy Principle is stated in terms of the entropy. This fact will take on
an even more interesting character when we consider the comparison between
inertial and gravitating mass.
Another interesting fact is that if the First Law is considered for an
isolated system, one obtains

14

~ p
_E = 0 = dU - 2 d - F~ dx ; ( = 1,2,3)

so that
~
dU = F~ dx

( = 1,2,3,4)

(14)
When the energy integral of Eqn. (14) is evaluated one finds
~
1
1
U = c2 + v 2 +
( & )2
2
2 ( ao )2

(15)
where u4 = /a0 is used and it is assumed that the system is near rest. The
energy density in Eqn. (15) includes the rest energy term because the integral
requires it; not because of a constant of integration as in the Special Theory of
Relativity. Further, because the system was considered to be isolated, pE=0,
then the appearance of the rest energy term in the expression for the system
specific energy brings with it some sublities of interpretation not found in
Einstein's Special Theory where energy and mass are equated one-for-one. For
instance, the one-to-one correspondence between energy and mass exists only
for resting mass when mass is conserved. Also notice that the Special Theory of
Relativity energy equivalence may exist only for isolated systems. Also, if we
require the usual conservation of mass then d/dt=0 and Eqn. (15) reduces to
the rest energy plus the classical kinetic energy.
Gauge Fields
When the standard variational techniques are used on the metric for the
isolated, five-dimensional system, it is found6 that the gauge function yields a
gauge field with ten components as
iE1
iE 2
iE 3 iV4
0

0
B3 B2 V1
iE1
Fij = iE 2 B3
0
B1
V2
(16)

0
V3
iE 3 B2 B1
iV
V1 V2 V3
0

4
and eight partial differential equations, Eqn. (17),

15

_ B = 0
1 B
+ x E = 0
c t
1 E
V 4 J
x B + a0
=

c t
c

_ E + a0 V 4 = 4

+ _ J + a0 J 4 = 0
t

xV + a0

B
=0

1 V
E
= a0
c t

1 V 4
4
_V +
=J4
c t
c
V 4 +

(17)
which replace Maxwell's four equations and the equation of charge continuity.
However, there are four new field components appearing in these eight
field equations. When these are assumed to be zero the system of equations
collapses back to the Maxwell's equations of electromagnetism. It is no surprise
that the collapse of the eight equations produces the Maxwell system; this has
been shown by many researchers. The objective becomes one of how are these
new field components to be interpreted?
Initial investigations led into the five-dimensional quantum mechanics
and to a predicted magnetic moment for neutrally charged particles7 (discussed
later). Current theories ascribe these anomalous magnetic moments to the
strong nuclear force. This led to the erroneous interpretation that these new
field components must be related to the nuclear forces. This turned out to be
wrong when later research was conducted in which a closer look was taken at
the concept of fundamental particles.
Fundamental Particle Fields
The concept of fundamental particles might be rather loosely stated as
something like "smallest possible" or "cannot be further divided". But one
generations' fundamental particles have been divided by the next generation
until there now exists a plethora of "fundamental" particles and the search for
more continues. But how can the concept of "fundamental particle" be stated
with mathematical rigor? If a mathematical statement for this "state" can be
put forth, then the logic of mathematics may be used upon the field equations
and it should then be possible to determine what fields these "particles" or
"states" might have.

16

Consider the concept of a fundamental particle and look for a


mathematical definition for it. First, consider the realm of thermodynamics
where the very stable states are isentropic states and, therefore, suppose that
the fundamental particles are isentropic states. When one looks at the metric
for an isentropic state of an isolated system one finds that the condition which
the German physicist London imposed upon Weyl's theory in 1927 is required.
Namely, one finds that in order to satisfy the isentropic condition the line
integral formed by the gauge potentials and the differentials of the metric
variables must be quantized, or since E=0, then, from Eqn. (8), (d)=(d)0 so
that
e

j dxi

=1

(18)
which is satisfied only if
j dx j = 2iN

(19)
where N is an integer and i is the square root of minus one.
When a line integral is encountered in the class room the students are
generally asked to find the value of the line integral given a certain path. Here
though, one has a line integral that already has a value. There are then two
questions that might be asked. First, if the gauge potentials are given, what
are the paths allowed? London's work answered this question5. The only paths
possible are those given by the solutions to the quantum mechanical equations
of motion. Further consequences of this result will be discussed later. The
second question that might be asked of the line integral is; what gauge
potentials are allowed by the line integral if the value of the integral is
independent of the path? This is asking what potentials may be used in the
integral which will produce a quantized value for the integral independent of
the path considered? This is the same as asking "What fields may a particle
have if these fields are to be independent of the path?"
If the value of the integral is to be independent of the path, then Eqn.
(19) must be true even when all dxj are zero but one. Thus, the quantum
condition requires that
k dxk = 2iN,

(20)
where there is no summation on k. Eqn. (20) must be true for all k, and
because one is free to choose the path, the k must reflect the quantization
represented by the integer N. Therefore,

17

~
j = N j j

(21)
where there is no sum on j and the may not be quantized. Thus, Eqn. (21)
represents the first response to the question concerning what j are allowed for
fundamental particles; the gauge potentials must be quantized.
This is the first known quantization of the gauge potentials for particles
which is required by some fundamental condition, such as the isentropic state
requirement. Restating; this is the first display of a logical necessity for
quantization of electric charge based upon fundamental principles and obtained
by restrictive assumptions.
By using the mathematical approach of assuming a solution in the form
of a product of functions of independent variables and setting
1

log f 2 = f t f r f f f ,

the trial solution was run through the eight field equations of Eqn. (17)8. The
result produced for the radial function is
fr =

k -
e r.
r

(22)
Here depends upon the particle and the potential displays some familiar
attributes of the Maxwellian gauge potential and some that are, at first,
surprising.
The potential corresponding to the classical electromagnetic
potential
r =

Zk -
er
r

(23)
where Z is the quantum number required by the quantum condition, depends
only upon the radial distance from the particle, not just the usual 1/r
dependence. At first glance one is prompted to state that this is the Yukawa
potential. However, the exponent in the Yukawa potential goes as r rather
than 1/r. One may also note that this potential has no singularities for any
value of the radial distance r. At distances much greater than this potential
(herein called the Neo-Coulombic potential) has the familiar 1/r form from
electrostatics and Newtonian gravitation. When the radial distance equals
lambda the potential has its maximum absolute value. Because of the
overriding effect of the exponential the potential returns to zero as r tends to
zero. The Neo-Coulombic potential is so well behaved that all of its derivatives

18

are also non-singular. This property will prove to be of extreme value when
considering such a potential in quantum mechanic systems since no
renormalization is required. Therefore, the usual problems arising with
renormalization do not appear with this potential.
The Neo-Coulombic potential gives the electric field radial component a
long range 1/r2 dependence that we know for the electric field,
Er =

Zk -
1 - e r .
r2 r

(24)
It also requires that the electric field rise to a maximum absolute value as r
decreases from infinity, go to zero as r approaches lambda, reverse sign as r
becomes smaller than lambda, go to another maximum absolute value and then
approach zero as r tends to zero. This short range behavior is drastically
different from that of the usual electrostatic field and will have enormous
consequences for the nuclear phenomena wherein the radial separations are of
the order of the lambdas of the fundamental particles.
The next thing noticed about the gauge potentials arrived at by the
above method is that the new three dimensional vector field has two
multiplicative factors, for

1

V r = W (1 + bt ) 2 1 e r
r
r

(25)
The first factor has the same radial dependence of the electric field and hence
the long range 1/r2 dependence. If this is to represent a physical field other
than the electric field then it must be the gravitational field. To further confuse
the issue, the second multiplicative factor involves a dependence upon time. At
first this may seem to run counter to all knowledge of gravitational effects;
however, later it shall be shown that this time dependence is all important in
gravitational phenomena.
Is it possible then that the ten gauge field components may be made up
of the three electric field components, three magnetic field components, three
gravitational field components, and the gravitational potential? Only by
working through the predictions of the theory in the various areas of physical
phenomena can it be determined whether the predictions can be supported by
the experimental evidence or if the predictions run counter to the evidence. If
there exists experimental evidence that is in measurable direct conflict with the
predictions of the theory then the theory must be wrong. On the other hand, if
the predictions are supported by the evidence and predictions exist which may

19

be used to test the validity of the theory then the theory deserves more than a
offhand dismissal just because it disagrees with existing theories or beliefs.
Quantization Derived
The strength of the quantum-theoretical structure is such that it has
swept aside virtually every attack upon it. However, using classical definitions
of commutivity it may be shown9 that the anti-commutivity of the position and
momentum is dependent upon the metric approximating a flat metric. If a
realm of conditions exists that does not allow a flat metric approximation then
the commutators must depend upon the geometry. One finds that

j
[ x j , p k ] = ih g kl jl + x s
sl

where the
j


sl

are the Christoffel symbols. This much does not depend upon any theory
whatsoever, but only upon the mathematics of differentiation. Since the
quantum Poisson brackets must correspond to the classical Poisson brackets,
then they also depend upon the geometry in the same fashion. In the past it
has been possible to argue that if the only physical field that affects the
geometry is the Einsteinian gravitational field, then it is possible to ignore this
geometrical effect upon the commutivity of space and momentum in nuclear
phenomena. If, however, the gravitational field is described by a gauge field
then this argument is nullified because the gauge fields do play a large role in
the realm of nuclear physics.
The German physicist London produce a quantization of Weyl's theory in
1927. In his work, London showed that if the arc length of the metric was
required to return to its original value, a quantization was produced and that
the wave function was proportional to this arc length. However, there was a
difficulty with his work; it required an imaginary distance.
The proposed theory not only removes the difficulty of the imaginary
distance but further, logically produces the quantization conditions when the
system is placed under an additional restriction. The quantum condition, as
stated before, comes from restricting one's attention to systems which are
isentropic. The requirement that the system have a constant entropy is the
simplest restriction that produces London's quantization. The imaginary
distance appearing in London's work also appears here in the entropy
manifold. However, the attractive electromagnetic force comes from a negative
gauge function which couples the "distance" in the manifold with the Weyl
geometry to the entropy manifold. In the entropy manifold the change in

20

entropy is the distance and, therefore, distance must always be real and
non-negative for an isolated system because of the principle of increasing
entropy.
The proposed theory then logically produces London's assumption and
removes the difficulty with imaginary distances. Further, it is found that the
quantization conditions are limited to a system with a distance curvature, or
gauge function. Thus, the interpretation of universal application of a nonvarying, least unit of action coming from Heisenberg's Uncertainty Principle
rests with the existence, or lack, of a distance curvature and not with the
existence of a vector curvature. Equivalently, only forces that may be
expressed in terms of a gauge function, or distance curvature, may exhibit
quantization, while forces describable by only a vector curvature cannot be
quantized. If the above interpretation of the new field components as
gravitational field components holds up as gauge field components then
gravitational effects may be quantized as well as the electromagnetic effects.
This description of the derivation of quantum mechanics from
generalizations of the classical thermodynamics runs counter to the commonly
held belief that one may derive classical thermodynamics using statistical
methods and a variety of force laws. This contention is, however, without
rigorous support, as may be seen when one considers the development of
statistical thermodynamics. For instance, in order to talk of a statistical
temperature one must start by assuming Newtonian physics (this constitutes
three fundamental assumptions). Given Newtonian, or other physics, one can
talk of an energy distribution, canonical ensembles and statistical temperature;
however, one must make an additional fundamental assumption (the
Equipartition Law) before the statistical heat capacities may be obtained.
In order to obtain thermodynamics two more assumptions are required.
It was pointed out by Peter G. Bergmann10 that using the statistical approach
one may obtain an expression for the difference in the heat exchanged between
the system and the surroundings and the element of work done. In classical
thermodynamics this difference is the change in internal energy which is path
independent. In the statistical approach the difference is obtained without
reference to the internal energy. To claim that the statistically derived
expression is an exact differential is a logically new assertion; it constitutes the
First Law of Thermodynamics. In addition, the assumptions of statistical
thermodynamics allow the derivation of the fact that the differential of heat
exchanged must be greater than, or equal to, the multiplication of the
statistical temperature by the differential change in the statistical entropy.
This product of statistical thermodynamic properties is similar to an identical
product of thermodynamic properties. In statistical thermodynamics it is
asserted that the ratio of the statistical temperature and the classical
temperature is Boltzman's constant.
Once this assertion is made, the
statistical entropy may be related to the classical entropy. However, there is no
logical necessity that the ratio of temperatures be a constant from the

21

statistical approach; only if it is a constant can there be a one-to-one


correspondence between the statistical entropy and the classical entropy.
The above quantum condition establishes the conditions assumed by
London and, therefore, one may follow his work in deriving the Schrodinger
quantum mechanics. London's work establishes how quantum mechanics may
be derived within the framework of a larger theory and will not be repeated
here. Rather, a sketch of the five-dimensional quantum mechanics will be
presented11.
The variational principle required by the entropy principle is given by
Eqn. (11). Because multiplication by a constant does not change the problem,
one may write
c 2 ( dq0 )2 = 0.

(26)
By defining the velocity vector as uj = dxj/dq0 and the momentum as pj = cgjkuk,
where the fact that gjkujuk = 1 has been used, one may show that pjpj = 2c2,
which is the five-dimensional "momentum-energy" relation.
Because of the benefits of a first-order differential wave equation, Dirac
sought to find a first-order operator equation that also satisfied the second
0-order Klein-Gordon equation (the operator equivalent to the
momentum-energy relation). This can also be done in five dimensions by
taking the specific Hamiltonian operator to be

h = i 1 1 + 2 2 + 3 3 + 4 4 - .
x
x
x
x

(27)
By taking the four partial derivatives in Eqn. (27) as the components of the
four-vector specific momentum operator, one may write
h = - ( _ P + )

(28)
where natural units, h = c = 1, have been used.
If one takes the p0|> = h|> and requires that the alphas and beta are to
be chosen such that solutions of this equation are also solutions of Eqn. (28),
one finds the restrictions imposed upon the choice of the alphas and beta to be

22

( _ P )2 = P 2 ,
2 = 1,
+ = 0.

(29)
The set of 4 x 4 matrices satisfying the requirements of Eqn. (29) is given as
I 0
=
,
0 I
j 0
; j = 1,2,3, ,
j =
0

2
4 =
0

2
(30)

where I is the 2 by 2 identity matrix and the sigmas are the 2 by 2 Pauli spin
matrices.
Then the five-dimensional Dirac equation may be taken to be
i

(t) = i( _ - ) (x)
t

(31)
where we have used the four-dimensional vector operator. By defining
0 = ; j = j ;

j = 1,2,3,4,

then Eqn. (31) may be written as


(i j j + 1) (x) = 0.

(32)
Taking into consideration Eqn. (32) with the gauge fields of Eqn. (16),
one arrives at
1

k
k
jk
(i j - j )(i - ) - 1 - 2 i F jk = 0

(33)

23

where

jk

0
x& 1
x& 2
x& 3
x& 4

x& 1

x&

0
2 is 3
21 s
2 iu

2
1

x&

2 is
0
2 is 1

2 iu

x&
2

2 is
2 is
0
2 iu

4
1

2 iu
2 iu 2
21 u 3

(34)

and s is the usual intrinsic spin while u is a new spin appearing because of the
added dimension. By expanding, one finds that Eqn. (33) becomes
[(i j - j )(i k - k ) - 1+ 2 B s + 2V u + i E v-i V 4 x& 4 ] = 0.

(35)
Recalling the field equations of Eqn. (17), even a particle without an electric
charge (that is an electrically neutral particle) may have a magnetic moment
because, for = J = 0, one finds

E = - a0 V 4

x B -

1 E
V
= - a0
.
c t

(36)
If these new fields are to be interpreted as the gravitational fields then Eqn.
(36) may be interpreted as requiring a magnetic moment for spinning,
gravitating particles.
An interesting result occurs when one looks at the allowed fundamental
spin states. In the five-dimensional quantization of the space-time-mass
manifold, three spin vectors appear.
One of these is the familiar
three-component spin vector of relativistic quantum mechanics; the second of
the three is a new three-component spin vector; the remaining is a
four-component spin vector defined below.
Using the theorem, if satisfies 2 = a2 where a is a number, then the
eigenvalues of are +a, it is not difficult to show that the component
eigenvalues are:
s =

1
1
3
2
, u = , S j = ; = 1,2,3 j = 1,2,3,4.
2
2
4

(37)
If, in analogy with the eigenvalues for the total angular momentum, one writes

24

Sj =

3
= S j ( S j + 1)
4

then the possible eigenvalues become


s = + _

1
1
1 3
, u = + _ , S j = ,- .
2
2
2 2

However, the following relations, which specify the components of a


four-dimensional spin vector which, when added to the angular
four-momentum, commutes with the specific Hamiltonian, restrict the number
of possible combinations of these eigenvalues.
S 1 = s1 - u 2 - u 3 , S 2 = s 2 + u 1 - u 3
S 3 = s 3 + u 1 + u 2 , S 4 = s1 - s 2 + s 3 .

The question to be asked now seems to be, how many combinations of the above
eigenvalues are allowed? The answer may be shown to be octets. This
predicted result compares with the experimental findings of Gel Mann.
By deriving the quantization conditions and using London's derivation of
the quantum mechanics from this condition one obtains classical atomic physics
by assuming that the effects of the gravitational gauge field components may be
neglected. Thus, there appears to be no effect of the proposed theory upon the
atomic physics that is now known.
There is an astonishing effect of the Neo-Coulombic potential upon how
one might describe nuclear phenomena. One of the first features noted about
the potential was its return to a zero value as the radial value approaches zero.
This has the effect of producing a force given by,
1
q k
F = 12 1 1 e r .
r
r
(38)
If this force is repulsive when r is infinite for like particles, it becomes zero
when the separation is at the distance lambda and becomes a strongly
attractive force when the separation becomes less than lambda. This is just the
sort of behavior found when proton-proton scattering was first done at high
enough energies to see a deviation from Coulombic scattering. The expression
for the Neo-Coulombic scattering cross-section was found to be

q1 q 2 2 sin d

d =
,
2
2mV 0 sin 4 ( )

25

where
2

1

4E
1+ 6
sin 4 ( )1+ ( - ) tan( )
k
2 2
2

=
.
4
2
3 4E

2
1 +
sin ( ) sin ( - )
2
2 k

This scattering cross section for like-particle interaction appears to have the
right dependencies to explain the scattering data. It remains to compare
prediction with existing experimental data to determine the validity of the
predictions and the ability of the Neo-Coulombic potential to explain the Strong
Nuclear Force with that portion of its radial dependence that causes the value
of the potential to return to zero.
When unlike particles are considered care must be taken to keep the
lambdas in the forces straight. The force on any charged particle due to the
presence of another, second particle, is the product of the charge of the first
particle and the field of the second particle. Thus, the force on the first particle
goes to zero at the lambda of the second. For the force on a proton due to the
field of an electron
k e - e
F p = q p E e = - 2 1 - e r
r r

(39)
while the force on an electron due to the field of a proton is
k p p
F e = q e E p = 2 1 - e- r .
r
r

(40)
By looking at proton and electron like-particle scattering data it would appear
that the lambda of the proton must be much larger than that of the electron. If
this is the case then the force on the electron due to the near presence of the
proton goes to zero while the proton is still attracted to the electron. Any
further decrease in the separation causes the electron to experience a repulsive
force; although the proton is still attracted to the electron. This immediately
raises the eyebrows. Can it be that Newton's Third Law, concerning the equal
and opposite forces, does not hold in Nature? The answer is, certainly.
Newton's Third Law does not hold in high-speed electromagnetic interactions
when viewed by the retarded potentials; it was found to be violated during beta
decay until the hypothesis of the neutrino reinstated the summation of particle
spins. Should one then throw out the unlike-particle forces because they violate
Newton's Third Law without seeing what predictions these forces might lead
to?

26

If one proceeds with the unlike-particle forces, he finds very quickly that
it appears possible that the proton might find a very close orbit, at a separation
from the electron by a distance lambda, in which it could settle down into a
Bohr orbit around the electron. On the other hand the electron would
experience no force from the orbiting proton. Such a state might cause one to
think of the neutron. Here one runs into the question of particle spins that beta
decay brought out and which led to the hypothesis of the neutrino. Also, the
argument is offered that Heisenberg's Uncertainty Principle requires that the
electron could never be in an orbit so tightly bound that the orbit is less than
nuclear separations. This argument hinges upon the unit of action being
Planck's constant. But remember the dependence of the Poisson brackets upon
the geometry?
Another argument against the neutron being an electron and proton in
nuclear-sized orbits is based on an argument that the principle of angular
momentum cannot be conserved. The neo-coulombic forces, which require that
the force between the electron and proton be directed on a line between them,
also requires that the angular momentum be conserved. However, the unit of
action depends upon the gauge function and this requires that, when Bohr-type
orbits are considered, there is an effective unit of action for the electron orbit
and a different effective unit of action for the proton orbit. Thus, the effective
unit of action for the electron orbit requires that in the neutron the orbital
angular momentum would be given by he and its intrinsic spin angular
momentum would be (1/2)he. Similarly, for the proton the orbital angular
momentum would be hp and the spin (1/2) hp.
After the neutron decays, the angular momentum is the sum of the two
particles' intrinsic spin angular momenta, which is given by
because both
particles are free and therefore, each has an intrinsic spin angular momentum
of (1/2) h. Thus, the conservation of angular momentum is expressed as
1
(+_ h e + _ h p ) + he + h p = h.
2

(41)
Experimental evidence of orbital and/or spin angular momentum is contained
in the experimental magnetic moments. If one equates the intrinsic and orbital
magnetic moments of the electron and proton while they are in the orbital
configuration to the experimental value of the neutron's magnetic moment they
have
+_

hp
1 he
B + _ n 2 h
h

(42)

27

where B is the Bohr magneton and n is the nuclear magneton. Eqn.s (41) and
(42) require that he =8.0517 x 10-4h and hp = 0.66585h. Thus, within the
proposed theory the neutron appears to be a proton in orbit around an electron.
Not surprisingly then, it is possible to build a nuclear model of the
protons-around-electrons, and electrons-around-positrons, states that allow one
to predict the masses of the nuclei which have a mass number less than 10 amu
with better RMS error than the best of the semi-empirical mass formulas have
for mass numbers greater than oxygen13. This should possibly be considered all
the more significant since the semi-empirical mass formulas have ever
increasing errors for the low mass numbers and are not even used below an
amu of 16. It remains for this nuclear model to be extended to the higher mass
numbers, but it appears from the work done thus far that one can only expect
that the correspondence with experiment will improve with increasing mass
numbers.
Is it possible that the Neo-Coulombic forces can explain the phenomena
associated with the weak forces? Certainly the nuclear mass predictions argues
that a nuclear model based upon these orbits does not miss far and is a much
cleaner model than currently used. Initial looks at the neutrino experiments
using the proposed theory offer other explanations for these experimental
results but are too lengthy to include here. It should be remembered that these
experiments must be explained by the proposed theory if the unlike forces are
to fully account for phenomena that the weak nuclear forces are now thought to
explain.
The long range 1/r2 dependence of the new three-dimensional vector
gauge field component suggests that these components are the components of
the gravitational field. If this is to be the case the proposed theory must then
explain the same phenomena that the General Theory of Relativity predicts.
First, note that the gravitational field components in the gauge field tensor
must have units equivalent to the electric field components. Following up on
this, one finds that a charge-to-mass ratio is needed to convert the gravitational
field units from the familiar units of acceleration to the volts/meter units used
in the gauge field tensor. By considering the new fields and comparing them to
the currently used fields one finds that this ratio is given by the square root of
the product of the gravitational constant and the dielectric constant, or
= G = 2.4296x 10-11 coul/kg.

(43)
An interesting result follows immediately. If the fundamental charge-to-mass
ratio works as it appears to, and electrically neutral spinning bodies have
magnetic moments, then the predictions of magnetic moments for electrically
neutral bodies may be made by determining the effective charge density of the
rotating gravitating body using the charge-to-mass ratio and the spin of the
body. A simple calculation of the earth's magnetic moment, assuming uniform

28

mass distribution, by this method produced a prediction of the magnetic


moment 1.06 times the actual value14. This prediction seems surprisingly close
considering the uncertainties in the density measurements of the mass
distribution of the earth.
One of the predictions of Einstein's General Theory of Relativity
concerns the tendency of light from stars and other objects in the heavens to be
shifted towards the red color end of the spectrum. Looking at the emission and
reception of light within the framework of the proposed theory one finds that
the unit of action, which establishes the energy of any state, depends upon both
the relative time and the gravitational field at the time and place of the
emission and reception. This is so because the theory holds the gravitational
field to be a gauge field and it is the gauge function that determines the
applicable unit of action. It is not difficult to show that

j
[ x j , p k ] = ih g kl jl + x s .

sl

(44)
Thus, for a metric with only a gauge function the effective unit of action would
be given by
h = h exp[2 f t f r f ].

(45)
By recalling the gauge gravitational field of Eqn. (25), one may use Eqn. (45) to
find the expression for the unit of action for emission of a photon to be
W e (1 + bt e ) - e
e Re
h e = h exp
Re

(46)
where the subscript, e, denotes emission. Similarly, the unit of action for the
reception of a photon can be found to be
W r (1 + bt r ) - r
e Rr .
h r = h exp
Rr

(47)
If photon energy is conserved between emission and reception then
h e e = h r r .

(48)

29

If one sets te = 0, tr = L/c, W = (-GM/c2), and b = -H, then they find the shift in
frequency given by
By looking at the first order approximations of this prediction one finds
that the time dependence of the gravitational field produces the linear
dependence and is given by Hubble's constant while the gravitational potential
produces the same prediction that comes from Einstein's theory.

= exp 2
e
c

r
Rr
M e Re
M r e
e

Re
Rr

r
HL Rr
+ c e 1.

Looking a little closer one finds that the time dependence of the red shift
produces an experimental number, H-1 = (5.6+0.6) x 1017 sec. (1.61 x 10-18 sec-1 <
H < 2.0 x 10-18 sec-1), that corresponds to the same time dependence that has
been measured and reported for the moon's orbit15 (b=1.9 x 10-18 sec-1), well
within experimental error. It is somewhat pleasing that a prediction coming
from the same time dependence originating in the gauge function leads to a
comparison of phenomena involving cosmological distances agrees with
phenomena involving the much shorter distance involved in the moon's orbit.
Another possible plus to this prediction is that, because the prediction involves
an exponential dependence upon time and gravitational potential between the
emission and reception of the light, then the distances that are currently
ascribed to distant bodies by their red shifts may be much greater than the
actual distances. Also, the possible red shifts from dense gravitating bodies
may be much greater than is now believed possible thereby removing the
mystery from many objects.
The time dependence of the gravitational field stems from the principle
increasing entropy and is a direct result of this inflation-like effect imposed
upon the universe by the denial of perpetual motion. An additional implication
follows for the use of dating processes which depend upon radioactive processes
in that the unit of action changes with time in accordance with that same time
dependence. The results would be that all of the dates would have to be
adjusted downward.
The prediction of the advance of the perihelion of the planetary orbits is
the one prediction of Einstein's General Theory of Relativity that requires the
entire formal theory. Within the proposed theory one obtains an advance to the
planetary orbital perihelion by simply using the low velocity Newtonian
equations of motion with the Neo-Coulombic gravitational potential, which is

30

(49)

3 GMm2
.
2
L2

(50)
The perihelion advance predicted by the General Theory of Relativity is given
by16
3 G2 M 2 m2
.
GTR = 2
2 2
c L

(51)
Thus, the lambda of the sun would have to be given by
sun =

GM
c

(52)
if the proposed theory is to be identical in its prediction of planetary perihelion
advance to Einstein's General Theory of Relativity.
Currently there is much discussion of experimental evidence of the need
for a fifth, and even a sixth, force in Nature. The evidence points to a decreased
gravitational strength when compared with Newtonian gravitation. Consider
the Neo-Coulombic gravitational force which must go to zero at some value of
distance that is representative of the body in question. The obvious conclusion
is that the gravitational force in the proposed theory must become less than the
Newtonian value as distance is decreased. Thus, a new independent force may
not be necessary at this time.
There are numerous implications of this feature of the Neo-Coulombic
force which will have large effects upon the concept of the universe presented
by the proposed theory. For example, a gravitational force which becomes
repulsive with decreasing distance denies the type of gravitational collapse now
discussed by cosmologists. Neither can it support the singularities now called
Blackholes. The possibility of the existance of distant bodies so massive that
light cannot escape their gravitational pull has not yet been investigated.
A number of possible experimental tests have been considered. A few of
these will be presented here.
The proposed theory presents a picture of the universe in which the
electromagnetic and gravitational fields are components of a single gauge field
tensor and, therefore, are fields on equal footing and also, more importantly,
inductively coupled. This implies that manipulation of one field will inductively
produce another of the fields. It is this type of inductive coupling which causes
a magnetic field to be created by the flow of current. The electric field which is
the source of the voltage in the alternator providing the power for home use was

31

inductively created by passing a conductor through a magnetic field. Is it not


then possible to create a gravitational field by the manipulation of the
electromagnetic fields if the inductive coupling presented in the proposed
theory exists?
Where might this inductive coupling most likely show up? One area of
phenomena is in wave properties such as electromagnetic or, in this case,
electromagnetogravitic waves. The five-dimensional wave solutions have an
additional transverse field component17 which is opposite in direction to the
electric field component. This additional component is the gravitational field
component. One of the results of the possible existence of this gravitational
component is that while the wave energy density depends upon the sum of the
squares of all wave components, the radiation pressure depends upon the sum
of the squares of the electric and magnetic components, but the square of the
gravitational component is subtracted from the sum of the others. This implies
that the radiation pressure would always be a little less than the energy
density rather than always equal to it. The initial experiments on radiation
pressure and energy density showed just this difference, however, the
difference was within experimental error. To date the known experimental
techniques do not appear to have sufficient accuracy to measure the expected
difference in these quantities.
Another experimental technique which has a much better chance of
detecting the new wave component is the neutron interferometer device. Here
the gravitational component is directed opposite the electric and a polarized
laser beam may be used to deflect one leg of the neutron's path without causing
an interaction between the magnetic component and the neutron's magnetic
moment. The sensitivity of the interferometer is such that even an extremely
small amount of energy in this component might be discernable.
The phase velocity of the five-dimensional waves are found to be
dependent upon any divergence in the flow of a medium through which the
wave is passing. This allows the possibility of slowing down the wave
significantly by causing a divergent flow. The divergence possible from nozzles
in continuous flow is too small to allow for other difficult factors affecting the
speed of light, such as the index of refraction, to be accounted for with sufficient
effect left over for clean measurement of the slow down. On the other hand, if
the divergent flow were created using explosives, the one might be able to slow
down gamma rays to about half the speed of light. This would involve all the
usual difficulties of one-shot testing plus some other possible problems.
The very nature of the five-dimensional manifold places restrictions
upon some phenomena. For instance, when looking at shockwaves in material
using the proposed theory it is found that the phenomena is predicted on a
four-dimensional hypersurface embedded by the conservation of mass within
the five-dimensional space. This has an effect that appears like a viscosity and
puts a very distinctive anti-symmetrical profile into the shock front. The recent
advances in our abilities to measures differential times and distances make it

32

appear possible to measure the rise of a shock by using a Laser Velocity


Interferometer with an electronic streak camera. The predicted asymmetry is
such that it would be easy to discern it from the classical symmetrical
Newtonian viscosity.
The preceding has presented the fundamental laws of the proposed
theory and how each of the existing theories may be shown to be either within
the scope of the new theory or superseded by it. All of the results of the
assumption of the laws presented here have, of course, been arrived at through
rigorous mathematical logic using these laws as the starting point. The
mathematical derivations have been left to later chapters in order to provide
for a better flow of the overview discussion and to limit the length of the
overview.
The single most important concept hoped to have been conveyed in the
preceding is that the classical laws of thermodynamics contain within them the
generality, applicability, and strength to allow them to provide the basis for a
description of a nature much more general than the sum of all the currently
known theories and contain within it the current theories as subsets.
Starting from its general five-dimensional form, the theory provides a
metric in the form of the stability conditions. Two variational principles are
given by the basic laws. The first, and the more general, is a Principle of
Minimum Free Energy, and has not been pursued in the above discussion.
Secondly, the Entropy Principle has been used throughout the preceding
discussion to limit the realm of phenomena to those for which current theories
are used. The basic laws determined the type of geometry of the metric. The
variational principle stemming from the Entropy Principle was used to obtain
the equations of motion and the field equations when the appropriate
restrictions were imposed. It was the restrictions employed that allowed the
concentration upon phenomena related to certain current theories.
To help interpret the five-dimensional field components, the isentropic
restriction was imposed. This restriction required a quantization from which it
was found that one could derive quantum mechanical equations of motion,
following London's work. When the characteristics of fundamental particle
gauge potentials were sought which satisfied this quantization condition it was
found that the Neo-Coulombic potential appeared, requiring that the
gravitational field and potential be components of a gauge field on equal footing
with the electric and magnetic fields. Thus, the isentropic restriction produces
the subset in which Quantum Mechanics and the fields of the fundamental
particles are found.
If the system is restricted to be an isolated one and one looks at the
trajectories required by the Entropy Principle, he finds that they are given by
equations of motion in five dimensions. By saying that mass density may be
written as a function of space and time one finds that the trajectories lie on a
four-dimensional,
space-time,
hyper-surface
embedded
within
the
five-dimensional manifold. If one further restricts his attention to those events

33

near equilibrium states, the metric may be approximated by a flat metric, and
one finds the equations of motion to be those of Einstein's Special Theory of
Relativity. A further restriction to slow moving things brings about the
reduction to Newtonian equations of motion.
Turning from equations of motion to the forces of Nature, the proposed
theory presents only one type of force, the gauge force, which shows up in three,
three-component vector fields plus a scalar field. These fields correspond to the
fields now known as the electric, magnetic, gravitational fields and the
gravitational potential. Because the proposed theory displays the three forces
together in a single five-dimensional field one probably should refer to all three
as components of the electromagnetogravitic (EMG) force.
The theory appears to describe the phenomena currently described by
the Strong Nuclear Force by the Neo-Coulombic electrostatic force which
reverses its sign as the separation of like particles is reduced. For the Weak
Nuclear Force the theory offers the asymmetrical unlike-particle force.
The Neo-Coulombic gravitational force not only provides the classical
gravitational predictions plus the planetary perihelion advance prediction, but
includes a prediction which appears to correspond to the recently observed
experimental results which have brought forth talk of a fifth force in nature.
Currently used cosmological and gravitational red shifts were found to
be the first-order approximations to the red shift predictions from the proposed
theory. The full exponential character of the time and gravitational potential
dependence of the red shifts may find usefulness in helping to describe the
universe by helping to clear up some of the mysteries of the cosmos.

34

References:
1

O.F. Mosotti, 'Sur les forces qui regissent la constitution interieur des corps',
Turin (1836). Mossotti's essay was translated into English and published in
Taylor's Scientific Memoirs, 1, 448 (1839).

P. E. Williams, 'The Dynamic Theory: A New View of Space, Time, and Matter',
Los Alamos National Laboratory Report, LA-8370-MS, pp. 11-18 (1980). (see
also Section 2.2)

pp. 39-49 of Ref.[2]. (see also Section 2.2)

H. Weyl, 'Space-Time-Matter', (1922).

F. London, 'Quantum Mechanische Deulung der Theorie, von Weyl', z. Physik,


{\bf 42\/}, 375-389 (1927).

pp. 87-91 of Ref.[2]. (see also Section 3.4)

pp. 111-119 of Ref.[2]. (see also Section 4.6)

P. E. Williams, 'The Possible Unifying Effect of the Dynamic Theory', Los


Alamos National Laboratory Report, LA-9623-MS pp.57-61 (1983). (see also
Section 4.9)

pp. 76-77 of Ref.[8]. (see also Section 4.13)

10

P.G. Bergmann, 'Basic Theories of Physics: Heat and Quanta', Dover (1950).

11

pp. 111-124 of Ref.[2]. (see also Sections 4.1-4.7)

12

pp. 119-124 of Ref.[2].

13

pp. 83-99 of Ref.[8]. (see also Section 4.13)

14

D. Halliday and R. Resnick, 'Physics', Third Edition, Wiley (1978). (see also
Section 5.1)

15

T. Van Flandern, 'Is Gravity Getting Weaker?', Sci. Am. (1976).

16

R. Adler, M. Bazin, and M. Schiffer, 'Introduction to General Relativity', Second


Edition, McGraw-Hill (1965).

17

pp. 38-54 of Ref.[8]. (see also Chapter 6)

35

CHAPTER 2

NEW THEORETICAL FUNDAMENTALS

The Dynamic Theory uses a different viewpoint, or approach, to


present a description of physical phenomena. Therefore the first criterion
that it must meet is that it must not be in conflict with existing theories in
a field of physics where the existing theory gives an adequate and accurate
description. To show that the Dynamic Theory meets this criterion, this
section will present the adopted laws and then proceed to show how the
fundamental principles of existing theories may be obtained from these
laws.
A.

General Laws

In the following development physical concepts are necessary, as are


symbols for these concepts. Because this development will merge certain
thermodynamic conceptualizations into mechanics, a notational dilemma
must be faced.
On the one hand, it is desired to preserve the
thermodynamic conceptualization by using familiar symbols from that
theory. On the other hand, descriptions of mechanical systems are also
sought. The formulism then looks either like thermodynamics with
familiar thermodynamic quantities replaced by mechanical quantities, or it
looks like mechanics into which thermodynamic quantities intruded. In
either case there is danger of confusion. One could avoid the dilemma by
choosing entirely different symbols for the variables of the theory. But
then the whole takes on an artificially abstract character. Since the
purpose of this formulation is to bring out the power of the thermodynamic
conceptualization, it was decided to use the suggestiveness of the
thermodynamic or mechanical symbols whenever convenient; the reader is
asked to keep an open mind and not make premature association with the
symbols used.
2.1

First Law.

The concept of conservation of energy is fundamental to all branches


of physics and therefore represents a logical beginning for a generalized
theory. Therefore, in terms of generalized coordinates or independent
variables, the notion of work, or mechanical energy, is considered linear
forms of the type
_W = F i ( q1 ,..., q n ,u 1 ,...,u n ) dq i (i = 1,2,..., n),

where the forces Fi may be functions of the velocities (dqi/dt = ui) as well as
the coordinates qi and the summation convention is used. The inclusion of
velocities in forces reflects the belief that forces should depend upon the
velocities. This will become clearer when these work terms are included in
the first law.
36

The line integral c Fi dqi then represents the work done along the
path C by the generalized forces.
A system may acquire energy by other means in addition to the work
terms; such energy acquisition is denoted dE.
The system energy, which represents the energy possessed by the
system, is considered to be
U( q1 ,..., q n , u 1 ,...,u n ).

dU will be assumed to be a perfect differential.


With these concepts, then the generalized Law of Conservation of
Energy, which is adopted as the first law of the Dynamic Theory, has the
form
_E = dU - _W
= dU - F i dqi (i = 1,..., n).

(2.1)
Positive dE is taken as energy added to the system by means other than
through the work terms and Fi is taken as the component of the
generalized force acting on the system which caused displacement dqi.
In the First Law the dimensionality is n + 1 and is determined by the
system considered. There is no limitation on the quantity or type of
variables that may be used. However, in this presentation and in practice,
it will be beneficial to place restrictions upon the type and number of
allowed work terms. A system with only one work term, which is the pdv
expansion work of classical thermodynamics, will be called a
"thermodynamic" system and the dimensionality will be two. A system
with three or less fdx work terms will be called a "mechanical" system with
the appropriate dimensionality. Obviously, if there are three mechanical
work terms, the dimensionality will be four. A system with a combination
of thermodynamic and mechanical work terms will be considered later.
In an infinitesimal transformation, the First Law is equivalent to the
statement that the differential
dU = _E + F i dq i

is exact. That is, there exists a function U whose differential is dU; or the
integral dU is independent of the path of the integration and depends only
on the limits of integration. This condition is not shared by dE or W .
The path dependence of W is another reason that the generalized forces
are assumed to be functions of velocity as well as position. In Newtonian
mechanics forces are usually assumed to be dependent on position only so
that the simplicity of path independence may be used. Though even in
Newtonian mechanics certain forces are taken as velocity dependent.
Friction forces are an example.
37

This statement of the generalized First Law is consistent with the


First Law of thermodynamics in that if there is only one generalized force,
which is taken to be the pressure, and one generalized coordinate, the
volume, then Eqn. (2.1) becomes
_Q = _E = dU + Pdv

where F = -P with the convention that work of expansion is work done by


the system on its surroundings. Here the system energy, U, is the
thermodynamical internal energy. There should then be no confusion
when Cartheodory's statement of the second law is applied to this
thermodynamic system. However, when considering the application of
generalizations of the classical thermodynamic laws to mechanical systems
some confusion may be expected. During the initial portion of this
development, it is desired to demonstrate the applicability of the
generalized laws to mechanical systems. Therefore, it may help avoid
confusion to think of the generalized coordinates of a mechanical system
as the space coordinates of a mass point. Obviously, there exists systems
in nature that may be considered to consist of a continuous distribution of
mass points. Such a system may be thought of as a composite system of
an infinite number of subsystems and, therefore, involve an infinite
number of "generalized coordinates," or "degrees of freedom." However,
just as in classical mechanics, we may later make the transition from mass
points to matter in bulk; then the generalized coordinates, qi, used here
may better be termed independent variables.
To explore some of the consequences of the exactness of dU,
consider a system whose variables are F, q and u. The existence of the
state function U, or an equation of state, means that any pair of these
three parameters may be chosen to be the independent variables that
completely specify the system. For example consider U = U(F,q) then
U
F

dU =

dF +
q

U
q

dq.
F

The requirement that dU be exact immediately leads to the result


q

U
F

=
q F

U
q

.
F q

The "energy capacity" of a system at the position q with dq = 0 may be


defined as
Cq =

_E
_u

=
q

U
u

and the "energy capacity" of a system under a constant force is defined as


38

CF =

2.2

_E
_u

=
F

U
u

.
F

Second Law.

There are processes that satisfy the First Law but are not observed
in nature. The purpose of the dynamic second law is to incorporate such
experimental facts into the model of dynamics.
The statement of the Second Law is made using the axiomatic
statement provided by the Greek mathematician Caratheodory, who
presented an axiomatic development of the Second Law of thermodynamics
that may be applied to a system of any number of variables. The Second
Law may then be stated as follows:
In the neighborhood (however close) of any equilibrium state of
a system of any number of dynamic coordinates, there exist
states that cannot be reached by reversible E - conservative
(dE = 0) processes.
When the variables are thermodynamic variables, the E-conservative
processes are known as adiabatic processes.
A reversible process is one that is performed in such a way that, at
the conclusion of the process, both the system and the local surroundings
may be restored to their initial states without producing any change in the
rest of the universe.
Consider a system whose independent coordinates are a generalized
displacement denoted q, a generalized velocity u (with u = dq/dt), and a
generalized force F. It can be shown that the E-conservative curve
comprising all equilibrium states accessible from the initial state, i, may be
expressed by (u,q) = constant, where
represents some as yet
undetermined function. Curves corresponding to other initial states would
be represented by different values of the constant.

39

Figure 1. If two reversible E-conservative curves could


intersect it would be possible to violate the Second Law by
performing the cycle i, f1, f2, i.
Reversible E-conservative curves cannot intersect, for if they did, it
would be possible, as shown in Figure 1, to proceed from an initial
equilibrium state i, at the point of intersection, to two different final states
f1 and f2, having the same q, along reversible E-conservative paths, which
is not allowed by the Second Law.
When the system can be described with only two independent
variables, such as on the E-conservative curve, then if these variables are q
and u and F is a generalized force,
_E = dU - Fdq.

Regarding U = U(q,u), then


U
u

_E =

U
q

du +
q

- F dq
u

where all quantities on the right-hand side are functions of u and q.


An E-conservative process for this system is

U
u

du +
q

40

U
q

- F dq = 0.
u

Solving for du/dq yields

du
=
dq

U
q
U
u

-F
u

.
q

(2.2)
Figure 2. The First Law fills the (u,q) space with slopes. The
curves represent the solution curves whose tangents are the
required slopes. The Second Law requires that these curves do
not intersect.

The right hand member is a function of u and q, and therefore, the


derivative du/dq, representing the slope of a E-conservative curve on a
(u,q) diagram, is known at all points. Equation (2.2) has therefore a
solution consisting of a family of curves, see Figure 2, and the curve
through any one point may be written
= (u, q) = constant.

A set of curves is obtained when different values are assigned to the


constant. The existence of the family of curves (u,q) = constant, generated
by Eqn. (2.2) representing reversible E-conservative processes, follows from
the fact that there are only two independent variables and not from any
law of physics. Thus it can be seen that the First Law may be satisfied by
any of these
= constant curves. The Second Law requires that these
curves do not intersect. Therefore the Second Law, together with the First
law, leads to the conclusion that through any arbitrary initial-state point,
all reversible E-conservative processes lie on a curve, and E-conservative
curves through other initial states determine a family of non-intersecting
curves.
To see the results of this conclusion consider a system whose
coordinates are the generalized velocity u, the generalized displacement q
and the generalized force F. The First Law is
_E = dU - Fdq

where U and F are functions of u and q. Since the (u,q) surface is


subdivided into a family of non-intersecting E-conservative curves (u,q) =
constant where the constant can take on various values 1, 2, ..., and
points on the surface may be determined by specifying the value of along
with q, in all regions where the Jacobian of the transformation does not
41

vanish, so that U, as well as F, may be regarded as functions of


Then
U

dU =

U
q

d +
q

and q.

dq

and
_E =

U
q

d +
q

- F dq.

Since and q are independent variables this equation must be true for all
values of d and dq.
Suppose d = 0 and dq
0. The provision that d = 0 is the
provision for an E-conservative process in which dE = 0. Therefore, the
coefficient of dq must vanish. Then, in order for and q to be independent
and for dE to be zero when d is zero, the equation for dE must reduce to
_E =

d ,
q

with
U
q

Defining a function

= F.

by
U
q

then dE = d where = ( ,q).


Now, in general, an infinitesimal of the type
Pdx + Qdy + Rdz + ...,
known as a linear differential form, or a Pfaffian expression, when it
involves three or more independent variables, does not admit of an
integrating factor. It is only because of the existence of the Second Law

Figure 3. Two reversible E-conservative curves, infinitesimally


close. When the process is represented by a curve connecting
the E-conservative curves, energy dE = d is transferred.

that the differential form for dE referring to a physical system of any


number of independent coordinates possesses an integrating factor.
Two infinitesimally neighboring reversible E-conservative curves are
shown in Figure 3. One curve is characterized by a constant value of the
function A, and the other by a slightly different value A + d = B. In any
42

process represented by a displacement along either of the two


E-conservative curves dE = 0. When a reversible process connects the two
E-conservative curves, energy dE = d is transferred.
The various infinitesimal processes that may be chosen to connect
the two neighboring reversible E-conservative curves, shown in Figure 3,
involve the same change of but take place at different . In general is a
function of u and q. However, it is obvious that may be expressed as a
function of
and u. To find the velocity dependence of
consider two
systems, one and two, such that in the first system there are two
independent coordinates u and q and the E-conservative curves are
specified by different values of the functions
of u and q. When dE is
transferred, changes by d and dE = d where is a function of and q.
The second system has two independent coordinates u, and q' and
the E-conservative curves are specified by different values of the function '
of u and q'. When dE is transferred, ' changes by d ' and dE = 'd ' where
' is a function of ' and u.
The two systems are related through the coordinate u in that both
systems make up a composite system in which there are three independent
coordinates u, q, and q' and the E-conservative curves are specified by
different values of the function c of these independent variables. To help
visualize the situation it may be noted that the composite system is, in
essence, two particles joined together and traveling with the same velocity
but not sharing the same location.
Since = (u,q) and ' = '(u,'), using the equations for and ', c
may be regarded as a function of u, and '.
For an infinitesimal process between two neighboring
E-conservative surfaces specified by c and c + d c, the energy transferred
is Ec = cd c where c is also a function of u, and '. Then
d

du +

d +

(2.3)

Now suppose that in a process there is a transfer of energy dEc


between the composite system and an external reservoir with energies dE
and dE' being transferred, respectively, to the first and second systems,
then
_ E c = _E + _E

and
c

= d + d

or
d

d +

Comparing Eqns. (2.3) and (2.4) for d

c
c

then

= 0.

43

(2.4)

Therefore c does not depend on u, but only on


c( , '). Again comparing the two expressions for d
=

also

and '. That is


we find

Therefore the two ratios / c and / c are also independent of u, q


and q'. These two ratios depend only on the 's, but each separate must
depend on the velocity as well (for example, if depended only on and on
nothing else, the dE = d would equal f( )d which is an exact differential).
In order for each to depend on the velocity and at the same time for the
ratios of the 's to depend only on the 's, the 's must have the following
structure:
= (u) f( ),
= (u) f ( ), with
).
c = (u) g( ,

(2.5)

(The quantity cannot contain q, nor can contain q', since and / c must
be functions of the 's only.)
Referring now only to the first system as representative of any
system of any number of independent coordinates, the transferred energy
is, from Eqns. (2.5),
_E = (u) f(

)d .

Since f( )d is an exact differential, the quantity 1/ (u) is an integrating


factor for dE. It is an extraordinary circumstance that, not only does an
integrating factor exist for the dE of any system, but this integrating factor
is a function of velocity only and is the same function for all systems. It
would be nice if there were a simple way of deriving the functional form of
(u). In thermodynamics we opted to take the easy way out by assuming
that the integrating factor was simply the reciprocal of the temperature.
However, for mechanical systems we will find the functional form of the
integrating factor when we determine the equations of motion.
The fact that a system of two independent variables has a dE that
always admits an integrating factor regardless of the axiom is interesting,
but its importance in physics is not established until it is shown that the
integrating factor is a function of velocity only and that it is the same
function for all systems.

2.3

The Absolute Velocity and Einstein's Postulate.


44

The universal character of (u) makes it possible to define an


absolute velocity. Consider a system of two independent variables q and u,
for which two constant velocity curves and E-conservative curves are

Figure 4. Two constant velocity energy transfers, E3 at u from b


to c and E3 at u3 from a to d, between the same two conservative
curves 1 and 2.

shown in Figure 4. Suppose there is a constant velocity transfer of energy


E between the system and its surroundings at the velocity u, from a state
b, on a E-conservative curve characterized by the value 1, to another state
c, on another E-conservative curve specified by 2. Then since it is seen
that
_E = (u)

2
1

f(

)d

at constant u.

For any constant velocity process between two other points a to d,


at a velocity u3 between the same E-conservative curves the energy
transferred is
_E( u 3 ) = _ E 3 = ( u 3 )

2
1

f(

)d

at constant u 3 .

Taking the ratio of


E = (u) = a function of the vel. at which E is transferred.
E3 (u3) same function of vel. at which E3 is transferred
Then the ratio of these two functions is defined by
(u) = E (between 1 and 2 at u)
(u3)
E3 (between 1 and 2 at u3)
or
_E =

_ E3
( u3 )

(u).

By choosing some appropriate velocity u3 it follows that the energy


transferred at constant velocity between two given E-conservative curves
decreases as (u) decreases, or the smaller the value of E the lower the
corresponding value of (u). When E is zero (u) is also zero. The
corresponding velocity u0 such that (u0) is zero is the "absolute velocity".
Therefore, if a system undergoes a constant velocity process between two
E-conservative curves without an exchange of energy, the velocity at which
this takes place is called the absolute velocity.
The definition of the absolute velocity requires constant velocity
processes be considered. All Galilean frames of reference will display this
process as one of constant velocity. Further, if all reference frames are to
be of equal status then observers in all Galilean reference frames must
share the dE = 0 constant velocity process equivalently. Furthermore,
45

each observer will have the same value for the absolute velocity or else one
of the frame will enjoy a privileged nature.
Just as the absolute temperature in classical thermodynamics is a
limiting quantity we may suspect that the absolute velocity will also turn
out to be a limiting quantity. Because of our experimental evidence that
the speed of light behaves as a limiting velocity when electromagnetic
forces are involved and the absolute velocity is independent of the force or
type of system and is therefore unique, it must be the speed of light. Thus,
the first two laws of the Dynamic Theory require Einstein's postulate
concerning the speed of light.
To be more specific, the absolute velocity is unique for all Galilean
frames of reference. There is one such velocity already known and that
velocity
is the speed of light, c. Therefore, the absolute velocity must be the speed
of light and the same for all Galilean observers. This is Einstein's
postulate.
The above may be put on a more rigorous basis by observing that
for the E-conservative process
_E = 0 =

U
u

du +

U
q

-F

dq .

If dE = 0 is to be invariant for all points q then we must have


U
q

- F =0

and thus du = 0, for all = 1,2,3. Thus the allowed transformations are
those with constant velocities. This, of course, was just what was required
by the statement of, or restriction to, a constant velocity process. Then all
Galilean observers will agree upon the identification of an E-conservative
system in absence of any work on the system.
Now let us suppose that at the time t, a system is at point p(q1, q2,
q3) in Q. If our system is E-conservative and traveling at the absolute
velocity, c, then in dt seconds it will be at the point q + dq where dq = u
cdt. Now the speed is given by
u u =

g u u = c,

where g is the metric for the space and the metric is parameterized using
the absolute velocity, c, which is the only velocity with an adequate
definition thus far.
Now an observer in another frame Q' sees the system at the point P'
given by (q1, q2, q3) at the time t. In dt seconds the system will have moved
to a point given by q + dq and the speed will be given by
u u

u u

= c

or
g

dq dq

46

= c 2 dt 2

since the process must specify the E-conservative process at the absolute
velocity, c.
But, since the Q observer must be Galilean then
= a q + a4 t

t = a 4 q + a 44 .

If we specialize so that g =
, g' =
, (ie. Euclidean) and we
specify the relative motion between q and q' to be only in the q' direction,
then our transformation is of the form
q = a11 q + a14 t
t = a14 q + a 44 t.

(2.9)

Substituting equation (2.9) into (2.7) we find that


1 1
2 4 4
a 4 a1 = c a1 a 4

( a11 )2 = 1 + c 2 ( a14 )2
2
4 2
2
1 2
c ( a4 ) = c + ( a4 ) .

These are three equations in four unknowns.


relation. But for the moment we have

We need one further

( a14 )2 = c 2 ( a 44 )2 - c 2 = c 2 [( a 44 )2 - 1]

and
( a11 )2 = 1 + c 2 ( a14 )2 .

If the q' moves with constant velocity v with respect to q then


1
1
1
1
dq
a dq + a 4 dt
a1 u + a 4
= 41
=
.
4
4
4
dt
a1 dq + a 4 dt
a1 7u + a 4

For u = 0,
1
dq
a
= 44 = - v.
dt
a4

Thus a41 = -va44 which implies


( a14 )2 = v 2 ( a 44 )2 = c 2 [( a 44 )2 - 1],

or
+_ 1

a4 =

+_ .

1 - v2
c

Then a41 = + v, and since


1
a1 =

1 + c 2 ( a14 )2

we have
4
a1 = + _

v
c

and
1
a1 = + _ .

We now have
47

1
1
q = +_q -v t
2
2
q = q
3
3
q = q

t = +_

v
c

u + t.

Now if (dq'/dt) = v, this implies


dq
=
dt

+ _v + _ v
2

+ _ v2 + _ 1
c

= 0.

This means we must take the + sign for a11. If (dq'/dt) = 0 we find
dq
=
dt

+ _v
2

+ _ v2 + _ 1
c

= -v

if we take the - sign for a41. If (dq'/dt) = c then


dg
=
dt

c-v
2

cv
2
c

= c
1

if the sign of a14 is taken as - and the sign of a44 is taken as +.


Thus, we have
1

a1 =
1
a4 = - v
4
a1 =
4
a4 =

or

48

- v
c

1
1
q = ( q - vt)
2
2
q = q
3
3
q = q

t =

t-

v
c

1
q .

(2.10)
Equations (2.10) are the transformations of Einstein's Special
Theory of Relativity, which, in Einstein's derivation needed only his
postulate concerning the speed of light and the requirement that physics
be the same for all Galilean observers. Here, in the Dynamic Theory we
have shown that the Second Law requires Einstein's postulate and the
transformations of Special Relativity for Galilean observers.
It should be noted that since the absolute velocity (or the speed of
light) is unique the answer to whether there may be a different limiting
velocity for different fundamental forces is answered by the Second Law.
The Second Law states that there is only one limiting velocity independent
of the type of force considered. Note that the function defined above as
goes to zero as v tends to c. This is a property required of the integrating
factor (u) and raises suspicions concerning he functional form we will
ultimately determine for .
2.4

The Concept of Entropy.

In a system of two independent variables, all states accessible from


a given initial state by reversible E-conservative processes lie on a (u,q)
curve. The entire (u,q) space may be conceived as being filled by many
non-intersecting curves of this kind, each corresponding to a different
value of . In a reversible non-E-conservative process involving a transfer
of energy dE, a system in a state represented by a point lying on a surface
will change until its state point lies on another surface + d . Then dE
= d , where 1/ , the integrating factor of dE, is given by = (u) f( ), and
therefore dE = (u)f( )d or
_E
= f(
(u)

)d .

Since
is an actual function of u and q, the right-hand member is an
exact differential, which may be denoted by dS; and
dS =

49

_E
(u)

where S is the mechanical entropy of the system and the process is a


reversible one.
The Dynamic Theory's Second Law may be used to prove the
equivalent of Clausius's theorem, which is stated here without proof.
Theorem:

In any cyclic transformation throughout which


the velocity is defined, the following inequality
holds:
_E
(u)

0,

where the integral extends over one cycle of the transformation. The
equality holds if the cyclic transformation is reversible. Then for an
arbitrary transformation
B
A

_E
(u)

S(B) - S(A),

with the equality holding if the transformation is reversible. The proof of


this statement may be seen by letting R and I denote respectively any
reversible and any irreversible path joining A to B, as shown in Figure 5.
For path R the assertion holds by definition of S. Now consider the cyclic
transformation made up of I plus the reverse of R. From Clausius' theorem
_E
I

_E

0,

or
dE
I

dE

S(B) - S(A).

Another result of the Second Law is that the mechanical entropy of an


isolated (dE = 0) system never decreases. This can be seen since an
isolated system cannot exchange energy with the external world because
dE = 0 for any transformation. Then by the previous property of the
entropy,
S(B) - S(A)

where the equality holds if the transformation is reversible.


One
consequence of the Second Law is that of all the possible transformations
from one state A to another state B the one defined as the change in the
entropy is the one for which the integral
I

B
A

_E

is a maximum. Thus
S(B) - S(A) = max

50

B
A

1 _E
d ,
d

where
or

is a parameter that indicates position along the path from A to B,

B
A

S(B) - S(A) = max

1 dU
F dq
d
d

d .

If U = U( ,q,u,du/d ), then the change in the entropy is given by the


integral
_S =

B
A

1 dU
F dq
d
d

d .

The u and q which maximize S will be denoted as v and x then, with U =


U(x,v), F = F(x,v), and (v) the v and are given by the solution of the system
of equations
d
d

G
G
=0
x
x

d
d

G
G
=0
v
v

where
G=

-F

dx
,
d

x' = dx/d and v' = dv/d .


Thus, the Dynamic Theory's Second Law provides an answer to the
question that is not contained within the scope of the First Law: In what
direction does a process take place? The answer is that a process always
takes place in such a direction as to cause an increase of the mechanical
entropy in the universe. In the case of an isolated system, it is the entropy
of the system that tends to increase. To find out, therefore, the equilibrium
state of an isolated one-dimensional system, it is necessary merely to
express the entropy as a function of q and u and to apply the usual rules
of calculus to render the function a maximum. The equations, which
describe the path the system takes toward the maximum of entropy, are
the equations of motion for the isolated system. When the system is not
isolated, there are other entropy changes to be taken into account.
2.5

Third Law.

The Second Law enables the mechanical entropy of a system to be


defined up to an arbitrary additive constant. The definition depends on the
existence of a reversible transformation connecting an arbitrarily chosen
51

reference state 0 to the state under consideration. Such a reversible


transformation always exists if both O and A lie on one sheet of the state
surface. If two different systems are considered, the equation of the state
surface may consist of several disjoint sheets. In such cases the kind of
reversible path previously mentioned may not exist. Therefore, the Second
Law does not uniquely determine the difference in entropy of two states A
and B, if A defines a state of one system and B the state of another. For
this determination a Third Law is needed. The Third Law may be stated,
"The Mechanical Entropy of a system at the absolute velocity is a universal
constant, which may be taken to zero."
In the case of a purely
thermodynamic system the absolute quantity is the absolute zero
temperature, while for a mechanical system the absolute quantity is the
absolute velocity. The Third Law implies that any energy capacity of a
system must vanish at the absolute velocity. To see this, let R be any
reversible path connecting a state of the system at the absolute velocity u0
to the state A, whose entropy is to be found. Let CR (u) be the energy
capacity of the system along the path R. Then, by the Second Law,
S(A) =

uA
uO

C R (u)

du
.
(u)

But according to the Third Law, S(A)


0 as uA
u0. Hence it follows that
CR(u)
0 as u
u0. In particular, CR may be Cq or CF.
The statement of the Third Law above reflects the restriction to
mechanical work terms.
A general statement of third law that is
independent of the number or type of variables is "The generalized entropy
of the system, when the integrating factor vanishes, is a universal
constant, which may be taken to be zero."
B.

General Relations

2.6

Energy and Maxwell's Relations.

In thermodynamics a discussion of equilibrium and stability


conditions is best done if the enthalpy, Helmholtz's and Gibb's functions
are defined first. Therefore, the mechanical analogues of these functions
are defined here.
Each branch of physics such as thermodynamics and particle
dynamics has its own developed procedures. If both branches can be
described by the same basic laws, then the procedures developed in
thermodynamics may prove to be useful in particle dynamics and vice
versa.
Once the mechanical enthalpy, mechanical Helmholtz's and
mechanical Gibbs' functions are defined, it is then easy to write down the
resulting mechanical Maxwell and mechanical energy capacity relations.
To begin the development of the Maxwell relations, the mechanical
entropy was defined as
52

_E
.
(u)

dS

Then, since dE = dU -Fdq,


dS =

dU

dq,

where
dU = (u)dS + Fdq

(2.11)
Define the mechanical enthalpy as H = U - Fq. Then
dH = (u)dS - qdF.

(2.12)
Therefore
H
S

= (u) ;
F

H
F

= - q.
S

The mechanical Helmholtz's function can be defined as K = U (u)S, and


dK = dU -

d[ (u)]
Sdu - (u)dS
du

or, with '(u) = d /du,


dK = - S (u)du - Fdq.

(2.13)
This leads to
K
u

then

= - S (u) ;
q

K
q

= (u)F.
u

The mechanical Gibb's function may be defined as G = H -

(u)S

dG = - (u)Sdu + qdF,

(2.14)
so that
53

G
u

G
F

= - (u)S ;
F

= q.
u

From the differential Eqns. (2.11), (2.12), (2.13), and (2.14) the
Maxwell relations for a mechanical system may be written:
(u)

u
q

(u)

u
F
S
q

(u)

F
S

= q

q
S

F
u

= u

S
F

(u)

q
u

=
u

.
F

(2.15)
The energy capacity at the position q can be defined as
Cq

E
u

S
u

= (u)
q

.
q

Define the energy capacity with a constant force as


CF

E
u

S
u

= (u)
F

then
( Cq - CF ) =

(u)
(u)

q
u

F
u

,
q

and
CF
=
Cq

F
q

F
q

The three generalized laws have been formulated and a few results
of these laws have been seen. The next step is to derive the stability
conditions to obtain the quadratic forms necessary for a metric. The
derivation of the equilibrium and stability conditions is identical to the
derivation of the thermodynamic equilibrium and stability conditions with
the variables changed to represent the mechanical variables q, u, S and F
instead of the thermodynamic variables T, V, S and P.
54

2.7

Equilibrium Conditions.

To establish the criteria for equilibrium, consider, Clausius's


theorem
B
A

_E

_E

B
A

0,
R

or
B
A

_E

B
A

_E

S(B) - S(A).
R

For an E-conservative system dE = 0, then S 0, or S(B) S(A). Therefore


the mechanical entropy tends toward a maximum so that spontaneous
changes in an E-conservative system will always be in the direction of
increasing mechanical entropy.
Now by First Law E = U F q. Therefore \ S
U - F q, which
is analogous to the Clausius inequality in thermodynamics. Now consider
a virtual displacement (U,q)
(U + U, q + q), which implies a variation S
S + S away from equilibrium. The restoration of equilibrium from the
varied state (U + U, q + q)
f(U,q) will then certainly be a spontaneous
process, and by the Clausius inequality (- S) > -( U - F q). Hence, for
variations away from equilibrium, the general inequality
U -F q-

S >0

(2.16)
must hold. The inequality sign is reversed from the sign in Clausius'
inequality because hypothetical variations
away from equilibrium are
considered rather than real changes toward equilibrium.
In a spontaneous process,
S

Erev = U + work done by the system.

The "work" consists of two parts. One part is the work done by the
negative of the force F. It may be positive or negative, but it is inevitable.
Only the rest is free energy, which is available for some useful work. This
latter part may be written as
A=

E rev - U + F q.

The maximum of A is

55

Amax =

S - U + F q,

(2.17)
which is obtained when the process is conducted reversibly.
The least work, Amin, required for a displacement from equilibrium
must be exactly equal to the maximum work in the converse process
whereby the system proceeds spontaneously from the 'displaced' state to
equilibrium (otherwise a perpetual motion machine may be constructed.
Corresponding to Eqn. (2.17) then,
Amin = U - F q -

S.

The equilibrium criteria may then be expressed as Amin 0. In words: At


equilibrium the mechanical free energy is a minimum. Any displacement
from this state required work.
2.8

Stability Conditions.

To decide whether or not an equilibrium is stable, the inequality


sign in Eqn. (2.16) must be ensured. The conditions for stability may take
different forms depending upon which variables are taken as the
independent variables.
To derive the stability conditions when q and S are taken as the
independent variables consider the terms of second order in small
displacements beginning with the general condition
U -F q-

S > 0.

Choose U = U(q,S), which, because of the identity


dS = dU - Fdq

is a natural choice for the independent variables, and expand U in powers


of the q and S
U =
+

1
2

2
q +2

S+F q
2

U
q S

q S+

+ terms of third order...

(2.18)
The inequality (2.16) then shows that in Eqn. (2.18),
second order terms + third order terms + . . . > 0.
56

Retaining only the second order terms, the criterion of stability is that a
quadratic differential form be positive definite;
2

2
U
U
q S+
2
q S
S

2
q +2

2
S > 0.

(2.19)
If this is to hold true for arbitrary variations in q and S, the coefficients
must satisfy the following:
2

>0 ;

U
q S

> 0.

An alternate approach is seen when u and q are considered to be the


independent variables, a quadratic form in u and q may be found by
using K = U - S so that
K = U-

S-

d
S udu

The terms S u cannot be neglected because in Clausius's inequality,


which is the actual stability condition, the variations are finite, and
therefore, from Eqn. (2.16) the following is obtained:

Expanding in powers of u and q,


K=F q-

d
1
S u+
du
2

2
q +

K
1
q u+
q u
2

and
S u=

1 U
u

2
u +

But
K
U
=
u
u

and
K
= F.
q

57

U
- F q u.
q

2
q + ...

Therefore
2

K
=
u q

F
1
= u
u

U
- F q u.
q

U
.
u

and
2

= -

d 1
du

S-

Then
2

d
du

S u= -

S+
2

( u )2 -

K
u q
u q

and the quadratic form in u and q is


2

( q )2 -

+2

S ( u )2 > 0.

Since
K
q

= F,
u

then
2

F
q

> 0.
u

Other quadratic forms may be derived by using different


independent variables; however, these two quadratic forms will suffice for
this development.
C.

Geometry

2.9

Geometry Required by the Fundamental Laws.

There is nothing that specifies which of the quadratic forms coming


from the stability conditions should be adopted as the metric. Thus the
choice may be based upon simplicity and/or applicability. However, it
becomes obvious that if we choose one of the forms using the velocity as
our metric and then obtain equations of motion, then the equations of
motion will become third order differential equations in time since the
velocity is itself first order and the equations of motion are second order
differential equations.
58

The fact that these equations of motion will become third order
differential equations in time displays a time asymmetry that appears to
correspond to nature. However, third order equations are difficult or
impossible to solve.
To avoid the difficulty of third order equations of motion, suppose we
adopt the quadratic form of Eqn. (2.19) as the metric for our system. Thus
we are adopting a manifold with coordinates of space and mechanical
entropy. This choice is not totally arbitrary because we wish to choose a
metric that will display the metric of Einstein's Special and General
Relativity as subsets of our metric. Looking toward this objective guides us
in the choice of metric.
It now becomes desirable to extend our system beyond the
dimensionality used thus far. Such an extension brings up a question
concerning the integrating factor. With one work term the differential of
the entropy was written as
_E

dS =

= f id

Then if for each dimension the exchange of energy is denoted by dEj, then
_ Ei

dS i =

= f id

where there is no summation intended for fi d i. Since each dSi is a perfect


differential, then the total change in mechanical entropy may be written as
dS =

dS i =

_ Ei

f id

However, the question which arises is whether there exists a single


integrating factor such that
dS =

_E

_ Ei

f id

To see this consider the element of work considered before as


_W =

i
F i dq ; i = 1,..., n.

Since each dUi is in itself a perfect differential, then dU =


_E =

dU i -

i
F i dq =

or
_E =

59

_ Ei .

( dU i - F i dq i )

dUi so that

If the system is total E-conservative in the sense that


_E =

_ E i = 0,

then dE = 0 is a Pfaffian differential equation. This equation is integrable


and has an integrating factor . The integrability is guaranteed by the
Second Law since it is impossible to go from one initial state to any
neighboring state. Then, just as in the one-dimensional case, the perfect
differential follows:
_E

dS =

_ Ei

But since
_E =

dS =

f id

then
fi

Now following the same argument presented in Section 2.2


concerning the composite system, dE = d where is a function of all the
i and the ui. Therefore, since dEi = id i, then
_E =

d +d

Now
d

so that dE =

idEi

or d =

i id i

i
du +

and
d

i
i

It follows that the / ui = 0 and that the ratios i/ are also independent
of the qi. Therefore the 's have the form i = fi and = F( i, i, ..., n)
and also
_S = fd

i
i

f id

The right hand side is a perfect differential and therefore so is the left.
60

Since each i/fi is an integrating factor and /F is also an integrating


factor, it follows that (u1, u2, ..., un) is an integrating factor for the dE as
well as for dEi = idEi. Therefore
dS =

_E

_ Ei

The importance of this question may be seen in terms of the


difficulty that would be created if a universal integrating factor could not
be found. For then each additional work term would require its own
integrating factor to be determined individually.
Thus assured that an overall integrating factor exists, then the
existence of an overall entropy function is guaranteed so that
dS =

_E

dU

Fi

dq

for any i and the quadratic form may be extended to include three spatial
work terms and thus becomes
2

(dS )2 +
2

S q

(dS)( dq ) +

q
, = 1,2,3.

( dq )( dq ) > 0 ;

Adopting this quadratic form as the metric of a general system whose


thermodynamic variables are held fixed, we may then write this metric as
(dS )2 = hij dqi dq j ; i, j = 0,1,2,3,

(2.20)
where the summation convention is used and
2

hij =

U
q

with
q0 =
S/F0, the scaled mechanical entropy for dimensional
correctness.
Thus, the stability conditions provide a metric in the
four-dimensional manifold of space-mechanical entropy. However, the
existing relativistic theories are theories in a space-time manifold.
Therefore, if these theories are to be contained within the Dynamic Theory,
then the space-time manifold must be found within the Dynamic Theory.
The arc length s in the space-mechanical entropy manifold may be
parameterized by choosing ds = u0dt = cdt, where u0 = c is the unique
velocity appearing in the integrating factor of the second postulate. There
are two reasons for choosing the unique velocity. First, it is the only welldefined velocity we have thus far. Secondly, we may look ahead to the
61

metric of the Special Theory of Relativity. The metric may now be written
as
2
i
j
2
c (dt ) = hij dq dq ; i, j = 0,1,2,3.

(2.21)
Now suppose the systems considered are restricted to only
E-conservative systems. Then the principle of increasing mechanical
entropy may be imposed in the form of the variational principle
( dq0 )2 = 0.

In order to use this variation principle, Eqn. (2.21) may be expanded,


solved for (dq0) and squared to arrive at the quadratic form
( dq0 )2 =

1
h00

2
2
c (dt ) + 2 h0 Adtdq - h dq dq ,

(2.22)
where
A=

h0
u +_
h00

with u = dq /dt.
By defining x0 = ct, x = q ;
written as
( dq0 )2 =

h
c
h0
+
( u )2
u u +
h00 h00
h00

= 1, 2, 3, then Eqn. (2.22) may be

1
g dxi dx j ; i, j = 0,1,2,3
f ij

(2.23)
where f = h00. This metric obviously reduces, in the Euclidean limit of
constant coefficients, to the metric of Minkowski's space-time manifold of
Special Relativity. It is interesting to note that in the metric of Eqn. (2.22)
the difference in the sign on the time and space elements of the metric
come from stability conditions given in terms of space and mechanical
entropy while the variational principle was taken to be the Entropy
Principle. In this fashion the Second Law guarantees the limiting aspect
found in Einstein's Special Theory of Relativity.
In his General Theory of Relativity, Einstein assumed the space-time
manifold to be Riemannian. However, this assumption involves the a priori
assumption that the scalar product be invariant. This assumption was
later questioned by Weyl in his generalization of geometry. From the
viewpoint that the adopted postulates of the Dynamic Theory should
62

contain the other theories it then becomes desirable to determine whether


or not these postulates specify the geometry of the (dq0)2 space-time
manifold. More particularly do the adopted postulates lead to a geometry
that includes the geometry of current theories? To arrive at a more general
geometry would not be a limitation for it would certainly include the
others.
Recalling Eqn. (2.23), we can define
( dq0 )2 =

1
g dxi dx j
f ij

1
(d
f

g ij dxi dx j .

(2.24)
Now the Second Law guarantees the existence of the function mechanical
entropy and that dq0 be a perfect differential; therefore
0
dq = q0 i dxi ,

(2.25)
where q0i =

q0/ xi. Then the exactness of dq0 is stated by


q

0
i| j

- q0 j|i = 0.

(2.26)
By defining the parallel displacement of a vector to be
d

is

dx

s
v

(2.27)
and using Eqns. (2.26) and (2.27) it may be seen that the connections
must be symmetrical, or
v

ik

ki

(2.28)
This result should not be taken to mean that only symmetric
connections need to be considered. Rather it means that given the ij's that
maximize (dq0)2, then the connections are symmetrical. However, since a
variational principle must be used to determine the ij 's, then both
symmetric and antisymmetrical connections will have to be considered.
In Weyl's generalization of geometry he found it necessary to
assume the symmetry of the connections. He proved a theorem showing
that the symmetry of the connections guaranteed the existence of a local
63

Euclidean limiting manifold and used this theorem in support of the


symmetry assumption. Here we find that the Second Law requires that the
connections formed by the solution coefficients must be symmetrical thus
guaranteeing, through Weyl's theorem, the existence of a local Euclidean
geometry within the Dynamic Theory.
Suppose now we consider whether the order of differentiating the
change in entropy makes any difference. This means that we must use
symmetric connections since the actual change in entropy will be
determined by the metric coefficients that generate a maximum. Therefore,
consider the difference
2

_( dq0 )2 =

( dq0 )2
x

Since (dq0)2 = q0iq0jdxidxj from Eqn. (2.25), using Eqn. (2.24) we find q0iq0j
= gij. Then
( dq0 )2
x

= [ q0 j|k q0 i + q0 i|k q0 j ] dxi dx j + (d q0 k )2 .

Thus
2

( dq0 )2
x

= [ q0 j|k|l q0 i + q0 j|k q0 i|l + q0 i|k|l q0 j + q0 i|k q0 j|l ] dxi dx j


+ 2 q0 l|k q0 l + 2 q0 k|l q0 k .

Likewise
2

( dq0 )2
x

= [ q0 j|l|k q0 i + q0 j|l q0 i|k + q0 i|l|k q0 j + q0 i|l q0 j|k ] dxi dx j


+ 2 q0 k|l q0 k + q0 l|k q0 l .

Therefore the difference must be


_( dq0 )2 = [( q0 j|k|l - q0 j|l|k ) q0 i + ( q0 i|k|l - q0 i|l|k ) q0 j ] dxi dx j .

Using the definition Eqn. (2.27) we see that


d q0 i =
q

= q0 r

i|k

0
k |i

Now

64

is

s 0
dx q r ,
r

= q0 r

ik

, also
r

ki

i|k |l

= q0 r|l
= q0 s

= q0 r

ik

ik
l
r

+ q0 r

ik

sk

ik
r

+ q0 r

ik

rl

[ q0 r

ik
l

ik
l

Similarly
q

0
i|l|k

= q0 r

sk

il

il
k

Therefore
0
0
0
0 0
q j [ q i|k|l - q i|l|k ] = q i q r

ik
l

il
k

sl

ik

Then defining the vector curvature as


r

ilk

ik
l

the difference may be written as


_( dq0 )2 = [ q0 j q0 r R r ilk + q0 i q0 r R r jlk ] dxi dx j .

However, recall that q0iq0j = ij; then


_( dq0 )2 = [ g jr R r ilk + g ir R r jlk ] dxi dx j .

But gri =gir and Rijkl = girRrjkl, so that


_( dq0 )2 = [ R jilk + Rijlk ] dxi dx j .

So the difference will vanish if Rjilk = -Rijlk. Now since


( dq0 )2 = q0 i q0 j dxi dx j = g ij dxi dx j ,

differentiation will result in


d( dq0 )2 = d( q0 i q0 j dxi dx j ) = d( g ij dxi dx j )

or

65

sk

il

d q0 i q0 j dxi dx j + q0 i dq0 j dxi dx j + q0 i q0 j d( dxi dx j )


= dg ij dxi dx j + g ij d( dxi dx j )

which can be written as


r

is

s 0
i
j
0
0
dx q r q j dx dx + q i

js

s 0
i
j
i
j
0 0
dx q r dx dx + q i q j d( dx dx )

= dg ij dxi dx j + g ij d( dxi dx j ).

But gij = q0iq0j. Therefore


r

is

s
dx g rj +

js

s
dx g ir = dg ij

or
g rj

is

+ g ri

js

g ij

and
jis

ijs

g ij
x

(2.30)
Now interchange jis to sij to get
+

sij

jsi

g is
x

(2.31)
Then interchange jis to isj so that
isj

sij

g si
x

(2.32)
Add Eqns. (2.31) and (2.32) and subtract Eqn. (2.30).
sij

jsi

or

66

isj

sij

jis

sij

g si

1
2

g sj

g ij

(2.33)
and
r

ij

rs

= g rs

sij

g si

g
2

g sj

g ij

Now by using the symmetry of gij it can be shown that


R jilk = - Rijlk

and therefore (dq0)2 = 0. This is the necessary and sufficient condition


that the differential entropy change may be transferred from an initial
point to all points of the space in a manner that is independent of the path.
The distinguishing feature of Riemannian geometry is the invariance
of the scalar product under a vector transplantation.
Therefore to
0
2
determine whether the (dq ) space is a Riemannian space, consider the
vector i and i. Now since i = gij j and
d

is

dx

is

s
dx g rk

g ij

s
dx + g ij d

then
g ij d

Or, since gijgij =

i
i

is

s
dx g rk

g ij

s
dx .

= 1 and
g ij
x

jis

ijs

then

Thus the change in the covariant and the contravariant vectors are given
by

Now consider the change in the scalar product


67

i.

Then

d(
=

i
is
r

dx
is

)= d
i

s
r

dx

+ i (i

s
r

rs

dx

rs

dx

s
r

s
i

Renaming the indices in the second term yields


d(

i
i

)=(

i
is

i
is

) dx s .

Thus the geometry of the (dq0)2 manifold is Riemannian.


Next consider the question of what is the geometry of the (d )2
space? Equation (2.24) shows that we may write (d )2 = f(dq0)2, which is
reminiscent of Weyl's generalized geometry. Further we have
g

ij

= f g ij .

Then in the sigma space an arbitrary vector


by the self-scalar product
i

l = || | |2 = g

ij

would have a length given

(2.34)
where l is the length of the vector in the entropy space.
If we differentiate Eqn. (2.34), we have
2l d l = l

f
x

i
dx + 2fldl.

However, in the entropy space the length of the vector is unchanged under
parallel displacement so that
dl =

1 f
1 lnf i
i l
=
dx
dx l .
2 xi
f
2 xi

(2.35)
Comparing Eqn. (2.35) with the definition of the parallel displacement of a
vector, Eqn. (2.27), we find that
i

lnf
x

plays a role similar to that of the connections ijk in the definition of


parallel displacement of a vector. Therefore we shall define the change in
the length of a vector under displacement to be

68

dl = (

i
dx )l.

(2.36)
This is the same definition Weyl made in his generalization of
geometry. However, there is a difference in the way it was obtained. Weyl
chose this definition in analogy with the connections and the definition
then led to the second more general metric. In this theory the fundamental
laws lead us to two metrics and Eqn. (2.35) for the change in the length of
a vector under displacement. Therefore, we have no choice.
Thus within the Dynamic Theory Eqn. (2.35) is a derived equation
and Eqn. (2.36) only renames the logarithmic derivative.
Using Eqn. (2.36) we may obtain, in general,
2
2
dl = 2 l (

= g ij|k

k
dx + g ij

i
dx ) = d( g ij
l

lk

k
dx + g ij

j
j

)
i

lk

k
dx .

Renaming the various summation indices, rearranging terms, and using


the length of a vector, we obtain
[ g ij|k + g lj

ik

+ g il

jk

Since this must hold for arbitrary choice of


( g ij|k - 2 g ij

) + g lj

k
dx = 2 g ij

ik

k
dx .

and dxk, we conclude that

+ g li

jk

= 0.

This is the same system of linear equations for the connections ijk as Eqn.
(2.30) except that the inhomogeneous term ijk has now to be replaced by gijk
- 2gij k. Therefore the same linear algebra as before leads to
i

jk

= -

i
jk

+ g li [ g lj

+ g lk

(2.37)
where (ijk) is the usual Christoffel symbol of the second kind.
Now, since the entropy space is Riemannian, then in the entropy
space we have 'i=0 and 'ijk = -(ijk) and the length l of a vector is unchanged
under parallel displacement. However, the same displacement law in the
sigma space, with metric gij, leads to the relation

69

dl = + _d g

= l

= + _d fg ij

ij

dx

x
1 lnf k
= +_
dx l .
2 xk

(2.38)
Thus (1/2)( lnf/ xk) plays the role of k in Eqn. (2.36). It follows then
that the ordinary connections -(ijk) constructed from ij are equal to the more
general connections ijk constructed according to Eqn. (2.37) from gij and k
= (1/2)( lnf/ xk): This can also be seen by direct computation from Eqn.
(2.36)
(2.39)
g ij = fg ij .
and
i

jk

jk

(2.40)
We may interpret the change of metric from ij to gij by Eqn. (2.40) as a
change of scale for the length at every point of the Riemannian manifold by
the variable gauge factor f.
This transformation is called a gauge
transformation, and k is called a gauge vector field.
The generalized geometry thus separates the problem of measurement of angles from that of measurement of length. For instance, the
angle between the two vectors i and i at a given point of the space is
measured by the ratio
i
i

|| || || ||

g ij
( g ij

)( g ij

This ratio does not change under the gauge transformation Eqn. (2.40).
The gauge transformation is therefore an angle-preserving, or conformal,
change of metric. On the other hand, the length of vectors will change
under Eqn. (2.40) according to Eqn. (2.35). Thus the metric tensor ij
determines angles, while one needs also the gauge vector k to measure
length.
Considering the sigma space, which is characterized by the tensor
field ij and gauge vector k. The same argument as before shows that we
may, without changing the intrinsic geometric properties of vector fields,
replace the geometric quantities by use of a scalar field f as follows:

70

ij

= fg ij ,

1 lnf
,
2 xk

jk

jk

(2.41)
That is, in the new metric, vectors will have the same law of affine
transplantation and the angle between different vectors at the same point
of the manifold will be preserved, but the local lengths of a vector will be
changed according to
2
2
l = fl .

Thus the general Weyl geometry of the sigma space admits also a
conformal gauge transformation.
D. Mechanical Systems Near Equilibrium
2.10 Special Relativistic and Classical Mechanics
Classical mechanics describes the motion of a system, which could
be a particle, for which the energy of the system is a constant. The
equations of motion yield trajectories resulting from the action of forces;
they may also be obtained from the Principle of Least Action. When the
action integral is treated as a variational problem with variable end points,
the method of Lagrangian multipliers yields the same equations as does
Hamilton's Principle. However, if the variational problem is transformed to
a new space in which the new variational problem has fixed end points,
then the metric for this space is displayed, and the equations of motion are
geodesics in this space.
In classical mechanics the Principle of Least Action as formulated by
Lagrange has the integral form
A=

P2
P1

mv d s .

(2.42)
In curvilinear coordinates the integral assumes the form
A=

P2
P1

mg

dx
dx =
dt

t( P 2 )
t( P1 )

mg

where , = 1, 2, 3.
Defining
T =

m
g
2

71

dx dx
,
dt dt

dx dx
dt,
dt dt

the integral becomes


A=

t( P2 )
t( P1 )

2Tdt.

Then the principle of least action may be stated as:


Of all curves C' passing through P1 and P in the neighborhood
of the trajectory C, which are traversed at a rate such that, for
each C', for every value of t, T+V=F, that one for which the
action integral A is stationary is the trajectory of the particle.
The transformation of variables may be carried out to display the
metric
(ds )2 = h dx dx

(2.43)
where h =2m(E0-V)g . Here different particles in the same field and with
different energies E0 would appear to have different geometries, a situation
which has been previously taken to be impossible and therefore precluded
the geometrization of dynamics(see page 6 of ref. 46). However, in view of
Weyl's generalization of geometry, treating the variational problem in the
Principle of Least Action as transformed to a new space in which the variational problem has fixed end points, in effect, is a transformation into a
space with Weyl geometry where the gauge function is 2m(E0-V). Thus
changing the energies does not change the geometry since it will still be a
Weyl space.
Suppose now that the concepts of classical mechanics are compared
with the concepts from the point of view of the Dynamic Theory. The
energy of the system in classical mechanics is a constant of the motion and
therefore the change in kinetic energy is the negative of the change in
potential energy, which may be written as
dH = dT + dV = 0.

However, for classically conservative forces dH is a perfect differential.


Therefore for this system with only one work term the force is a function of
position only.
This suggests the association of the classical energy of the system,
H, with the system energy, U, which is also a perfect differential. Now if
the system is isolated, or E-conservative, then
0 = _ E = d U - F dq.

But if dU=dH=0 then F must be zero. This points out an important


difference between classical physics and the Dynamic Theory. A classically
72

conservative system is one for which the system's energy is a constant of


the motion. However, the E-conservative system, within the Dynamic
Theory, is one for which dE = 0. Thus an E-conservative system which is
also conservative in the classical sense must have no forces F which may
depend upon velocity as well as position but may have forces which arise
from -( U/ q) = F and must be functions of positions only.
Suppose we now turn our attention to the mechanics of Special
Relativity. In the Special Theory of Relativity Einstein sought to put
Newtonian mechanics into a form that would leave the speed of light
invariant. The resulting dynamics exhibits the notion of a unique velocity
in a similar sense to the previously defined absolute velocity.
Within the Dynamic Theory we may display the appearance of the
Special Theory's foundations by using the generalized entropy principle
rather than being required to assume the existence of Newton's equations
of motion on an a priori basis.
Newtonian mechanics is displayed in its simplest form for particles,
so we shall make the restrictive assumption that the mass density, 0, such
that
0

d(vol) = m0

We will also assume that the gij are constants, thus


(d

and g =

2
2
) = c 2 (dt ) - g dx dx ,

= 1,2,3,

. Our variational problem depends upon the integral


S=

S
0

dS =

S
0

(dS )2 =

mq0
0

0 2
2
m0 ( dq ) =

2
1

2
j k
m0 f g jk u u d ,

where we have used the definition


j
u =

dx
.
d

Because we have assumed that the mass, m0, is a constant we can write
our integral as
0
q =

S
m0

2
1

f g jk u j u k d .

We can make a change of variables by letting


ds =

g jk u j u k d ,

(2.44)
so that the integral becomes
73

0
q =

S2
S1

f dS.

We can now define a new function


f
2
m0 c ,
2

T =

and then consider a further change in variables such that

dS = c f , d =

2T
m0

d .

If we substitute this new variable into our integral we find


0
m0 c q =

( P2 )
( P1 )

2T d ,

(2.45)
with the auxiliary selection that 2T - m0c2f = 0.
The problem of determining geodesics has now been converted into a
statement of the principle of maximum generalized entropy:
Of all curves C', passing through P1 and P2 in the
neighborhood of the trajectory C, which are traversed at a rate
such that, for each C', and for every value of , 2T = m0c2f, that
one for which the generalized entropy integral, q0, is maximum
(stationary) is the trajectory of the particle.
reads

When stated in the form of a variational equation, this principle


( P2 )
( P1 )

2T d = 0,

with the auxiliary condition 2T-m0c2f=0, on C'.


The dynamics is unaffected by the addition of a constant to the
gauge function; therefore, let
V = h - m0 c 2

f
,
2

where h is a constant. The auxiliary condition now reads


T + V - h = 0, on C .

74

We can solve this variational problem by making use of the


Lagrangian multiplier method for a problem with constraints.
We
construct a function G=2T+ , where =T+V-h=0, and determine the
solution of the system of equations
G
x

d
d

G
u

= 0,

j = 0,1,2,

, n, with

T + V - h = 0.

(2.45)
This system has a solution for which =-1Sokol, and it follows that the
trajectory C is determined by the solution of the system
d
dt

T
u

T
x

= -

j = 0,1,2,

, n.

(2.46)
We assumed that the gjk were constants; therefore, if we make the
definition
Fj = -

V
x

then the equations in Eqn. (2.46) become


d
2
k
m0 c g jk u = F j ,
d

because d =cd and =m0gjkujuk. If we multiply these equations by gli and


sum them, we obtain
d
d

lj
l
l
m0 u = g F j = F , l = 0,1,2,

, n.

(2.47)
From the metric with constant coefficients we get
d
=
dt

2
c - g

x& x& ,

, = 1,2,3,

or
d
=
dt

2
2
c -v .

(2.48)
75

Substituting Eqn. 2.48 into Eqn. 2.47 we find that


F

l
d
dt
2 dx
m0 c
dt
d
d

d
m0 c 2 u l
d

d
dx l dt
m0 c 2
dt d
v 2 dt

d
v 2 dt

c2
c

m0 c 2
c

vl

where vl=dxl/dt. Thus we have


1

l
F =

1-

d
dt

m0 v
1-

(2.49)

where =v/c. Now Eqn. (2.49) can be rewritten as


2

m0 v l

d
dt

.
(2.50)

Because =0 in a local coordinate system x,


l
F = m0

d x
= m0 a l ,
2
dt

(2.51)
where, in a local reference frame x, al=(1/c2)(d2xl/dt2). In Eqn. (2.51) we
have the form of Newton's second law in classical mechanics.
We may rewrite Eqn. (2.50) in the form
1-

l
f =

d
dt

m0 v
1-

l
2

(2.52)
These equations are the equations of motion of the Special Theory of
Relativity and come from the geodesic equations of the variational problem,
76

in Eqn. (2.42), based upon the generalized entropy principle with the
restrictive assumptions that the mass density, , be a constant and that
the metric coefficients, gjk, are independent of the mass distribution. Thus,
we have shown that the special relativistic equations of motion and the
Newtonian equations of motion are required by the generalized entropy
principle, but that they represent a limited subset of the entropy principle.
2.11

Energy Concepts

Newtonian and relativistic mechanics talks of potential and kinetic


energy while classical thermodynamics, which forms the basis of the
Dynamic Theory, contains concepts with units of energy such as entropy,
enthalpy, and free energy. We may use these common fundamental
principles within the Dynamic Theory to explore how the mechanical
energy concepts fit among the general thermodynamic energy functions. It
seems that this will be of more than a little benefit when trying to keep all
of the energy-based concepts in proper perspective.
First, let us recall the First Law, with n=1,
_E = dU - Fdx,

(2.53)
whereas the differential change in the generalized entropy is
dS =

dU

Fdx

(2.54)
where dS is a total differential. If we suppose that the system energy, U,
may be a function of position, x, and the velocity, v=dx/dt, then we may
write
dS =

1 U
1 U
F
dv +
dx - dx.
v
x

Because dS is a total differential, then


1
v

U
-F
x

This requires that

77

1 U
.
v

U
x v

(2.55)
because = (v) from the second law. Further, dU is a total differential so
Eqn. (2.55) becomes
F
v
U
-F
x

= - v = function of velocity only.

Now consider the functional form of the force from the equations of motion
in Eqn. (2.52),
F =

1-

F(x)

F(x),

(2.56)
where F(x) is strictly a function of x because it came from the gauge
function. Then
F
d
=
F (x).
v
dv

Thus, Eqn. (2.55) may be written as


d
F ( x )
dv
U
F ( x )
x

d
dv .

(2.57)
In order to satisfy Eqn. (2.57) we find = and U=U(v).
By substituting these results into the differential expression for the
entropy, Eqn. (2.54) we find

dS =

dU
dv
dv
1-

- F(x) dx,

(2.58)

78

which is a perfect differential whereby we have found that U is strictly a


function of velocity.
Now consider the First Law for an isolated system, or
_E = 0 = dU - Fdx,

but, using Eqn. (2.56) this may be written as


dU =

1-

F(x) dx.

P
P0

F dx.

Then by integrating we find that


U -U0 =

This is Einstein's energy integral which, because of the equations of


motion, becomes
m0 c

U =

2
2

1-

+ constant.

(2.59)
In his Special Theory of Relativity Einstein interpreted the right-hand
side of Eqn. (2.59) to be the kinetic energy; therefore, he chose the
integration constant to be -m0c2 in order that T=0 when v=0. Here, Eqn.
(2.59) is the energy of the system and, therefore, will not be zero when v=0.
Thus, the constant of integration should be taken as zero, giving the
energy by
m0 c

U =

1-

2
2

(2.60)
If we differentiate Eqn. (2.60) with respect to the velocity, we find
U
v

1
2v
m c
2
c

m v

(1

(1 B )

Substituting this result into Eqn. (2.58), the change in entropy


becomes
dS =

m0 vdv
- F(x)dx.
(1 - 2 )2

This expression may now be integrated because


79

m v

S
1
m c
2

(1

d(
(1

)
)

dv

2
1

F ( x )dx

V ( x ),

and
1
2
m0 c
2
S =
+ V(x) + constant.
(1 - 2 )

By setting the constant of integration at 1/2 m0c2, we get


1
1
2
2
2
m0 c
m0 v
2
2
S =
+ V(x) or S =
+ V(x).
(1 - 2 )
(1 - 2 )

(2.61)
Thus, the generalized entropy for a purely mechanical system has
two parts. One, depending entirely upon the velocity, and which we may
call kinetic entropy, is given by
1
2
m0 v
2
T =
.
(1 - 2 )

(2.62)
The second term in the mechanical entropy is a function of position only
and may be called the entropy potential, V(x).
We may look at the kinetic entropy differently if we go back to the
variable changes during the presentation of the maximum entropy
principle, because there we had
ds =

but d = cd

g jk u j u k d

2t
m0

d ,

1 - v 2 / c 2 199 therefore,
T =

j
k
1
1
du du
2
j k
2
m0 c g jk u u = m0 c g jk
2
2
d d

1
dt
2
j k
m0 c g jk x& x&
2
d

But, by Eqn. (2.48) this becomes


80

1
2
m0 v
2
T =
,
(1 - 2 )

which is the kinetic entropy of Eqn. (2.62). Thus, we find that it is the
mechanical entropy, S, that must have a constant value along any
trajectory for an isolated system, because
T +V - h = 0

for the trajectory, and therefore, S=h=constant.


Thus we have established the following for the trajectories of an
isolated system:
Mechanical Entropy: S = h = constant
m0 c 2
Kinetic Entropy : T
2
(1
)
Potential Entropy : V ( x )
F ( x )
m0 c 2

Energy of the system : U

Work done by the system : W

x2
x1

Kinetic energy of the system : T


Force : F j

or 0

m0 vdv
(1

3
2

m0 c 2

1
1

F j ( x )

Re st energy of the system : U ( v


First Law : d E

F j dx j

m0 c 2

F j dx j ,

dU
1

0)

F j dx j

2.12 Non-Isolated System


Thus far we have consistently required the system to be isolated.
Obviously there are a large number of physical phenomena for which this
restriction may not be used, even as an approximation. Therefore,
relaxation of this restriction should provide description of a large and
important class of systems. The remainder of this book involves the
investigation of the predictions of the Dynamic Theory assuming the
system is isolated. This may give the implication the non-isolated system
is less important or less interesting. This is not the intention of the
81

presentation. Rather, the presentation is aimed at displaying the fact that


existing theories are subsets of the Dynamic Theory. In order to do this we
must stay with the assumption of the isolated system.
One of the benefits of the Dynamic Theory is the capability of using
procedures currently used in one branch of physics in another where prior
to the unification displayed here would have been thought impossible. A
system in which this procedure should produce significant results is a
nonequilibrium thermodynamic system. Thermodynamics tells us that we
must minimize the free energy, but the ability to use this as a variational
principle to obtain equations of motion is a procedure that the Dynamic
Theory now makes possible for this thermodynamic system.
E. Quantum Mechanics
2.13 Quantum Mechanics Derived
In 1927 F. London derived quantum principles from Weyl's
geometry.(47) However, the results of his work made it difficult to define
length as a real number and because of this Weyl later interpreted the
mathematical formalism of his unified theory as connected with
transplanting a state vector of a quantum theoretical system.
Suppose that we consider an isolated, or E-conservative, system so
that dE = 0. Then, because of the Second Law dq0 0 which is the principle
of increasing mechanical entropy. Then certainly (dq0)2 0 and also, since
(dq0)2=f(d )2, then f(d )2 0. However, if f<0, then (d )2<0 since it is the
product that must remain greater than, or equal to, zero. In this case
0
dq =

-f

- (d

2
) .

But
d(d ) =

k
dx (d )

and
d(d )
=
(d )

k
dx ,

which implies that the element of arc (d ) is given by


(d ) = (d

)0 e

k dx

where (d )0 is some initial value of the element of arc.


Now suppose an equilibrium, or reversible, state is desired so that
dq0=0. Thus, the desired condition is a null trajectory of the (dq0)2
82

manifold. Then, if f 0, the desired condition is also a null trajectory of the


(d )2 manifold. This implies that d(d )=0 or (d )=(d )0, so that
k dx

= 1,

which is satisfied only if


k
dx = 2 iN,

(2.63)
where N is an integer. This is the quantum condition London introduced.
To illustrate how this condition arises from the Dynamic Theory's
approach, suppose a description of a hydrogen atom is desired. A
hydrogen atom is in a stable condition and, if isolated, satisfies the
conditions dE = 0 and dq0=0. These conditions along with f 0 establish
the quantization of the integral in Eqn. (2.63). To show how the Dynamic
Theory removes from London`s work the difficulty of defining length as a
real number, consider an elementary presentation of London's. Suppose
the field of a proton to be given by
1
0

0; i

0.

Equality of forces for the simple case of circular motion requires that
2

mv
e
= 2.
r
r

Thus the period is given by T=2 r/v and the velocity by


e

v=

mr

Now
k

k
dx =

cT = 2 iN,

so that
1

cT
=
r

c2

mr
= 2 iN.
e

Solving for the radius shows that the allowed radii are
r =

- N 2 e2
.
( 1 )2 m c 2

By choosing
83

2 i e2
hc

i
137

i ,

where h is Planck's constant, then the possible radii become


2

r =

N h
,
4 2 e2 m

which are the Bohr radii.


The imaginary 1 presented the difficulty, in London's work, of
defining length as a real number. In the Dynamic Theory real distance, or
length, may be defined, and properly should be, in the (dq0) manifold.
Recalling that the definition of the potentials is
k

= +_

ln f
x

1
2

it may easily be seen that if f<0, then k becomes imaginary as does the
length of arc in the (d )2 manifold since the length of arc is given by
=

(d

2
) .

However, the arc length in the (dq0)2 manifold is real since dq0 0 by the
Second Law.
It should be noted that the conditions for quantization are not
restricted to dE=0, dq0=0, and f<0 as used here. Any set of conditions
which results in the final element of arc (d ) being equal to the initial
element of arc (d )0 results in quantum conditions. It is particularly
significant to note that the quantization involves only forces, which may be
described in terms of the "distance curvature" and does not involve forces
describable by a vector curvature. Thus interpreting the gauge potentials
be electromagnetic potentials provides quantum effects for
k to
electromagnetic forces.
Here, again, is a distinction between curved and Euclidean
manifolds, though here it appears slightly different. The Dynamic Theory
requires a quantization. However, this quantization depends upon the
existence of a gauge function and appropriate restrictive conditions. Thus
a curved space may exhibit quantum effects but only if the curvature is
accompanied by a gauge function or a distance curvature.
Thus the Dynamic Theory, through London's quantization, not only
supports the contention that "God does not play with dice all the time" but,
further, may supply the answer to the question, "What is waving in the
wave function?" London showed that the wave function is directly related
to the element of the arc length in the sigma manifold. Therefore the
"waving" is the tendency of this element of arc length to increase and
84

decrease around a closed path. Using the calculus of complex variables,


the quantum number becomes the order, or multiplicity of the zero of (d ).
This may also be stated in terms of null trajectories. Einstein's null
trajectory was the path light travels and remains so here. However, this is
the zeroth order null trajectory. The remaining null trajectories for the
complex arc length are given, as London showed, by the equations of
Quantum Mechanics.
2.14 On The Derivation of Thermodynamics from Statistical Mechanics
It is commonly believed that one can "derive" thermodynamics from
a variety of force laws using the techniques of statistical mechanics. This
belief is not supported when one considers the development of statistical
thermodynamics.
For instance, in order to talk of a statistical
temperature, , one must start by assuming Newtonian physics (this
constitutes three fundamental assumptions). Given Newtonian physics
one can talk of an energy distribution, canonical ensembles and statistical
temperature; however, one must make an additional fundamental assumption (the Equipartition Law) before the statistical heat capacities may
be obtained.
In order to obtain thermodynamics we need two more assumptions.
To display the assumption necessary for the first law of thermodynamics
let me quote from page 85 of "Basic Theories of Physics: Heat and Quanta"
by Peter G. Bergmann. "The difference between the heat transferred to the
system and the work performed by it,
_Q - _W = C d +

n
i=1

( J i - Y i )d Ri ,

(2.64)
is, according to our previous discussions, the increase in u. But in a
systematically thermodynamic approach (that is, using only macroscopic
observations and concepts), we get the differential expression, Eqn. (2.64)
without reference to u. From that point of view, to claim that this
expression is an exact differential is a logically new assertion; and this
assertion constitutes the First Law of Thermodynamics."
The assumptions of statistical thermodynamics allow us to derive
the differential of the heat exchanged in the form
_Q = d ,

(2.65)
where
is the statistical temperature and d
Further, it may be shown that

85

is the statistical entropy.

_Q

(2.66)
By comparing Eqn. (2.66) with the classical thermodynamic statement
_Q = TdS,

(2.67)
and
dS

_Q
,
T

(2.68)
we find that the statistical expression, Eqn. (2.66), is analogous to the
classical expression, Eqn. (2.68), for the second law of thermodynamics.
Also we may equate Eqns. (2.65) and (2.67) to obtain
d

= TdS,

(2.69)
and Eqns. (2.66) and (2.68) are simultaneously satisfied provided that
/T>0. In statistical thermodynamics it is asserted that /T=kB, where kB
is Boltzman's constant. Once this assertion is made then we have
kBd =dS, hence S=kB . However, there is no logical necessity that the ratio
/T be a constant from the statistical approach, and only if it is a constant
can we have a one-to-one correspondence between the statistical entropy
and the classical thermodynamic entropy.
The misconception that classical thermodynamics may be derived
from Newtonian mechanics without the necessity of making additional
assumptions is further entrenched by authors, such as Kittel, who in his
text Thermal Physics says the following on p. 49, "We show in Chapter 8
that is proportional to the conventional absolute temperature which is
measured in degrees Kelvin"; (This implies a logical necessity) On p. 427
the author states, "By analogy with the relation dQ = d we `assume' that
the Kelvin temperature T has the property dQ = kBTd
for a reversible
process; here kB is a constant to be determined and is the entropy." (The
implied logical necessity is reduced to an assumption.)
We are so familiar with Newtonian mechanics and its basic validity
that it is difficult for us to consider that it might be derivable from some
other physical concept and its associated fundamental assumptions.
Further, classical thermodynamics, even before statistical mechanics gave
86

rise to the distinction between microscopic and macroscopic views, never


talked of anything resembling equations of motion. Added to these factors
is the somewhat long logical progression from the adopted laws of the
Dynamic Theory to Newtonian mechanics.
On the one hand the First and Second laws adopted by the Dynamic
Theory give rise to a generalized entropy principle that requires that any
dynamics for an isolated system must occur so as to seek a maximum of
the generalized entropy. Thus we have a variational principle based upon
maximizing the entropy. On the other hand the laws produce, through the
stability conditions, a metric to be used in this variational principle in
which the type of geometry is specified and need not be assumed as in
Newtonian and relativistic mechanics. The Euler equations resulting from
this variational principle taken in the non-relativistic, Euclidean,
three-dimensional limit, for particles become
2

d x
= Fi ,
2
dt

or inertial mass, m, times the acceleration, d2xi/dt2, must equal the force,
Fi. Thus the adopted laws, through restrictive assumptions, do lead to
Newton's Second Law. Newton's First Law comes from considering the
motion in the absence of any force. To arrive at Newton's Third Law one
must show that all of the forces, allowed by the adopted laws of the
Dynamic Theory, must be symmetrical. A violation of this symmetry of
forces that has recently been found will be shown in Section 4.11 for forces
within the nucleus; and, therefore, the Dynamic Theory does not require
Newton's Third Law to hold within the nucleus, but does for atomic forces
and macroscopic matter.
F. Summary
2.15 Summary of new theoretical fundamentals
When this investigation was initiated, it was concluded that
Einstein's postulate of the constancy of the velocity of light could not be
adopted since it was felt that experimental evidence in electromagnetism
alone did not justify applying it as a limiting velocity to all types of forces.
However, we find that Einstein's postulate is required by the Dynamic
Theory which approaches physical phenomena from a different way. The
new viewpoint indeed supports Einstein's every contention with regards to
Special Relativity and his uneasiness concerning quantization. Further,
the Dynamic Theory supports Einstein in such a way that it seems only
the early successes of his theories kept Einstein himself from coming to the
same realization.
This is, of course, speculation, but it was Einstein who returned to
very fundamental concepts in order to establish a basis for his relativity
87

theory. He was also known to be aware of the tremendous strength of


classical thermodynamics since he wrote, "A theory is the more impressive
the greater the simplicity of its premises are, the more different kinds of
things it relates, and the more extended is its area of applicability.
Therefore the deep impression that classical thermodynamics made upon
me. It is the only physical theory of universal content concerning which I
am convinced that, within the framework of applicability of its basic
concepts, it will never be overthrown." Thus it seems only the fact that
Caratheodory's statement of the second law, which is the key to the
development of the Dynamic Theory, did not make its appearance before
the relativistic theory had achieved such stupendous successes kept
Einstein from eventually investigating its possible extended application.
The key points in the development of the Dynamic Theory seem to
be the recognition of the generality of the thermodynamic laws and their
independence upon the number or type of variables considered and the
recognition that the quadratic forms associated with the stability
conditions from natural metrics leading to a geometrical description of the
dynamics of the system independent of the variables used in the
description.
There are numerous conclusions and implications that could be
reiterated here; however, only a few of the seemingly more significant ones
will be discussed. The first one is the existence of an integrating factor for
any system describable by the First Law, particularly an integrating factor
independent of the type of force considered. It is this fact that ultimately
leads to a unique limiting velocity for all forces. However, in speaking of
the absolute velocity for mechanical systems, care must be taken to point
out that, as far as the three laws are concerned, it does not represent an
absolute barrier. Rather the laws only state that, for a mechanical system
with only three work terms representing the work done by three spatial
forces the absolute velocity represents an upper and lower limit. Thus
solutions with velocities greater than the speed of light are also allowed.
However, so long as the system is subjected to only these three forces, then
its velocity may never cross this barrier. This absolute barrier effect may
be expected to change if another force term representing an additional
dimension is found necessary.
The reduction in the number of fundamental laws or postulates is
significant. This together with the unifying effect of the three laws
promises to simplify the study of physical phenomena by founding the
entire realm of physics upon a common set of conceptualizations.
In this chapter it was shown that the Special Theory of Relativity was
a special case obtained from the fundamental laws adopted by the
Dynamic Theory. Einsteins postulate concerning the speed of light was an
immediate result of the Second Law. Further, it was shown that Quantum
Mechanics resulted when stable isentropic states were required. This
displays a different relationship between these theories than has previously
88

been envisioned. Here there is no question concerning one being more


fundamental than the other. The determining factor is not which is more
fundamental, but what restrictions are placed upon the system. For
example, if the system is in an "isentropic state" that state is to be
determined by the equations of Quantum Mechanics. If the system is not
required to be isentropic or otherwise restricted so that the entropy must
return to its original value after completing a loop the quantum Mechanics
does not describe the system. Rather, in this case one must turn to the
equations of Einstein's Special Theory of Relativity.
A further note should be considered here. The equations of motion
that have been derived here and the Quantum Mechanic equations of
motion which London derived from the isentropic condition describe the
system as if "tends" toward an equilibrium. This is the origin of the
motion. That is, the tendency to seek a maximum of entropy for the
isolated system.
In this chapter there is no clear way to improve our understanding of
nuclear phenomena, nor is there any clarification of gravitational effects.
Further, ramifications of the theory will be pursued in the next three
chapters, which will bear on these points.

89

CHAPTER 3 - FIVE-DIMENSIONAL SYSTEMS


During the preceding development of the Dynamic Theory, there did
not appear to be anything that approached a description of nuclear effects.
Of course quantum theorists may respond that the nuclear effects lie
within the realm of quantum theory. This, however, does not seem to be a
strong argument since current nuclear theory appears to depend upon a
number of ad hoc postulates.
If it is supposed that nuclear theory cannot be extracted from some
aspect of the preceding four-dimensional world view, then how might the
Dynamic Theory produce a foundation for nuclear theory? At this point
there may appear to be no obvious way. Therefore, let us proceed on a
different tack.
Thus far we have constantly adhered to the policy of dividing
systems into two types: thermodynamic systems with only a work term of
the pdv type and mechanical systems with three mechanical, or spatial,
work terms. The generality of the adopted laws places no restrictions upon
the number or type of variables used. In particular, there is no restriction
coming from the laws themselves which says we cannot use four work
terms, one the thermodynamic pdv term and three mechanical Fdq terms.
Obviously pdv itself is just another Fdq type term with the pressure as the
generalized force and the volume as the generalized displacement.
When we teach thermodynamics we write the First Law with the
right hand side of the equation being the change in internal energy (system
energy), the thermodynamic work term, and three spatial work terms. We
then tell the students that since the right hand side of the equation
involves five unknowns we must have five independent equations in order
to have a solvable system. The first equation offered is the conservation of
mass which we state guarantees that we may write mass density as a
function of space and time. But is this really true for all space and for all
time?
The rub comes in attempting to visualize a world description in five
dimensions. Many arguments may be envisioned which tend to imply only
a four-dimensional manifold is needed. The kinetic theory of gases relates
the pressure to the average velocities of the particles contained. Does that
not imply that thermodynamics ultimately rests on a four-dimensional
manifold? Recall that the system in the kinetic theory is basically in
equilibrium.
Statistical thermodynamists may claim that thermodynamics is
basically statistical in nature and is fundamentally tied to order and
disorder and hence to the four-dimensional world of quantum theory. But
remember that the overall system, to which the statistical approach is
applicable, is a composite system made up of many subsystems each in an
equilibrium state. What happens to this argument when the number of
individual particles is not infinitely large?
90

Still there seems to be no substantial support for a five-dimensional


world from the point of view of current theories. This is to be expected
though in view of the difficulties experienced in the transition from the
classical three-dimensional world to the four-dimensional space-time of
Einstein's theories. Obviously had the extension of the universe been
restricted on a priori grounds to three-dimensional Euclidean space,
Einstein's theory would have been rejected on first principles. On the
other hand, as soon as we recognize that the fundamental continuum of
the universe and its geometry cannot be posited a priori and can only be
disclosed to us from place to place by experiment and measurement, a vast
number of possibilities are thrown open.
Among these the
four-dimensional space-time of relativity, with its varying degrees of
non-Euclideanism, has found a place. So also may the five-dimensional
view of the Dynamic Theory be found within the possibilities. Ultimate
judgment upon its necessity, or applicability, should rest upon a
comparison of the theory's predictions with reality.
A. Systems Near an Equilibrium State
The metric coefficients are made up of the second partial derivatives
of the system energy function and, therefore, if the system remains near an
equilibrium state, then the value of these derivatives evaluated at the
equilibrium state may be used as a first approximation for the metric
coefficients. In this case the geometry will be Euclidean and, from the
preceding four-dimensional development, the Euclidean manifold produced
by applying the E-conservative restriction was Minkowski's space-time
continuum of Special Relativity.
Therefore, suppose we begin an investigation of the five-dimensional
world by staying very near an equilibrium state so as to simplify the
description to a five-dimensional generalization of Minkowski's space-time
manifold.
3.1

Equations of Motion

Suppose that we consider some sort of system requiring four work


terms and for the moment not concern ourselves as to exactly what this
system might be. Thus, for our system we will have thermodynamic as
well as mechanical variables and the First Law becomes
~
~
_E = dU + Pdv - F~ dq ;

= 1,2,3.

Where the v and are considered as specific quantities. That is, these
quantities are related to a unit of mass such as is customary in
thermodynamics.
The specific volume is the reciprocal of the mass density . Using the
mass density instead of the specific volume the First Law becomes
91

~
~ P
d E = dU - 2 d - F~ dq ;

= 1,2,3.

This law now requires that the system's specific energy U be a function of
five independent variables so that U = U(S,q1,q2,q3). Thus, the First Law
requires a five-dimensional manifold of specific entropy, space, and mass
density for a general system. Since the system under consideration needs
both thermodynamic and mechanical variables, we can no longer refer to
the entropy as mechanical or thermodynamic; however, the limiting case
where the mass is held fixed must produce the mechanical entropy.
The procedure established by the Dynamic Theory is to take the
stability condition quadratic form as the metric for a stable system. Thus,
the coefficients of the metric become the second partial derivatives of the
energy function. In order to simplify the metric, suppose for the present
that we restrict our system to be very near an equilibrium state so that we
may consider the second partial derivatives to be constants. This is in
essence considering a local Euclidean manifold; the symmetry of the
geometric connections guarantees that we may do this.
Since the metric coefficients are constants, a transformation may be
found such that the cross terms are zero. Then in this coordinate system
and when
q

and q 4

F0

a0

the metric becomes


2
0 2
4 2
2
c (dt ) = ( dq ) + dq dq + ( dq ) ;

= 1,2,3.

(3.1)

If we again consider the restriction d = 0 so that we are talking of an


E-conservative system for which the principle of increasing entropy holds,
then we have the variational principle given by
(3.2)

(dS )2 = 0.

Solving Eqn. (3.1) for dq0 and squaring we get


( dq0 )2 = c 2 (dt )2 - dq dq - ( dq 4 )2 ;

= 1,2,3,

or
dq
dt

= c2 - g

dq
dt

dq
dt

, = 1,2,3,4.

The entropy manifold given by Eqn. (3.3) is a five-dimensional


Minkowski-type manifold with coordinates of space-time-mass. We may,
92

(3.3)

therefore, follow the procedure Minkowski and Einstein used in the Special
Theory of Relativity.
First, to avoid confusion, let us rename the coordinates as
x

1
q , x2

ct, x1

2
q , x3

3
q , and x 4

4
q .

Then define the five-dimensional velocity vector as

dx
; i = 0,1,2,3,4
0
dq

and define the five-dimensional acceleration vector as


f

d x

u
0
q

dq

i dx j dx k
.
0
0
jk dq dq

+
2

Now the specific entropy is the arc length and the variational
principle is based upon the entropy. Therefore, if we multiply the specific
entropy by the mass density, we have the entropy density. The variational
problem becomes
2

( dq0 )2 =

(3.4)

( dq0 )2 = 0.

Notice how the mass has entered our variational problem. It has entered
because our metric was in terms of the "specific entropy", or entropy per
unit mass. The variational problem is based upon the entropy, not the
specific entropy. Thus, the mass density is required in the variational
problem to correct this difference. The importance of this lies in the fact
that this is the origin of the "inertia" which appears in the following
equations of motion.
The Euler equations for this problem are
g ij u j

d
dq

g ij u i u j

g ik
-

g ij u i u j -

uu

x
g ij u i u j

=0

or

a u

g ij u j
g ij u i u j

g ij u u +

g ij u j

d
dq

g ij u i u j

g ik
-

Using the fact that gijuiuj = 1, the Euler equations become

93

i u u
x
g ij u i u j

= 0.

i
f =

- a0 u 4 g ij u j

(3.5)

where the Fi are force densities.


Obviously if we hold the mass density fixed, u4=0, then the volume
integral of this equation becomes the force-mass- acceleration relationship
of Special Relativity.
Now Since
f =

dq
dt

u
and
0
q

= c2 - u u ;

= 1,2,3,4

then
i

2
2
c -v

i
u dt
t dq0

u
=
0
q

i
F =

dx
t dq0

where v 2 = u u :

= 1,2,3,4.

Then
F =

= c2

c -v

dx
c - v dt

1
t

1-

1-

dx
,
dt

where = v/c with v being the four-dimensional speed.


The force density equation may now be written as
1-

F =

1-

dx
.
dt

Consider

but

1-

dx
dt

1
dt

1-

dx
+
dt

1
t

1-

dx
;
dt

/ t=a0v4, so that the force density equations may be written as

94

1-

F =

1
c

dx
dt

1-

a0 u
2
c

1
2

1-

dx
.
dt

We may define
1
2

1-

as the effective mass density or "relativistic" mass density; then


1-

By defining F

1-

F =

1
c

dx
dt

a0 v v

1-

F 23 so that

F =

dx
dt

a0 v v
1-

(3.6)

we see that this force density becomes Einstein's special relativistic force
density when v4 = 0, or for constant "rest mass." Thus, the equations of
motion, Eqn.s (3.6), reduce to Einstein's special relativistic equations of
motion when d /dt=0.
3.2

Energy Equation
Now for our system the restriction that
~
~ P
_E = 0 = dU - 2 d - F~ dx ,

= 1,2,3

requires that
p
~
dU = 2 d + F~ dx ,

or if p/

= 1,2,3,

is considered as another generalized force density, then


~
dU = F~ dx ,

= 1,2,3,4.

Thus, by integrating the expression for the system's specific energy


change, we should arrive at the Einstein energy equation if we hold dx4/dt
0. Therefore, we shall perform the integration using the force densities
given by Eqn. (3.6) to get the system's energy, or
95

~
U - U~ 0 =

p
p0

~
F dx =

d
dt

t
t0

d
dt

p
p0

u u +

1-

1-

dx
dt

du
dt

1-

a0 u u
1-

dx

dt.

But c22 = u u and c2(d/dt) = u du /dt; therefore,


~
U - U~ 0 =

d
dt

t
t0

1-

= c2

3
2 2

(1 -

&

1-

dt

Now depends upon u and not upon x4 or ; therefore


~
U - U~ 0 =

c
1-

2
2

or
~
U =

2
2

1-

+ constant.

If the internal energy is considered to be the system's energy when


the spatial velocities u ;
= 1, 2, 3 are taken as zero, then the internal
energy density given by
~
U =

2
4

+ constant.

1- u
c

At the condition where u4 is also zero the internal energy density is then
~
U =

2
c + constant.

By taking the constant of integration to be zero, this internal energy


density then is seen to correspond to Einstein's "rest energy" where here
the "rest energy" is in terms of a four-dimensional "at rest" state.
If we make the usual approximation of allowing 2<<1, then the
system's energy density is approximately given by
96

~
U =

2
c +

1 2 1
( &)2 ,
v +
2
2 ( a0 )2

where here u4 = d /dt is used. This displays the classical limit system
energy density for an E-conservative system very near equilibrium.
B. Systems With Non-Euclidean Manifold
Suppose now we relax the assumption that the system is very near
an equilibrium point so that the second partial derivatives are no longer
constants but are functions. This is essentially the same transition as
Einstein made going from his Special to General theory; however, the logic
of the transition is much simpler here. The only change in the logic
appears in the relaxation of the assumption of nearness. There is, of
course, a drastic increase in mathematical difficulty since the metric components are no longer constants.
3.3

General Variational Principle

We shall consider a system that must be described by both


thermodynamic and mechanical variables. When written in terms of the
mass density, the First Law for this system may be written as
~
~ P
_E = dU - 2 d - F~ dq ,

= 1,2,3,

where the tilde denotes specific quantities.


Following the prescribed procedures of the Dynamic Theory we shall
take the stability condition quadratic form as the metric for our system.
Thus, the metric coefficients will be given by the second partial derivatives
2

hij =

U
q

, i, j = 0,1,2,3,4,

where q4= /a0. The metric may then be written as


2
0 2
0
2
c (dt ) = h00 ( dq ) + 2 h0 dq dq + h dq dq ,

where ,=1,2,3,4.
Imposing the restriction that the system be E-conservative,
results in the principle of increasing entropy, so that

(3.3A)
E=0,

(dS )2 = 0.

In terms of the specific entropy the variational principle may be


written as
97

( dq0 )2 =

(3.4)

( dq0 )2 = 0.

Solving the metric given by Eqn. (3.3A) and squaring yields the
expression
( dq0 )2 =

1
h00

{ c 2 (dt )2 + 2 h0 Adtdq - h dq dq },

, = 1,2,3,4,

(3.7)

with
A=

h0

2
2
c h & & ( h0 q& )
.
q q +
( h00 )2
h00 h00

q& + _

h00

This metric in a five-dimensional manifold of space-time-mass may


be rewritten as
( dq0 )2 =

2
) ,

(d

h00

where
( dq0 )2

g ij dxi dx j , i, j = 0,1,2,3,4,

and
(d

q ij dxi dx j , i, j = 0,1,2,3,4,

with x0=ct, x1=q1, x2=q2, x3=q3, and x4= /a0. Thus we may write
( dq0 )2 = q ij dxi dx j =

1
(d
f

2
) =

1
g ij dxi dx j .
f

(3.8)

Having established the metrics in Eqn. (3.8) in the manner


prescribed by the Dynamic Theory, the geometry must be Weyl geometry;
wherein the potential five-vector is defined as
+_

lnf
x

1
2

(3.9)

and the field tensor is given by


F ij

98

i, j

j,i

(3.10)

3.4

Gauge Function Field Equations

In order to isolate the field equations resulting from a gauge function


from the field equations produced by a vector curvature, let us consider a
local Euclidean manifold for (d )2.
Now the field tensor given by Eqn. (3.10) has 25 components. We
would like to determine the field equations for these components. The
quickest, though not the only, way is to consider the five dimensions to be
x0=ict, x =x , =1,2,3,4. The field tensor is then defined to be
0

iE 1

- iE 1

iE 2

B3 - B2

- iV 4

V1

B1

V2.

- B1

V3

-V 1 -V 2 -V 3

F ij = - iE 2 - B 3
- iE 3

iE 3 iV 4

B2

Using Bianchi's identities


F ij
F jk
F ki
+
+
=0
k
i
j
x
x
x

and the various combinations of the indices 0, 1, 2, 3, 4 we obtain the field


equations
B=0
xE +

1 B
=0
c t

xV + a0
V4+

=0

(3.11)

1 V
E
+ a0
= 0.
c t

The definition of the five-vector current density


F ij
i
x

4
Ji
c

(3.12)

yields the equations


E + a0

V4

= 4

1 E
V
4 J
=
+ a0
c t
c
1 V4
4 J4
V+
= .
c t
c

xB -

99

(3.13)

In addition to these field equations there is the statement of


conservation of charge where
Ji
= 0, i = 0,1,2,3,4,
i
x

so that
t

J4

J + a0

(3.14)

= 0.

For ease in future reference to these eight field equations they may
be rewritten as
B=0

[a]

1 B
+ xE = 0
c t
xB -

V
4 J
1 E
=
+ a0
c
c t
V4

E + a0
t

J + a0
xV + a0

1
c
1
V+
c

V0+

3.5

[b]

= 4
J4

[c]
[d]

=0

(3.15)

[e]

=0

[f]

V
E
[g]
= a0
t
4 J4
V4
= [h]
t
c

Energy-Momentum Tensor

If we follow the approach of relativistic electrodynamics, we may


define the tensor {T} in terms of the field tensor {F} according to
T jk

1
4

k
F j F +

1
4

jk

F st F st

Using the field tensor to calculate the components of


energy-momentum tensor we find that the components are given by

100

the

-1
[( E x B ) + V 4 V ] , = 1, 2, 3 ,
4
1
[ E 2 + B 2 + V 42 + V 2 ] ,
T 00 =
8
i
[E V ] ,
T 04 =
4
1
=
[ V 4 E + ( V x B ) ] , = 1, 2, 3 ,
4
1
[ V 42 + B 2 - E 2 - V 2 ] ,
T 44 =
8

T0 =

T4

and
T

1
1
{ E E + B B -V V 4
2

[ E 2 + B 2 + V 42 - V 2 ]} ,

where
, = 1, 2, 3.

Equations (3.15) form a set of eight Maxwell-type equations which


obviously reduce to Maxwell's four equations.
The wave equations for the new field quantities may be derived using
standard assumptions.
t

V )+

1
c

V4 4
=
2
c
t

V 1
+
dt c

J4
=
dt

V4
2
t

and
( V 4 )+

1
c

V
= - a0
t

1
c

V
+
dt

Therefore,

V4-

1
c

V4 4
= 2
2
t
c

J4
- a0
t

For the vector field we have:


(

V )+

4
V4
=- (
J4 )
t
c

and
x( xV ) +

V+

101

1
c t

V4=

4
c

J4 ;

V .

therefore
2

But

E=4

- a0

V4

V-

4
c

a0
c

J4+

+ a0 (

) .

70, so that
2

and x B -

V4-

V4 4
= 2
2
t
c

J4
- a0
t

- a0

V4
t

1 E 4
V
72 , so that
J - a0
c t c
2

V-

1
c

4
c

4
E
V
J +2
- a0
.
c
t

J 4 + a0

Now the wave equations for the usual vector and scalar potentials
are
2

A-

1
c

=-

4
J
c

and
2

=-4

We may differentiate these with respect to the mass density and substitute
them into our wave equations and get
2

V4 -

1
c

V4=

4
c

J4
+ a02
t

V4
2

and
2

V-

1
c

4
c

J 4 + a0

E
V
- a0
t

where
V 4 V 4 + a0

and V = V - a0

The field energy density may be defined by


1
[ E E + B B + V V + V 42 ] ,
8

102

and the electrical Poynting vector may be defined by


c
(E x B) .
4

SE

Now the electrical Poynting vector represents the outward flow of the
electromagnetic field energy through a surface. If we take the total vector,
whose components are T0 , to be the total flow of energy, then the vector
with components

c
V 4 V 81must be the outward flow of energy due to changes of the
4

mass density within the surface. Let us designate the mass energy vector as
Sm

c
(V 4V ) ,
4

so that the total energy vector is


S = SE + Sm

whose components are


S =

c
c
[ ( E x B ) +V 4V ] = - ( ) T 0
4
i

The Dynamic stress tensor may be defined as the three-dimensional


tensor whose elements are
T

1
1
{ E E + B B -V V 4
2

[ - 2 V 2 ]}

The Maxwell stress tensor is defined in electrodynamics as the


three-dimensional tensor with elements
T

1
1
{E E +B B 4
2

[ E 2 + B 2 ]}

In terms of the Maxwell stress tensor, the Dynamic stress tensor may be
written as
T

= T M - {V V -

1
2

Then in terms of the above defined quantities

103

[ V 2 - V 42 ]} .

1
S
c

i
S
c

{T D }

i
(E V )
4

[ V 4 E + ( V x B )]

{T} =

i
(E V )
4
i
V 4 E + ( V x B )] .
4
1
[ V 42 + B 2 - E 2 - V 2 ]
8

Suppose we calculate the trace of the energy-momentum tensor:


D
t R {T} = T jj = T jj + +

1
[ V 42 + B 2 - E 2 - V 2 ]
8

1
[ B 2 + V 42 ] + T 4
4
1
1
3
=
[ B 2 + V 42 ] +
[ E 2 + B 2 - V 2 - ( E 2 + B 2 + V 42 - V 2 )]
4
4
2
1 1 2 1 2 1 2
=
[ B - V - V4]
4
2
2
2
1
=
[ B 2 + V 2 - E 2 - V 42 ] .
8
=

3.6

Force Density Vector.

The force density vector may be defined in terms of the divergence of


the energy-momentum tensor.
Therefore, suppose we calculate the
five-dimensional divergence of the tensor {T}, or

1
T jk
=
k
4
x

[ F j F k +

1
4

jk

F st F st ] .

Because of the antisymmetry of Fjk, the first term may be written as


F j
F j
=
.
k F k
k F k
x
x

By interchanging the indices k and l


1
F j
F kj
=
=
k F k
F k
2
x
x

F j
F j
+
F k .
k

x
x

Using the Bianchi identity


F j
F kj
F k
+
+
=0 ,
k

j
x
x
x

the terms contained within the parentheses may be written as


104

1
F j
=
k F k
2
x

1 ( F k F k )
F k
=.
j F k
i
4
x
x

Substituting this back into the expression for the divergence, the last term
will be canceled because l, k, s, and t are dummy indices. Then the
divergence becomes
1
T jk
F k
=
.
F j
k
k
4
x
x

By interchanging the indices k and l on the right-hand side we obtain


1
T jk
=
F jk
xk 4

F k
.

(3.16)

The Dynamic force density five-vector may now be defined as


K

Div 5 {T} .

Therefore, the components of K are given by


K j=

1
F k
.
F jk

4
x

But the five-vector current density is given by


F k -4
=
Jk .

c
x

The components of the five-vector force density become


K j=

-1
4
-1
J k = J k F kj .
F kj
4
c
c

Now, since Jk = (ic , J, J4), then


i
[ J E + J 4V 4 ] ,
c
,
1
J
K = [ E + ( u x B )] + 4 V
c
c
K0 =

(3.17)

and
K4= V 4 -

J V
,
c

where J = u. These then are the components of the force density


five-vector resulting from a gauge field in the Dynamic Theory. These
components reduce to the four components of the Lorentz force density
should V4 = V = 0.
105

With the interpretation that the four force density components with
subscript 1 through 4 are the force density vectors which appear in the
First Law as F , then the force density vector provides the connection
between the First Law and the geometry of the sigma manifold discussed in
section 2.9. Thus, the existence of the vector field i is also demanded by
the Dynamic Theory and need not exist as a separate assumption.
3.7

Equation of Energy Flow.

Consider the zeroth component of the Dynamic force density


five-vector
T 0k
T 00
T0
T 04
=
+
+
.
k
0
4
x
x
x
x

K0=

Then
i
i
[ J E + J 4V 4 ] =
c
(ict) c

i
S
+
4
x

(E V )
x

or
1
1
1 S
1
[ J E + J 4V 4 ] =
+
c
c t c x 4

(E V )
x

or, since x4 = /a0,


- J E - J 4V 4 =

S a0 c ( E V )
.
4
x

Rearranging the terms


div S +

= - J E - J 4V 4 +

a0 c ( E V )
4

and separating out the electrical Poynting vector leads to


div S E +

= - J E - J 4 V 4 - div S m +

a0 c ( E V )
.
4

This then is the five-dimensional energy flow equation.


3.8

Momentum Conservation

The expression for the conservation of momentum may be obtained


from the space portion of the force density five-vector
106

T
T k
T 0
T 4
=
+
+
,
k
0
4
x
x
x
x

K=

, = 1, 2, 3 .

But T 110 is the three-dimensional divergence of the Dynamic stress tensor {TD},
x

therefore,
1

K=-

S
a
+ div { T D } + 0
t
4

[ V 4 E + ( V x B )] .

If we consider a volume in which all the material is contained and


outside of which the field vanishes, then integrating over this volume yields
{K+
v

1
c

S a0
t 4

[ V 4 E + ( V x B )]} dv = div { T D } dv .
v

The integral of K gives the total force (i.e., the time derivative of the
mechanical momentum p less the vector &v / 1 - 2 113. Now define the vector
g

S
c

a0
{
4

[V 0 E + (V x B )+

&v
1-

} dt .

Then define
gd

G ,

so that
d
( p + G ) = div { T D } dv .
dt
v

Using the divergence theorem the volume integral may be converted to a


surface integral so that
d
( p + G ) = { T D } n da .
dt
s

If the field vanishes outside of V, it must do so also on the boundary


surface s, hence
d
( p + G )= 0 .
dt

Therefore, it is not the mechanical momentum p but the quantity p +


G which is conserved. Therefore, we must interpret G as the momentum of
the field and
107

g=

S
c

a0
4

V E +(V xB ) +

4 &v
1-

dt

as the momentum density of the field.


3.9

Gauge Field Pressure


The Dynamic stress tensor is given by
T

1
1
2
2
2
2
{ E E + B B -V V E + B + V 4 - V ]}
4
2
1
3 2
= [ E 2 + B2 - V 2 ] V4 .
8
8

Now separate the three-dimensional dynamic stress tensor into a traceless


and an isotropic tensor.
1
1
{ E E + B B -V V [ E 2 + B 2 + V 42 - V 2 }]
4
2
1
1
=
{ E E + B B -V V [ E 2 + B 2 + V 42 - V 2 ]}
4
8
1
2
1
=
[ E E + B B - V V ] - ( )(
)
[ E 2 + B2 - V 2 ]
4
3 8
1
1
- ( )(
)
[ E 2 + B 2 + 3V 42 - V 2 ]
3 8
t +

where
t

1
1
{ E E + B B +V V - (
)
4
3

[ E 2 + B2 - V 2 ]

and
-(

1
)
24

[ E 2 + B 2 + 3V 42 - V 2 ] .

Now
tr { t }= (

1
) [ E 2 + B 2 - V 2 - ( E 2 + B 2 - V 2 )] 0
4

and
tr {

}= - (

1
) [ E 2 + B 2 + 3 V 42 - V 2 ] .
8

108

Consider the definition


1
t
3

= -(

1
)
24

[ E 2 + B 2 + 3V 42 - V 2 ] .

Then
t= -(

1
) [ E 2 + B 2 + 3V 42 - V 2 ]
8

and
t 0 0
= 0

t 0

0 0

The isotropic part of the stress tensor is usually called the


"pressure." Therefore, define 3p = t in accordance with customary
notation, so that
p= -(

1
) [ E 2 + B 2 + 3V 42 - V 2 ] .
24

With the exception of the factor of 3 this reduces to the "radiation


pressure" for an electromagnetic field when V = V0 =0. Note that this
pressure may be zero since it is the sum and difference of squares, or p =
0, when
2
2
2
2
V = E + B + 3V 4 .

This may prove to be an important point when considering boundary


conditions in cosmology or the study of elementary particles.

109

Chapter 4.

QUANTIZATION IN FIVE DIMENSIONS

The preceding development provides a tremendous wealth of mathematical abstractions. However, there seems within it no readily apparent
method of interpreting the new fields. If there appears to be no physical
entity which may be associated with the new field quantities, then the
development will have gone for naught. On the other hand, with the notion
of nuclear fields in mind it seems that if the new field quantities are
included in a quantized picture, then perhaps the relation to nuclear fields
may be made.
In the following the requirement for quantization is provided by
appropriate restrictions upon a system whose description is taken from the
Dynamic Theory. However, the use of the five-dimensional Dirac equation
has not yet been shown to result from the Dynamic Theory. Schrodinger's
quantum mechanics may be obtained using London's work, but I am not
aware of a procedure to arrive logically at Dirac's equation even though I
feel that the method exists. As it now stands, the use of the generalized
Dirac equation must be accepted as an independent fundamental assumption.
4.1 Quantization.
The system under consideration now is a five-dimensional system
with arc element
( dq0 )2 = f(d

2
) .

Now since our system is an E-conservative, dE = 0, system the principle of


increasing entropy requires that (dq0)2 > 0 so that f(d )2 0. Introducing
the quantization conditions results in
j

j
dx = 2 in , j = 0, 1, 2, 3, 4 ,

where
nf
j

1/2
j

and x0 ct , x1 q1 , x 2 q 2 , x 3 , x 4

a0

If we restrict ourselves to a (d )2 space which is the local Euclidean space,


then (d )2 is the five-dimensional Minkowski-type manifold; using London's
work we would produce a five-dimensional quantum dynamical system.

110

4.2 Five-Dimensional Hamiltonian.


We previously showed that the principle of increasing entropy
resulted in
( dq0 )2 = 0

as the variational principle for a local Euclidean manifold.


Since
multiplication by a constant does not change the problem we may take our
variational problem to be
c

( dq0 )2 = 0 .

Defining the velocity vector as uj=dxj/dq0 and the momentum as


pj= L/ uj= gjkuk, where we have used the fact that gjkujuk = 1, then we may
form the contravariant momentum as
jk

p j = g pk = g

jk

gk u

so that
j
jk
p j p =( g j u j )( g

g k u )=

k
u g jk u

(4.1)

= ( g ik u j u k )
2

since
c2 =
gjkujuk.
Equation (4.1) is the five-dimensional
"momentum-energy" equation.
We may now follow London's procedure to obtain our wave function
for the five-dimensional system. However, a quicker way to investigate the
effect of the Dynamic Theory upon quantum mechanics would seem to be
that of adopting Dirac's equation in a five-dimensional form and following a
development analogous to standard four-dimensional relativistic quantum
mechanics. With this in mind, then we shall adopt the form
h = i ~1

+ ~2

+ ~3

+ ~4

(4.2)

to be the five-dimensional specific Hamiltonian operator. The partial


derivative operators are specific operators and hence are dimensionless in
natural units. In Eqn. (4.1) the 's and do not involve derivatives and
must be Hermitian in order that h be Hermitian.
By taking the four partial derivatives in Eqn. (4.1) as the four-vector
specific momentum operator we may write
h= -( ~ p +

111

) .

(4.3)

4.3 Five-Dimensional Dirac Equation.


If we take p0 |> = h |> 10 and require that the 's and are chosen such that
solutions of this equation are also solutions of Eqn. (4.3), we find the
restrictions imposed upon the choice of the 's and to be:
( ~ p )= p2 ,
~2
=1 ,

and
~~ + ~ ~ = 0 ,

(4.4)

where natural units, c = 1, are used.


A set of 8 x 8 matrices satisfying the requirements of Eqn. (4.4) is
~

, ~i =

i = 1, 2, 3, , ~ 4 =

(4.5)

where
=

0 -I

i = 1, 2, 3 , and A =

and
I=

1 0
0 1

0 1
1 0

0 -i
i

, and

0 -1

Then the five-dimensional Dirac equation may be taken to be


i

(x) = (i ~

(x) ,

where the 17is a four-dimensional operator. By defining


0

~~

( = 1, 2, 3, 4) ,

then Eqn. (4.6) may be written as


(i

j
j

+ 1)

(x) = 0 .
~

By virtue of the properties of the ~ s 20 and 21 plus the fact that


1 for j = k = 0
g = { - 1 for j = k = 1, 2, 3, 4 ,
0 for j k
jk

112

(4.6)

the anticommutator of the -matrices must satisfy


j

, i } = 2 g ji .

4.4 "Lorentz" Covariance.


Under a five-dimensional Lorentz transformation
j
k
x = L kj x

we shall suppose each component of the wave function


into a linear combination of all four components:
(x) LT

(x ) = S

(x) transforms

(x) ,

where S is a Dirac spinor satisfying


S

-1

S = L kj

(4.7)
By using an infinitesimal Lorentz transformation given by
j

where

j
k

(4.8)

j
k

Lk = g k + d

28 are a set of 16 numbers, then S ( ) may be shown to be given by


S ( ) = exp (T

d ) ,
0

where the matrix T is given by


T=

1
4

jk

Equations (4.7), (4.8), and (4.9) suffice to guarantee the Lorentz covariance
of the five-dimensional Dirac equation.
Following standard quantum mechanical procedure we shall adopt
the probability current density to be
k
j (x) =

(x)

(x)

with the requirements that:


i
k j = 0 ,
jktransforms

jk

as a contravariant vector, and


must be real.

113

(4.9)

4.5 Spin.
In the three-dimensional space the angular momentum is given by
the vector L as the cross product of the coordinates and momenta. We
shall then define the angular four-momentum to be the four-dimensional
cross product
L

ijk

j
x p

where x4 is the mass density and

ijk

0 if any two indices are alike,


1 for even permutation to align indices in ascending
order,
-1 for odd permutation to align indices in ascending
order.

Then the commutator of the components of the angular


four-momentum with the specific Hamiltonian is not zero; for instance
[ L 3 , h] = i

2
p -i

1
p +i

1
p -i

4
p +i

2
p -i

Now suppose there exists a four-spin vector S such that the sum of the
angular four-momentum and the four-spin vector commutes with the
specific Hamiltonian; then if we define a new three-spin vector u given by
the components ui = i 4 1, u2=i 4 2/2, and u3=i 4 3/2, and take the usual
spin vector s, given by s1=i 2 3/2, s2=i 1 3/2, and s3=i 1 2/2, the components
of the four-spin vector may be shown to be
S 1 = s1 - u 2 - u 3 ,
S 2 = s 2 + u1 - u 3 ,
S 3 = s3 + u1 + u 2 ,

and
S 4 = s1 - s 2 + s 3 .

In analogy with standard relativistic quantum mechanics the


eigenvalues of the four-spin components can be shown to be + _

3
36. It
4

may also be shown that the set of observables P, h, and S where P is the four-momentum
and S is the four-spin, form a complete set of commuting observables.

114

4.6. Dirac Equation with Fields.


In analogy with relativistic quantum mechanics we take the
five-dimensional Dirac equation to be
[(i

where

+ 1]

(4.10)

=0 ,

is five-vector potential. By operating on the left with


[(i j - j ) - 1] 38 and separating j k 39 into symmetric and antisymmetric parts as
j

1
= {
2

1
[
2

}+

ik
g +

jk

then Eqn. (4.10) becomes


[(i

Separating

)(i

) - 1 + (-

-i

-i

jk

= 0.

(4.11)

42 into symmetric and antisymmetric parts as


j

1
(
2

F jk =

1
)+ (
2

and defining the field tensor as


k

Eqn. (4.11) becomes


[(i

)(i

)-1-

1
iF jk
2

jk

=0 .

Now since
0
- x1
jk

= - x2

0 - 2i s1

- n1

2i s 2

2i s 3

- n2

0 n3
- n3

where
j

0 - 2i s 3 n 2 ,

2i s1

- x3 - 2i s 2
- x4

4 n j = 1, 2, 3, and

115

(3.28)

E1

E2

E3 V 0

- E1

B3

B2 V 1

F jk = - E 2

B3

E 3 - B2
-V 0

0 - B1 V 2
0 V3

B1

-V 1 -V 2 -V 3

plus recalling the seven Maxwell-type equations from Eqn. (3.15)


V4+
xB-

1 V
E
+ a0
= 0,
c t

V+

1 E 4 J
V
=
- a0
,
c t
c

B = 0,

x E+

4
V4
=J4 ,
t
c
x V + a0

1 B
= 0,
c t

E=4

= 0,

- a0

V4

(4.13)

then Eqn. (4.12) may be written as


[(i

)(i

) - 1 + 2 B s - i E x - i V 4 x4 - i n V ]

= 0.

and thus becomes the Dirac equation with fields E, B, V4, and V.
Suppose we consider a system without an electric charge so that p =
J = 0, then by Eqn. (4.13) we still have
E = - a0

V4

and

xB-

1 E
V
= - a0
c t

and, therefore, there will still be a magnetic moment.


4.7. Allowed Fundamental Spin States.
In the five-dimensional quantization of the space-time-mass
manifold three spin vectors appear.
One of these is the familiar
three-component spin vector of relativistic quantum mechanics. The
second of the three is a new three-component spin vector while the
remaining one is a four-component spin vector.
Using the theorem:
If
of

satisfies
are a.

= a2 where a is a number, then the eigenvalues

Then it is not difficult to show that the component eigenvalues are


s =+ _

1
1
3
, u = + _ , S 2j = , = 1, 2, 3 and j = 1, 2, 3, 4 .
2
2
4

116

If, in analogy with the eigenvalues for the total angular momentum, we
write
3
2
S j = = S j ( S j + 1) ,
4

then the possible eigenvalues becomes


s =+ _

1
1
1 3
,u = + _ , S j = ,.
2
2
2 2

However, the following relations, which were shown to be required


for S to commute with the specific Hamiltonian, restrict the number of
possible combinations of these eigenvalues.
S 1 = s1 - u 2 - u 3 ,
S 2 = s 2 + u1 - u 3 ,
S 3 = s3 + u1 + u 2 ,

and
S 4 = s1 - s 2 + s 3 .

The question to be asked now is, how many combinations of the


above eigenvalues are allowed?
For S1 = 1/2 the combination s1 = -1/2 and u2 = 1/2 is impossible.
For S1 = 1/2 the combination s2 = -1/2 and u1 = -1/2 is impossible.
For S3 = 1/2 the combination s3 = -1/2 and u1 = -1/2 is impossible.
For S4 = 1/2 the combination s1 = -1/2 and s2 = 1/2 is impossible.
For S1 = -3/2 only one combination is possible: s1 = -1/2, u2 = 1/2,
and u3 = 1/2.
For S2 = -3/2 only one combination is possible: s2 = -1/2, u1 = -1/2,
and u3 = 1/2.
For S3 = -3/2 only one combination is possible: s3 = -1/2, u1 = -1/2,
and u2 = -1/2.
For S4 = -3/2 only one combination is possible: s1 = -1/2, s2 = 1/2,
and s3 = -1/2.
Now because S4 is a combination of the first terms of each of the
components S1, S2, S3, not all of the above listed 16 combinations are
possible.
For S4 = 1/2 the following combinations of (s1, s2, s3, u1, u2, u3) are
possible.

117

(1)
(2)
(3)
(4)

( , , ;- ,
( , , ; ,( ,- ,- ;
(- , - , ;

,, )
, ,- )
,- ,- )

for S1 = S2 = S3 =
for S1 = S2 = S3 =
for S1 = S2 = S3 =
for S1 = S2 = S3 =

The remaining combinations are:


(5)
(6)
(7)
(8)

(- , , - ; - , - , - )
(- , , - ; , , )
( ,- ,- ;- - , )
(- , - , ; - , , )

for S4 = S3 - ; S1 = S2 =
for S4 = S1 - ; S2 = S3 =
for S2 = S3 = - ; S1 = S4 =
for S1 = S3 = - ; S2 = S4 = .

Thus there is an octet of possible combinations. There are also some


obvious symmetries in these combinations. An aid in seeing these
symmetries is the vector defined as t where
t 1 - ( u 2 + u 3 ) ; t 2 u1 - u 3 ; t 3 u 2 + u1

Then for each of the eight combinations above we find (t1, t2, t3) given by
(1)
(2)
(3)
(4)

t= (0,0,0)
t=(0,0,0)
t= (0,1,1)
t= (1,1,0)

(5) t= (1, 0, -1)


(6) t= (-1, 0, 1)
(7) t= (0, -1, -1)
(8) t= (-1, -1, 0)

Thus, the eight combinations correspond to four distinct t vectors which


carry a sign. Or
t1=(0, 0,0) ;t2 = (0, 1, 1) ; t3 = (1, 1, 0) ; t4 = (1, 0, -1)
For + t we have:
t1
(s;u) = ( , , ; - , , - )
t2 (s;u) = ( , - , - ; , , - )
t3
(s;u) = (- , - , ; , - , - )
t4
(s;u) = (- , , - ; - , - , - ) .
For -t we have:
-t1
(s;u) = (s,u) = ( , , ; , - , )
-t2
(s;u) = (s,u) = ( , - , - ; - , - , )
-t3
(s;u) = (s,u) = (- , - , ; - , , )
-t4
(s;a) = (s,u) = (- , , - ; , , )
Now by defining the vectors:
a=( , , );
c=( ,- ,- );

b= (- , , - )
d =( , , - )
118

We may write
t1
t2
t3
t4

(s;u) = (a;b)
(s;u) = (c;d)
(s;u) = (-d;c)
(s;u) = (b;-a)

-t1
-t2
-t3
-t4

(s;u) = (a;-b)
(s;u) = (c;-d)
(s;u) = (-d;c)
(s;u) = (b;a)

The octet is then made up of the combinations:


(a; b); (c; d); (b; a); (-d; c) .
The appearance of octets for basic quantum numbers is reminiscent of
elementary particle theory. Thus, the Dynamic Theory seems to give
promise to the hope of tying elementary particles to fundamental principles
in a new way.
B. Quantized Fields
Much difficulty was encountered in trying to find a solution to the
wave equations. This stimulated a return to thoughts of fundamental
particles. The motivation for this change was primarily the feeling that it
would be more productive to get away from the wave solutions for a while,
but also there was the haunting feeling, retained for some five years, that
the new fields played a role in particle structure. This feeling was based
primarily on the role the new fields appear to play in the five-dimensional
quantization and their role in the self-energy of charged particles.
4.8. Quantum Condition Applied to Particles.
The quantum condition
j

j
dx = 2 iN, j = 0, 1,

, n,

(4.14)

was required when generalized isentropic states were considered. Given


the thermodynamic basis for the three fundamental laws, it seems natural
to think that if the Dynamic Theory were to say anything about
fundamental particles then it should probably come from considering
generalized isentropic states. Thus, the quantum condition, Eqn. (4.14),
should play a crucial role. This also was the condition from which London
began his work, which showed that this condition produces quantum
mechanics, but quantum mechanics describes interactions between
particles such as electrons and nuclei. It does not specify what types of
particles are allowed.
119

That the three adopted laws must apply to individual fundamental


particles is tantamount to the notion that these three laws must specify
what particles are allowed and, thereby, must specify their allowed fields.
If we again look at the quantum condition, Eqn. (4.14), we see that it is
given as a line integral that must have a quantized value. Our usual first
encounter with a line integral involves the evaluation of a given line
integral when the path is specified. Because the quantum condition
represents a line integral that may only have certain values, London asked
a legitimate question when he asked what paths would be allowed given
the electrostatic potential in advance8. This points out that there are three
parts to any line integral, the integrand, the path, and the integral value.
Another question that may be asked of the quantum condition is
given that the integral value may only be 2 iN, what are the possible
allowed for a particle that must retain its identity along any path? This is
equivalent to asking what fields are fundamental particles allowed to have
if we are free to move them anywhere in the manifold? To be more specific,
we are asking what are allowed by the quantum condition if the dxj are to
be independent?
If the dxj are to be independent, then we may choose all dxj to be
zero except dxk. Then the quantum condition requires
k

k
dx = 2 iN (no sum on k).

(4.15)

Equation (4.15) must be true for all k, and because we are free to set the
path, then the k must reflect the quantization represented by the integer
N. Therefore,
j

=Nj

~
j

(no sum) ,

(4.16)

where the may not be quantized. Thus, Eqn. (4.16) represents the first
response of the quantum condition to the question concerning what j are
allowed for fundamental particles; the gauge potentials must be quantized.
The definition of the gauge potentials is
j

lnf

1
2

(4.17)

xj

where f was the gauge function. The field tensor was defined by
F jk =

j,k

k, j

(4.18)

where covariant differentiation is required. There are restrictions placed


upon these fields, for they must obey the set of eight differential equations
given by Eqn. (3.15).

120

4.9. Radial Field Dependence.


Any potential

=Nj

62 allowed by the quantum condition must also satisfy

Eqn.s. (3.15). Even so, Eqn.s. (3.15), (4.17) and (4.18) represent three stages of
differentiation, starting with the gauge function, f. In looking for the restrictions Eqn.s.
(3.15) place upon the quantized potentials, we may employ a technique of mathematics in
the solution of differential equations. We try to find a solution in the form of the product of
functions of the separate variables. However, our trial solution must produce potentials of
the form in Eqn. (4.16). Therefore, suppose we try to find a solution of the form
1

(4.19)

lnf 2 = FG

where
F= ft fr f f f

with ft = function of time only, fr = function of spherical radius only, etc.,


and the function G is defined by the system of partial differential
equations,
G
x

( N j - G) F
(no sum) .
j
F
x

The definition of the gauge potential, using the trial solution of Eqn. (4.19),
now produces
j

1
2

lnf

(FG)
x

=G

+F

G
x

but by the defining relations for G this becomes


j

=G

F
x

+ ( N j - G)

=Nj

F
x

(no sum).

If we define
~
j

then we have the proper form


j

=Nj

~
j

(no sum).

(4.20)

We may now use our trial solution to write the potentials in


spherical coordinates:

121

lnf
1

1
2=
N1 f t f r f f f = N1 F1 ,

r
1
lnf
1
2 = N2 f f f
=
2
t
r
r
r
1
3=
rsin

lnf
lnf

= a0

1
2

N3
rsin

f f =

ft fr f f

N2
F2 ,
r

f =

N3
rsin

F3 ,

1
2=
a0 N 4 f t f r f f f = a0 N 4 F 4 ,
1

lnf 2
N0
N0
=
ft fr f f f =
F0 ,
0=
(ilc)
ic
ic

(4.21)

where the notation f t 72 denotes dft/dt and Fj denoted


dF
dt
dF
F1=
dr
dF
F2=
d
dF
F3=
d
F0=

,
,
,
,

and
F4=

dF
.
d

Substituting the potentials given by Eqn. (4.21) into the definition of the
field tensor and using the required covariant differentiation, we obtain the
field components

122

( N1 - N0 )
F 01 ,
c
( N2 - N0 )
E =
F 02 ,
cr
( N3 - N0 )
E =
F 03 ,
crsin
a0 ( N 4 - N 0 ) F 04
,
V4=
c
( N 2 - N 3 ) F 23
- N 3 cot F 3
Br =
rsin
r
Er =

B =

( N3 - N1 )
N3 F3
F 13 rsin
r

( N1 - N 2 )
N2 F2
F 12 r
r
V r = ( N 1 - N 4 ) a0 F 14 ,

B =

(4.22)

,
,
,

V = ( N 2 - N 4 ) a0 F 24 ,

and
V = ( N 3 - N 4 ) a0 F 34

These field components reflect the quantization of the potentials.


However, the quantization of the fields is not a simple quantization
because each component depends upon the difference of quantum
numbers.
The field components given by Eqn. (4.22) must satisfy the differential equations of Eqn. (3.15). Therefore, if we substitute Eqn. (4.22) into
Eqn. (3.15), we will obtain the restrictions upon the quantum numbers, Nj,
and the functions ft, fr, f , f , and f required for these fields to be the fields
of a fundamental particle. We begin with the equation
(3.15a)

B=0 .

This equation becomes


1
r2

1
Br r 2
+
r
rsin

(sin B )

1
rsin

B =0

in spherical coordinates. Substituting from Eqn. (4.22) and simplifying, we


finally arrive at
1
r 3 sin

( N 3 - N 2 ) N 3 r cot ( f 3 + r f 31 ) + F 23 ( N 2 - N 1 ) N 2 + ( N 1 - N 3 ) N 3 = 0

This requires that


( N 2 - N 3 ) N 3 r cot ( F 3 + r F 31 = F 23 [( N 2 - N 1 ) + ( N 1 - N 3 ) N 3 ] .

Substituting from the definition of the Fj, we find


123

( N 2 - N 3 ) N 3 cot r f t f r f F f + r 2 f t f r f f

If f t f f

f = ft fr f

f ( N 2 - N1 ) N 2 +( N1 - N 3 ) N 3 .

0 78, this may be rewritten as

( N 2 - N 3 ) N 3 cot

rf r + r 2 f r f = f r f

( N 2 - N1 ) N 2 +( N1 - N 3 ) N 3

However, we may divide by frf cot and separate the equation into
[ rf r + r 2 f r ]
f
=
cot f
k fr

(4.23)

where
K=

( N 2 - N1 ) N 2 +( N1 - N 3 ) N 3
.
( N2 - N3 )N3

The left-hand side of Eqn. (4.23) is a function of r only, while the


right-hand side is a function of only. Therefore, Eqn. (4.23) must be a
constant. We can then write
[ rf r + r 2 f r ]
f k
=
=
cot f
fr

k=

(4.24)

where the constant n depends upon the set of quantum numbers, N1, N2,
and N3, for the particle.
Thus, n depends on the particle under
consideration.
The radial equation in Eqn. (4.24) may be integrated immediately
with the result
f R=

k e
r

(4.25)

The appearance of this exponential functional form for the radial


dependence is surprising and, at first, pleasing. The surprise is that this
functional form comes only from the gauge function, f, playing the guiding
role in the gauge fields and the field equation
B=0 ,

which is a purely classical equation.


radial function does not appear
dimensionality, but only upon the
four-dimensional approach would
function.

Thus, the exponential neo-coulombic


at first to depend upon the fifth
quantum condition so that even a
have produced this same radial

124

It is pleasing to see the appearance of the exponential neo-coulombic


radial functional of Eqn. (4.25) because the electron catastrophe has
haunted theoreticians since the inverse radial dependence of the columbic
potential was first seen. The radial function in Eqn. (4.25) is well behaved
everywhere; as r
, fr
0 and as r
0, fr
0. A quick glance at the
function might cause one to think it is the Yukawa potential, but a closer
look will show that the exponent is the inverse of the exponent in the
Yukawa potential.
The value for f may be obtained by integrating the remaining portion
of Eqn. (4.24). This integration produces
f = k 3 ( sin

If the exponential neo-coulombic function corresponds to reality,


then n must be small, less than 10-17m for electrons, and of the order of
magnitude of 10-15m for protons. Therefore, 0 must be very small, which
in turn implies f is very close to a constant.
The equations resulting from substituting the quantized potentials
into the remaining non-source equations [3.15b], [3.15e], and [3.15f],
produce the following restrictions
[3.15b]

( N2 - N3 )N3 f

f =0

( N3 - N1 ) N3 f

f =0

( N1 - N2 )N2 f

f =0

( N2 - N3 )N3 f t f =0
[3.15e]

( N3 - N1 )N3 f t f =0
( N1 - N 2 ) N 2 f t f = 0

[3.15f]
(satisfied identically).
When the potentials are substituted into the equations with source
terms, Eqn.s [3.15c-e] and [3.15h], the resulting equations are very
complex. To reduce the complexity of the equations, the assumption was
made that all source terms were zero; that is,
= 0; J = 0 ; J 4 = 0 .

This assumption reduced the complexity somewhat but still left a system of
equations that, thus far, is unsolved. However, an interesting aspect of
this assumption is the possible existence of a radial electric field without
the presence of any electric charge within, or upon, the particle. This
possibility rests upon the pressure of the term V 4 / 86 in the Eqn. [3.15d].
Much was learned about the interaction of charged particles by
considering only the radial dependence of the electric field while
temporarily neglecting the magnetic field or any potential variation of the
electric field with azimuthal angles.
This latter is the spherically
symmetric field assumption. Having not yet obtained a complete solution
to the system of equations that is the result of substituting the quantized
125

fields into the eight field equations, it proved beneficial to make the
assumption of spherically symmetric fields in which the only variation of
the fields is the radial dependence specified by the neo-coulombic radial
function.
Then, if we want to explore the radial dependence of static forces
between the fundamental particles allowed by the quantum condition, we
must consider the force law,
1
F jk J k ,
c

K j=

whose spatial components may be written


1
J
K = E + ( J x B )+ 4 V .
c
c

By restricting our concern to static forces we can concentrate on the force


density,
J4
V .
c

K= E+

(4.26)

Thus, the radial dependence of the electric field, E, and the V field are all
that need to be considered at the moment. Substituting the radial
function, Eqn. (4.25), into the field expressions, Eqn. (4.22), we find
Er =

Zk
r

1-

(4.27)

and
Vr=

Wg
r

1-

where Z = (N1-N0) and W = (N1-N4) so that the quantum number Z appears


in the radial electric field the same as it does classically.
From Eqn. (4.27), the electric field of fundamental particles allowed
by the Dynamic Theory is quantized by the quantum condition, and the
quantum steps may only be integer steps. This would necessarily preclude
any particles with fractional charge steps.
Substituting the radial fields of Eqn. (4.27) into the force law, Eqn.
(4.26), and integrating the charge density over the physical extent of the
particle, we find the radial force between two particles is
F=

q1 Zk
r

1-

g 2 Wg
r2

1-

(4.28)

where
q1 =

d(Vol) , g 2 =

J4
d(Vol) .
c

If we consider the electric force in Eqn. (4.28) and restrict our attention to r
such that r >> n, then the electric force becomes the columbic force
126

FE=

q1 Zk
r

Further, the other force term, based upon the V field, may be seen to also
have the same long-range form,
FV =

g 2 Wg
r

Thus, this force is also an inverse square long range force.


Therefore, the nonelectric force in Eqn. (4.28) cannot be a nuclear force.
What then must be the interpretation of this force, or must its appearance
be interpreted to mean that nature cannot have a five-dimensional
character?
The only force known in nature that has a long-range character, in
addition to the electrostatic force, is the gravitational force. But how can
we interpret the V field as the gravitational field when Einstein showed that
the gravitational field could be explained by a vector curvature in a
four-dimensional manifold and the V field is a gauge field in a
five-dimensional manifold? If the V field were to be considered as
gravitational, then the bending of light around the sun, predicted by
Einstein's General Theory of Relativity, must have another explanation. Is
this possible? If the V field is gravitational, then is there room in the
Dynamic Theory for an explanation of nuclear phenomena or must it also
follow the current approach to nuclear physics, thereby requiring similar
additional assumptions?
These and other questions occur when we see the long-range
character of the V field force. It is this theoretical quandary, presented by
the V field, that spoils the pleasant surprise of seeing a non-singular
electrostatic field emerge from the quantum condition. A number of
possibilities appear dependent upon the answers obtained to the previous
questions and/or others. Primarily, the possibilities pertain to the validity
of the five-dimensional view, for it is from this five-dimensionality that the
V field comes. One reasonable approach, in attempting to find a possible
way out of this quandary, seems to be to suppose the gravitational
interpretation is a possibility and then see how the Dynamic Theory
compares with measurable experimental evidence.
4.10 Self-Energy of Charged Particles.
One of the difficulties in Maxwellian electromagnetism is the infinite
self-energy that is predicted for a charged particle.
This "electron
catastrophe," or singularity, does not exist with the non-singular neocoulombic field, and the self-energy of a charged particle may be found.
In classical electromagnetic theory the self-energy of a charged
particle is discussed but its value has not been established. This is
127

because the expression for the self-energy is a function of the radius


associated with the physical extent of the charge distribution. Thus, the
radius of the charged particle must be known before the self-energy can be
determined.
Currently the self-energy of a charged particle is equated with the
energy associated with its inertial mass by E = mc2. Then the radius
associated with its energy is taken as the "radius" of the particle. There is
no intention that this radius be the physical radius of the particle though it
compares favorably with experimental values.
The question arises here of whether or not the Dynamic Theory, with
the five-dimensional viewpoint, can theoretically predict the self-energy
and/or the radius of the physical extent of the mass or charge distribution
of the particle. One of the beneficial aspects of the generalization of
physical theory as done in the Dynamic Theory is the possibility of using
conceptualizations and procedures developed in one branch of physics in
another branch. This aspect of the theory appears applicable here. The
self-energy of a charged particle is the notion that a certain amount of
energy be associated with the existence of the particle and its charge. This
notion
may be associated with the notion of free energy used in
thermodynamics, for, if the self-energy of the charged particle is its free
energy, then it represents the energy which may be "freed" upon converting
the particle into energy. Conversely, this would represent the energy
required to assemble the charged particle.
With the conceptualization of free energy the second law provides the
condition for a stable equilibrium state, namely that a charged particle in
an equilibrium state must exist at a minimum of its free energy. Thus, if
the self-energy, or free energy, of a charged particle is sought, then
minimizing its free energy will yield the desired result.
The free energy was defined, in analogy with the thermodynamic
case, as
G U- S-x F ,

(4.29)

where depends upon the applicable work terms which here will be taken
as the three spatial dimensions, so that = 1, 2, 3. The first law is given
by
d E = dU - F dx

while the second law yields


dS = dU - F dx

for a quasi-static, reversible process. Therefore, the differential change in


the system energy is
dU = dS + F dx .

128

(4.30)

Differentiating Eqn (4.29) gives the differential change in the free


energy as
(4.31)

dG = dU - dS - Sd - F dx - x dF .

Substituting Eqn (4.30) into (4.31) yields


(4.32)

dG = - Sd - x dF .

The force in Eqn (4.32) is considered to be the Lorentz force


F = q[ E + ( vx B ) ]

so that Eqn (3.48) becomes


dG = - Sd - x d{ q[ E + ( vx B ) ] }.

If we wish to consider the change in free energy with respect to a


change in the charge at a constant velocity, we find that, since
is a
function of velocity only, d = 0. The specification of constant velocity
stems from the desire to obtain the self-energy of a charged particle;
therefore, the particle should be considered as sitting still, so that it will
have no kinetic energy. The differential change of free energy for a
stationary particle is then
dG = - Sd - x d{q[ E + ( vx B ) ] },
= - Sd - x {dq[ E + ( vx B ) ] + qd[ E + [( vx B ) ] }.

so that for

= constant
G
q

= - x [ E + ( vx B ) ] - x q

[ E + ( vx B ) ]
q

but
E + ( vx B )

is independent of the charge q and, therefore,


G
q

= - x [ E + ( vx B ) ] .

If the charge is not in motion, then


G
q

129

=- x E

(4.33)

since v = 0.
If G is the self-energy of a charged particle, then by Eqn. (4.32)

G
= - r Er ,
q

the change in the self-energy is given with respect to a change in the


charge q. If we assume a spherically symmetric charge density, , then
dq = dv = 4 r 2 dr .

We may then find the free energy by the integration


q

dG = - r E r dq
q0

G0
R

= - 4 r 3 E r dr ,
0

where R represents the radius within which the charge density


contained.
The field Eqn. [3.15d] is
( E )= 4

- a0

( V4)

is

The entire right-hand side of this field equation behaves as a charge


density; therefore, we may equally write
( E )= 4

(4.34)

where it is understood that either ( V 4 )/ 109 is zero or is considered to be


a total effective charge density. In either event, Eqn. (4.34) gives us
r

( r2 Er )
=4
r

when we consider only a radially symmetric field Er. Thus,


dr =

r r2

d( r 2 E r ) .

(4.35)

Substituting Eqn. (4.35) into Eqn. (4.33) the self-energy is then found to be
G=-

(r E r )d( r 2 E r ) + G 0 .

If we now use the neo-coulombic electric field given by


130

(4.36)

Er =

e
4

1-

in the integral of Eqn. (4.36), we find


R

G=- e
0

e
1e
4 r
r

1-

+ G0 ,

which may be integrated so that


G=

- e2
2

2(4 )

3
4 R

1
1
- e2 R 4

2
R

+ G0 .

(4.37)

To find the specific value of the self-energy, we must find the R that
minimizes G. Therefore, set
G
=0 .
R

After carrying out the required differentiation and simplifying, this is


satisfied if
2
R +

3
5

R-

2
5

=0 .

(4.38)

Equation (4.38) only has one positive root, which is


R = 0.4

Substituting this result into Eqn. (4.37), the self-energy becomes


G=

- e2
(4 )2

(0.063379) + G 0 ,

or
G=

- 7.26235 x 10 -3 MeV - fermi


+ G0 ,
(fermi)

(4.39)

when

is given in units of fermi.


An example may be a proton for which is approximately 1 fermi if
the proton-proton scattering data is considered. Then if ~fermi,
Gp = -7.26235 x 10-3 MeV + Gop ,
so that
Gop = 938.263 MeV

(4.40)
131

is the part of the proton rest energy independent of its charge. The charge
energy of the proton would then be
Gcp = -7.26235 keV ,
which is negligibly small compared to the non-charge energy Gop.
What is the nature of the energy Gop? It is not energy caused by the
presence of electric charge on the protons. Also the self-energy, G, was
found for a resting particle. If we associate the resting self-energy, G, with
the rest mass as
Gp = mopc2 = Gcp + Gop ,
and Gcp is the portion of the proton's rest energy that is due to its charge,
then Gop must be that portion of the rest energy that is due to the proton
mass above. In this case, the proton mass energy, Gop, is given by Eqn.
(4.40).
Suppose we consider an electron and assume that e - ~ 10-3 fermi.
Then
Ge- = 0.511 MeV = -7.26235 MeV + G0e-,
or the mass energy of the electron would then be
Goe- = 7.773 MeV ,
whereas its charge energy is
Gce- = -7.26235 MeV
4.11 Nuclear Phenomena.
The electrostatic force, appearing in Eqn. (4.27), differs significantly
from the columbic force only when r becomes small enough to be of the
order of magnitude of the n. The first experimental evidence that the
scattering of charged particles by other charged particles was not always
columbic was the Rutherford scattering data. The appearance of the
exponential multiplier in the neo-coulombic force of Eqn. (4.27) prompts us
to ask whether or not the difference between this force and the columbic
force suffices to explain nuclear phenomena without resorting to the
postulation of a new short-range force such as the nuclear force.
An obvious starting point to explore the possibility that the neocoulombic force might apply to nuclear phenomena would probably be the
132

Rutherford scattering formula.


This may be done; however, the
appearance of the exponential term makes an analytical expression
difficult, if not impossible, to obtain. We may arrive at a solution of limited
usefulness if we assume that r >> . Further, in considering particle
scattering, we shall restrict our consideration to scattering of like particles,
only, so we are guaranteed that only one is involved.
The best way to investigate the scattering cross sections is to start
with the solutions of the equations of motion for planetary orbits in which
the force is given by the neo-coulombic force instead of the simple inverse
r2 force from Newton's gravitational force. We find the radial equation
becomes
2
Mk
d u
+ u = 2 (1 - u) e- u ,
d 2
L

where u = 1/r, k is the gravitational constant, and L is the orbital angular


momentum. The exponential function may be expressed in terms of a
power series, and our radial equation becomes
2
(- u )2
(- u )n
Mk
d u
+
u
=
(1
u)
1
+
(u)
+
+
...
+
+ ...
2
d 2
2!
n!
L

Mk
L

(n + 1)
1 + (n=1)
(- u )n
n!

by assuming r>> so that /r = u <<1, then we may neglect the terms


with n 3 in our radial equation. The result of this assumption is
2
d u
+
d 2

u=

Mk
L

Mk

2 L2

with
2

= 1+

2 Mk
L

This equation may be compared with the classical equation,


2
Mk
d u
+u= 2 ,
2
d
L

or with the general relativistic equation6,


2
Mk
d u
+ u = 2 + 3 Mku 2 ,
d 2
L

which has the identical form of our equation. Thus, the same method of
perturbations may be used to obtain a solution as was used for the
relativistic case. The result of this calculation is the solution,
133

1 Mk
u = = 2 1 + e cos(
r L

(4.41)

),

where
0

3 Mk 2
2 L2

is the increase in the perihelion.


Notice we have shown that the neo-coulombic force will predict an
advance in the perihelion of planetary orbits with the solutions to our
planetary orbits equation. We will discuss this further at a later time. We
needed the solution given by Eqn (4.41) in order to obtain an expression for
the scattering cross section of like particles.
If we now consider the solution, Eqn. (4.41), obtained with the assumption that r >> , then the scattering cross section may be expressed
as

d =

q1 q 2
2 mv02

2 sin d
sin

(4.42)

where
1+ 6
3
1+
2

E
k

sin
4 E
k

1 + 2(

- )

tan

2
4

sin

sin ( - )

The appearance of the factor expresses the first-order deviation of


the scattering cross section of the neo-coulombic force from that of the
columbic force. However, the assumption that r >> implies a limit on the
minimum impact parameter for which this cross section retains validity.
Therefore, a computer solution is probably necessary to really investigate
the scattering of charged particles using the neo-coulombic force.
Figure 6. Comparison of coulomb and neo-coulomb forces at short range.
Another way of visualizing the neo-coulombic force is to make a plot
of it and compare it with a plot of the columbic force. Figure 6 compares
these two forces plotted with the separation variable in fermions and
normalized so that the columbic force at one-fermion separation is unity.
Note that this plot compares the forces for like particles to ensure that is
the same for both particles. Figure 6 shows that the neo-coulombic force
134

is virtually indistinguishable from the coulomb force for separations


greater than approximately 10 . However, at a separation of exactly , the
force is identically zero. In terms of the classical notion of nuclear forces,
we would say that at separations greater than 10 , the nuclear force is
negligible, whereas at a separation of the magnitude of the nuclear force
was equal to the magnitude of the coulomb force. The neo-coulombic force
becomes an attractive force for separations less than . This is exactly the
behavior to be expected of a non-singular potential. For a potential to be
non-singular it must tend to zero as r goes to zero. Such a potential which
tends to zero for r tending to zero and for r tending to
must have a
maximum absolute value in between. At that maximum the force, being
determined by the slope of the potential, will go to zero and will be of the
opposite signs on each side of the zero.
Now let us look at the force between unlike particles, say a proton
and an electron. Consider the electron and proton to be placed on a
horizontal surface separated by a distance, r, with the proton to the right of
the electron. Thus, the long-range attractive forces between these two
particles will cause the proton to experience a force to the left while the
electron will experience a force to the right. We may than write the force
on the proton that is due to the positive charge of the proton being in the
electron field as
F p = q p Ee
=

-k
r

1-

( u x ) ,

(4.43)

where the electron field involving the electron lambda has been accounted
for. The electron force owing to the electron charge being in the proton
field is given by
F e = qe E p
=

k
r

1-

p
r

( u x ) .

(4.44)

Figure 7 plots both these forces as a function of the separation, r, where, p


= 10-15m, or p = 1 fermi has been assumed. The electron-electron
scattering data show that the electron-electron interaction behaves in a
coulombic manner even when separations are approximately 0.01-0.1 fm.
To be consistent with this data, we have assumed e = 10-3 fermi.
From this plot of the force on the proton and the force on the
electron, we see that for separations less than about 10 fermis the forces
become extremely unsymmetrical.
This immediately and visually
demonstrates that the neo-coulombic exponential force violates Newton's
135

third law requiring that the force on the proton be equal in magnitude and
opposite indirection to the force on the electron. The question arises
whether or not a violation of Newton's third law has ever been seen as the
result of an interaction between an electron and a proton? The answer,
based on a neutron disintegration from which a proton and electron
emerge, is definitely yes; Newton's third law was seen to be violated. To
reinstate Newton's third law in neutron disintegration and all other beta
decay, Pauli postulated the existence of the neutrino.
Fermi later
developed his theory of weak interactions,11 from which appeared the
necessity to talk of a fourth force in nature.
Can it be that the neo-coulombic force, which requires distinct
for
distinct fundamental particles, accounts for the action of the weak forces
also? The possibility that it might opens the theoretical flood gates and a
virtual tidal wave of questions surges forth. Does this mean the neutrino
does not exist? What about the experimental evidence submitted in
support of the capture of a free neutrino?12 Could this mean that the
neutron might be bound states of an electron and proton? This question
should be followed by, what about conservation of angular momentum in
neutron decay (i.e., spin), conservation of linear momentum, and
Heisenberg's uncertainty principle?
Figure 7. Neo-coulombic forces between unlike particles at short range.
The preceding questions do not begin to scratch the surface of the
theoretical questions that need to be answered as the result of considering
the possibility that the force law of Eqn. (4.27) with only a gravitational
force plus the neo-coulombic force might explain the phenomena now
thought to require four distinct forces in their explanation. However, the
appearance of a non-singular force with the apparent range of the neocoulombic form cannot be thrown out offhand. Therefore, it seems that the
only reasonable choice is to systematically and thoroughly explore the possibilities.
If we again consider the plots of the proton and electron forces in
Fig. 7. we see that, at atomic separations and greater distances, the forces
obey Newton's third law and the difference between the neo-coulombic and
columbic forces is so small that it could not be detected in atomic or
macroscopic phenomena. But as the separation becomes smaller, the
picture begins to change. When the r approaches p, the electron is no
longer attracted to the proton as strongly as the proton is attracted to the
electron. If the separation is exactly p, then the electron is indifferent to
the proton's presence. The proton, on the other hand, is still very much
attracted to the electron. If for the moment, we ignore the interpretation of
Heisenberg's uncertainty principle that would say it cannot be, then we
could easily imagine a circular proton orbit around a stationary electron,
during which the proton stays at a radius of p from the electron. The
136

electron should be stationary during such motion because it would


experience no force.
We now consider a separation between the electron and proton,
which is some simple fraction of p. Here, we find the electron repulsed by
the proton, but the proton is still attracted to the electron. Notice that the
force on both particles, from our initial positioning of the proton on the
right, is to the left. If both particles were given an angular momentum
such that they were placed into synchronized circular orbits, then because
their synchronous motion always results in the force on both particles
being directed along the line separating them and from the proton toward
the electron or from the electron away from the proton then, again ignoring
arguments from the uncertainty principle, circular orbits in which the
electron is in a small orbit about a space point could be imagined, where
the proton is in a much larger orbit about the same space point.
Let us follow this picture a little farther and write simple
Newtonian-like force laws for this situation. The situation envisioned is
presented in Fig. 3. The electron position is given by re from the origin,
and the position of the proton is given by rp. The separation between them
is
r = rp - re .

(4.45)

Because the force is always directed along the line separating the two
particles, we may write the radial equation of motion for the proton as
mp

vp
rp

-k
r

1-

e
r

(4.46)

where the assumed circular motion has been taken into account and vp is
the tangential proton velocity. The electron equation of motion is given by
me

k
p
ve
= 2 1e
r
re
r

(4.47)

Figure 8. Electron and proton orbits.


In both equations k < 0.
The right-hand side of Eqn. (4.46) and (4.47) are both functions of
the separation, r, whereas the two left-hand sides are individually
functions of re and rp. A solution is possible only when the three
equations, Eqn.s (4.45)-(4.47), are solved simultaneously.
An alternative approach is to add Eqn. (4.46) and (4.47) to obtain the
equation of motion for the center of mass,
mp

vp 2
rp

+ me

p
ve 2 k
= 2 1e
r
re r

137

- 1-

(4.48)

or
2
k
p
MV
= 2 1e
R
r
r

- 1-

where R = (mprp + mere)/(mp + me) and M = mp + me. From Eqn. (4.48) we


see that bound states, where the center of mass is in motion as the result
of the asymmetrical force, may only occur when the separation is less than
p.
All of these equations of motion exhibit a feature not usually found
in equations of motion. That is, because the force depends on the
separation, r, between the particles and not strictly on the position, re, for
the election, then the usual integration of the force over a change of
position, which produces the potential energy, cannot be readily done
because,
V( r e ) = - F e dr e
=-

K
r

1-

(4.49)

dr e .

However, Eqn. (4.45) and Eqn. (4.47) may be used to obtain re as a


function of r, or vice versa, so the integration of Eqn. (4.49) may be
completed. The transcendental function in the force law prohibits an
analytical solution of re as a function of r. Therefore, only numerical or
graphic solutions of these equations are possible.
4.12 Heisenberg's Uncertainty Principle and Geometry.
The suggestion that bound states of electrons and protons might
exist where the orbits are of the approximate order of magnitude of nuclear
dimensions, is essentially a return to the notion that a neutron might be
such a state. This idea gave way under arguments of conservation of
momentum and Heisenberg's Uncertainty Principle to the view that
electrons are forbidden to be found within the nucleus. Therefore, let us
take another look at those fundamental tenets of quantum mechanics, the
Poisson brackets.
The classical Poisson bracket is defined by
F,G =
j

F
qj

G
F G
pj
pj qj

where F and G are any two functions of the canonically conjugate variables
qj and pj. The special relations that occur when F and G are qj and pj,
respectively, are especially important in quantum mechanics; these are,
classically:
138

{qj, qk} = 0
{pj, pk} = 0
{qj, pk} = 0 ,

(4.50)

where jk is the Kronecker delta. The classical Poisson brackets of Eqn.


(4.50) are obtained when Euclidean spaces are assumed. However, the
definition of Poisson brackets remains valid for general metric spaces,
when the notion of covariant differentiation is used. If we now consider the
momenta, expressed in a general coordinate system, the covariant
components,
p j = mg ij x k

and

(4.51)
jl

p j = g pl = m x j

are the contravariant components. Covariant differentiation must be


carried out with respect to contravariant vector components. There, in a
general space the canonically conjugate variables to be considered are xj
and pk, and the Poisson bracket of the position and momenta becomes
k
x ,p =
j

j
j 2
x
+
x
l
sl
x

j
p
x
l
l
p
p

j s
x
jl +
sl

lk

k
n
p
nl

(4.52)

or
k
j
x ,p =

jk

j
s
x .
sk

Quantum mechanics adopts the operator,


_
i

pj ,

for the momentum. This, in general case, becomes the covariant operator
_
i

,j

pj .

The operator for the contravariant momentum components is then


139

(4.53)

_
jl
g
i

,l

(4.54)

Now if we look at the quantum Poisson bracket, where the operators are
operating on a scalar , then
[ x j , pk ]
= xj

_ kl
_ kl j
g
g (x
l
i
i
x

= xj

j
j s
x
+
x
l
s
l
x

_
_ kl
- g kl
g
l
i
i
x

_ jl j
g x
l
i
x
= i g kl

jl

j s
x
sl

), l

(4.55)

This may be written in terms of the classical Poisson bracket, Eqn. (4.52),
as
k
j
x ,p

= i _ g kl x j , p l

If the space is Euclidean, then the gkl become the Kronecker delta
and the Christoffel symbols vanish and the quantum Poisson bracket of
Eqn. (4.55) becomes
j
x , pk

= i_

jk

because pk = gkl pl = klpl = pk. However, from Eqns. (4.53) and (4.54), we
see that the metric does play a role in the quantum operators. This should
also be seen in the use of the operators in the Schrodinger Hamiltonian
operator, because
j
pj p =

_ 1
i g
=

_ 1
i g
x

g pj

_
jl
g
i

g g jl

(4.56)

becomes the operator to be used in a general space and, of course, is the


operator used in applying Schrodinger's equation to the hydrogen atom.
140

The geometrical effect may be seen also in Dirac's equation by


considering that the restrictions,
p )2 = p j p j ,

=1 ,

(4.57)

and
+

=0 ,

must be met in order for solutions of


o
p |> = H |> ,

where
H =-

p+ m ,

is also a solution of pjpj = m2 in natural units.


The first restriction may be rewritten as
(

j 2
j
jk
j k
p ) = p i p = m 2 g jk x j x k = g p p

by the definition of the momenta. Then, if we expand the left-hand side


and equate coefficients of the pjpk, we find that
( )2 = -g11 ,
( )2 = -g22 ,
( )2 = -g33 ,
1

= - 2 g 12 ,

= - 2 g 13 ,

= - 2 g 23 .

and

From Eqn. (3.77), for a Euclidian metric where gjk = jk, these restrictions
reduce to the usual restrictions. Any metric properties will affect these
restrictions and will therefore feed into the solutions.
Now, of what benefit it this discussion of geometrical effect upon
quantum mechanics in considering the neo-coulombic force? Recall that
the neo-coulombic force came from a gauge function in a Weyl space. A
gauge function has a geometrical effect that could be thought of as
effectively changing the unit of action in quantum mechanics. To see the
basis for this statement, let us recall the quantum Poisson bracket
operations on a scalar,

141

(3.77)

k
j
x ,p

= i _ g kl

j s
x
sl

jl

(3.74)

and let us define


_

jk

= _ g kl

jl

j s
x
sl

(3.78)

then we can write


k
j
x ,p

=i _

jk

which has the same form now used but the effective unit of action _ 158
depends on the geometry as seen by Eqn. (3.78).
We may look at the effective unit of action in yet another way.
Recall, from the principle of maximum entropy, that the generalized
entropy is the action. Thus, quantization of the action is a quantization of
the generalized entropy. But, because the entropy space is tied to the
sigma space, we have
dq

0 2

2
) .

= f(d

The gauge function is a function of the space point; therefore, the gauge
function varies continuously from point to point in the space. Thus, if the
generalized entropy is quantized, so must be . We may write
0
q = n_ = = n_ ,

where the difference between _ 161 and _ 162 contains the geometrical difference
between q0 and . But how can we determine the relationship between _ 163
and _ 164?
From the principle of maximum entropy, the potential energy
function was defined as the negative integral of the force through a
distance just as it is in classical mechanics, because
F j=-

dV
dx

but the potential energy plays the role of the gauge function. The
equations of motion for the asymmetrical forces between unlike forces
showed that an analytic form for the potential energy may be unobtainable
owing to the transcendental nature of the forces and because the forces
depend upon the separation between the particles, not their positions.
Thus, there appears no way, at the moment to obtain an analytical
expression for _ 166, and we must resort to a numerical solution for the unlike particle
case.
142

(3.79)

The absence of an analytical expression for the effective unit of


action, _ 313, does not completely stop us from considering the possibility that a neutron
may be a proton in a large orbit about an electron in a small orbit. We may, for the moment,
acknowledge the difficulty of obtaining an analytical expression for _ 314 by allowing the
_ 315, or unit of action, for the proton and electron to be a function of their orbit, and we
may designate _ e 316 to be the effective unit of action for the electron orbit in a neutron and
_ p 317 to be the unit of action for the neutron's proton orbit. If the effective unit of action
depends upon the orbit, as it appears here that it must, then the interpretation that
Heisenberg's Uncertainty Principle rules out the possibility of an electron being contained
within nuclear discussions is inapplicable.
Another argument against the neutron being an electron and proton
in nuclear-sized orbits is based on an argument that the principle of
angular momentum cannot be conserved. The neo-coulombic forces,
which require that the force between the electron and proton be directed
on a line between them, requires that angular momentum be conserved.
However, the effective unit of action for the electron orbit requires that, in
the neutron the orbital angular momentum would be given by _ e 318 and its
1
_ 319. Similarly, for the proton the orbital
2 e
1
angular momentum would be _ p 320 and the spin ( ) _ p 321.
2

intrinsic spin angular momentum would be

After the neutron decays, the angular momentum is the sum of the
two particles' intrinsic spin angular momenta, which is given by _ 322
because both particles are free and, therefore, each has an intrinsic spin angular momentum
1
2

of ( )_ 323. Therefore, the conservation of angular momentum is expressed as


1
+ _ _e + _ _ p + _e + _ p = _ .
2

(3.80)

Experimental evidence of orbital and/or spin angular momentum is


contained in the experimental magnetic moments. If we equate the
intrinsic and orbital magnetic moments of the electron and proton while
they are in the orbital configuration to the experimental value of the
neutron's magnetic moment we have
+_

where

2
2

_e
_

+_

_p

2
2

326 is a Bohr magneton and

_
n

_e
_

_p
_

= - 1.91315

327 is a nuclear magneton.

Equations (3.80) and (3.81) represent two equations in the two


unknowns, ne 328 and _ p 329, which may be solved to obtain the effective units of action
for the electron and proton orbits making up a neutron such that angular momentum in
conserved during neutrons' decay and that the correct magnetic moment of the neutron is
ensured. Substituting the experimentally measured values of intrinsic magnetic moments
143

(3

for the electron and proton into Eqn. (3.81) produces a more accurate solution because this
contains the anomalous magnetic moments. Then we would have
+ _ 2.002319

_e
_

+ _ 2.79275

_p
h

_e
_

_e
_

= - 1.91315

(3.82)

The only simultaneous solution of Eqns. (3.80) and (3.82), for which
_ e 331 and _ p 332 are both positive, are
_ e = 8.0517 x 10 -4 _ ,
p
_ = 0.66586 _ .

The values of the effective units of action for the proton and electron
given in Eqn. (3.83) show that angular momentum is conserved during the
decay of a neutron when the neutron is considered to be a proton in orbit
around an electron under the neo-coulombic force.
The third major argument against a neutron being a state of electron
and a proton orbits stems from the experimental evidence on the violation
of Newton's Third Law during decay. That is, the energy of the electron
emerging after decay is inconsistent with the equal and opposite columbic
forces between an electron and a proton. Here, we find that the neocoulombic forces are unequal in magnitude and opposite in direction; thus
the energy of an electron emerging as the result of crossing from such an
orbit cannot be consistent with Newton's third law. There now exists a
fourth argument against this picture of a neutron: the possible existence
of the neutrino. The above picture of the neutron produces no need to
postulate the existence of neutrinos. What then can be said about the
experimental evidence that has been put forward in support of the capture
of free neutrinos?12 A conclusive answer will need to await further
investigation.
4.13 Nuclear Masses
The difficulty produced by the asymmetry of forces that arises in the
interaction of an electron with a proton may be avoided if two protons are
considered to be in orbit about the single electron. If we think of a
snapshot of such a case we would find that the situation depicted in Fig. 9
allows us to visualize the forces.
Figure 9. Two protons in orbit about a single electron.
The force on the electron would be zero because it has a proton on
each side diametrically opposed to one another. The force on each proton
will be made up of two parts; one, the force that is due to the presence of
the electron, and the other, owing to the other proton. The symmetry
144

(3.83)

guarantees that each proton will experience an identical force, if circular


orbits are assumed, toward the center of rotation. The force on the proton
on the left would be
F p1 =

-|k |
r

1-

|k |
1 - e e(2r)
(2r )2

(3.83)

2r

To be sure, quantum mechanical procedure should be used;


however, it may be beneficial to begin by assuming circular orbits similar
to Bohr's initial approach to atomic structure. This may indicate the
potential utility of the force in Eqn. (3.83), as well as perhaps identifying
procedures to be used later.
Any nuclear orbits should probably be relativistic; therefore, in
cylindrical coordinates, where the velocity for motion in a plane is given by
,

v = rr + r

then we have
= 1-

2
( r2 + r2
v
= 1=
2
2
c
c

For circular orbits, this becomes


= 1-

(3.84)

Thus, the relativistic equations of motion for the proton become


d m p (rr + r )
|k |
= 2
dt
r

1
p
1e
4
2r

2r

- 1-

(3.8

Equation (3.85) separates into two equations


d mp r
|k|
= 2
dt
r

1
p
1e
4
2r

2r

- 1-

r .

and
d mp r
dt

=0 .

The second of these equations says that the angular momentum is given
by

145

(3.86)

m r

(3.87)

= L p = n_ ,

where _ 342 indicates that whereas the unit of angular momentum will be a constant for a
given orbit, it may be different for different orbits.
The first of Eqn. (3.86) is
( mp r - M p r
d mp r
=
dt

d
) m p r dt
|k |
= 2
2
r

1
p
1e
4
2r

2r

- 1-

but for circular motion r = 0 344, therefore,


mp r

-|k |
r

1
p
1e
4
2r

- 1-

2r

(3.88)

Substituting from Eqn. (3.87) into Eqn. (3.88) we have


2
2
-|k |
n ( )
=
3
2
mp r
r

1
p
1e
4
2r

2r

- 1-

(3.89)

The potential energy for one of the protons can be found by integrating the
force and is
V(r) = - . F(r) dr = - | k |

1
r

1
p
1e
4
2r

|k|
=
r

1 e
4

-e

2r

2r

- 1-

dr

(3.90)

Then the total energy of the three-body system, including rest energy,
would be
ET =

2|k |
r

1 e
4

2r

- e-

2 m p c2

+ me c 2 .

However, by substituting Eqn. (3.87) into Eqn. (3.84) and solving for , we
find

146

(3.91)

=
1+

n_

(3.92)

m p rc

Thus, substituting Eqn. (3.92) into Eqn. (3.89) produces a


transcendental equation whose solution gives r(n), which may then be used
in Eqn. (3.90) to obtain the total energy of the system. The mass of the
system should then be found from
M=

ET
2
c

Because this system has one electron and two protons, it has a total
electric charge of +1 and would have a mass of approximately 2 amu. This
is the same characteristic exhibited by the deuterium nucleus. If this is
the structure of the H2 nucleus, then the mass given by Eqn. (3.93) should
correspond to the mass of the ground-state nuclear mass for n = 1. If the
1e-, 2p+ case existing where n = 1 is the ground-state H2 nucleus, then is
the excited state represented by two protons in the n = 2 state or can it be
represented by one proton in an n = 1 orbit and one in an n = 2 orbit? The
equations developed here consider only the case when both protons are in
the same orbit. Any consideration of the protons being in different orbits
introduces an asymmetry in the forces and a similar difficulty faced in the
neutron case. Therefore, for the moment we will consider only the simpler
cases, where symmetry reduces the complexity of the solution. Notice,
though, that even in the simpler symmetric case, no analytical solution
exists of Eqn. (3.89) for r(n) because the force contains a transcendental
function.
By allowing 351to be different for each n, then for the ground and first excited
states of the H2 there are four quantities to be determined: p , e , _ (1) 352, and
_ 353. Of course, in theory, we could determine both p 354 and e 355 from scattering
experiments: then we would only have two, _ (1) 356and _ (2) 357. But if we discover
exactly how the _ 358 depends upon the orbit, then a solution of Eqn. (3.89) would
represent a pure theoretical prediction of both the ground state mass, m(n = 1), and the
excited state mass, m(n = 2), for then we would be able to express _ = _ (r) 359.
If we think about the possibility of adding an additional proton to the
1e-, 3p+ case, we are faced with a question. Can the additional proton be
placed in the n = 2 orbit without considering the asymmetry thus
introduced into the system, or must we consider a single orbit with three
protons symmetrically spaced? The answer lies partly in the solution of
the appropriate Schrodinger or Dirac equation, because this would inform
us of the number of protons that are allowed in a given orbit. This,
however, would not answer the question concerning how the asymmetry
introduced by a single proton in the n = 2 orbit affects the problem. To
147

(3.93)

obtain an answer to this question the situation should be addressed by


both methods and both results should be compared with the experimental
evidence.
Whichever approach proves to be correct, the result would be a
nucleus with a total charge of +2 and a mass number of 3, or Z = 2, and A
= 3. This corresponds to the He3 nuclei. If the third proton can be placed
in the n = 2 orbit of the solution for a single-electron case, and determining
p 360 and
e 361 by scattering experiments, plus determining _ (1) and _ (2) 362 by the
bound and excited states of the H2 nucleus, then the mass of the He3 is a pure
theoretical prediction. We shall consider this a little later.
The fact that the excited state of the H2 nucleus is a virtual state
implies that adding two protons may itself be nearly a virtual state and
thus a cutoff would occur in the number of protons that can exist in a
bound state for a one-electron core. To possibly look at other nuclei,
suppose first that we go back and consider the like particle force between
two electrons,
F ee =

|k |
r

1-

(3.94)

This force is repulsive until the separation between the electrons


becomes less than e, and then it becomes attractive. But can there be
bound states of two electrons? The answer lies in the effect of the gauge on
_ 364. If the _ 365 for unlike particles is less than _ 366, then the sign reversal between
the force of unlike particles and the force for like particles should mean that the _ 367 for
like particles should be larger than _ 368. Thus, bound states of like particles would be
forbidden.
As an example of the sign reversal effect on _ 369, suppose that the gauge
for the unlike electron, proton case is given by
1

lnf 2 =

-k
r

- e-

or
f = exp

-k
r

- e-

Then,
( dq0 )2 = g jk dx j d xk = f(d

2
) f g jk dx j dx k ,

or
g jk = f g jk .

148

(3.95)

Suppose g jk 374 =

jk 375;

then Eqn. (3.78) would become


_

jk

= _f g kl
= f_

jk

jl

+0

(3.96)

Then, from Eqn. (3.96), we find


_
=f .
_

Now e < < p 378; therefore, when r = p, the gauge function, f, from Eqn. (3.95)
is less than unity because k < 0. Thus the _ 379 given by this function is always
less than or equal to .
On the
other hand, for like particles where only one
is
involved,
the _ gauge function would be given by
f = exp

k e
r

(3.97)

Then, because k > 0, f 1, requiring _ _ 382.


Although like-particle bound states may be forbidden by the
uncertainty principle for large _ 383, bound states of unlike particles of nuclear
dimensions are allowed by an _ 384that may be much, much less than _ 385. Next, we
might consider possible bound states between electrons and positrons with subnuclear
dimensions. If the for the positron, e+, is less than the for the electron, e, then we could have bound states with two electrons in orbit about a
single positron that would be given by equations exactly like the equations
for the 1e-, 2p2+ states, where the positron replaces the electron in the 1e-,
2p+ equations and the two electrons in the 1e+, 2e- case replace the two
protons in the 1e-, 2p+ shell. It is possible now to consider a 1e+, 3e- core,
thus introducing questions concerning asymmetry aspects and other
possible questions. However, owing to the e + < e - << p, the core
structure is one where in electrons orbit about positrons and the electron
orbits are << p. For the protons in a shell orbit the interior core structure
may be negligible, just as the internal structure of the nucleus has almost
no effect on the atomic electron orbits.
We shall not explore the core structure here but shall consider only
the effect of different-core excess electron charge upon allowed proton shell
orbits. If we denote the excess electron charge of the core by the integer Y,
by which we mean the total number of core electrons less the number or
core positrons, then by denoting the number of shell protons in orbit
around this nuclear core, we find that the charge on the nucleus, Z, is
given by
Z = A-Y

149

(3.98)

Equation (3.98) indicates that the excess core electron number behaves
identically with the neutron numbers in current nuclear theory, although
there are no neutrons as such in this nuclear model. Indeed, the neutron,
in this picture, is simply another state, namely Y = 1 and A =1.
This suggests a picture of the nucleus in which there are protons in
orbits about a nuclear core. The number of protons are given by the
current mass number, A. The radii of the proton shell orbits are
approximately the value of p; that is, about 1 fermi. The core may be
made up of electrons in orbit about positrons and is sized approximately
the same as e-, which is much, much less than p. This view of the
nucleus is similar to that of the atomic view, but here the nuclear core
plays the role of the atomic electrons. The force law for the shell proton orbits would then be given, from Eqn. (3.85), by
F=

|k |
r

1
p
1e
4
2r

2r

-Y 1-

(3.99)

The equations specifying the proton shell orbits are


2
2
-|k |
n (h )
=
3
2
mp r
r

1
p
1e
4
2r

2r

-Y 1-

(3.100)

and
1

=
1+

n_
m p rc

The total energy of the nuclei would be given by


E(A, Y) =
n
-

Ye

R(n)

A(n) | k | 1 e
R(n) 4
+

A(n) m 2p
[R(n)

2R(n)

(3.101)

+ E c (Y)

where A(n) is the number of protons with the quantum number, n; R(n) is
the radius of the proton orbit with the number n; Ec(Y) is the energy of the
nuclear core for which Y is the excess electron charge; and [R(n)] is the
relativistic evaluated for R(n). The mass of the nuclei with energies given
by Eqn. (3.101) would then be
150

M(A, Y) =

1
c

E(A, Y) .

(3.102)

This approach has a simple look to it. For instance, if Y = 1 then Ec =


0.511 MeV, the rest energy of the electron. Then the ground state for 2H
would be
E(2, 1) = 2 E 1 + E c ,

whereas the excited state is


E(2, 1 )* = 2 E 2 + E c .

Using Ec = 0.511 Me V we find that the energy of a single proton in the n =


2 orbit would be
E2 =

E(2, 1 )* - E c
2

Table I. Experimental and predicted nuclear masses.


Experimental Predicted
M = Ep - EE
Predicted
MassMass
M/A
BE/A
Y Z A (MeV)
(MeV)
(MeV) (MeV) (MeV)
1 1 2
1875.0
1873.7
-1.3
-0.7
1 1 2
1877.9
1879.7
1.8
0.9
1 2 3
2808.3
2819.9
2.6
0.9
1 3 4
3749.5
3748.2
-1.3
-0.3
2 1 3
2808.9
2804.2
-4.7
-1.2
2 2 4
3727.3
3736.1
8.8
2.2
2 3 5
4667.5
4668.1
-4.8
0.1
2 4 6
5604.7
5600.0
-4.7
-0.8
3 2 5
4667.8
4668.3
0.5
0.1
3 3 6
5601.4
5600
-1.0
-0.2
3 4 7
6531.8
6532.3
-1.6
0.1
4 1 5
4691.8
4689.2
-1.6
-0.3
4 2 6
5605.5
5610.4
4.9
0.8
4 3 7
6533.3
6531.6
-1.6
-0.2
4 4 8
7454.3
7452.7
-1.6
-0.2
5 2 7
6545.7
6545.7
0.6
0.1
5 3 8
7471.2
7471.2
-1.5
-0.2
5 4 9
8392.2
8395.5
0.8
0.1
6 2 8
7482.5
7482.5
1.0
0.1
6 3 9
8406.7
8404.7
-2.0
-0.2
6 4 10 9325.0
9326.0
1.0
0.1

151

2.1
1.2
1.7
1.5
4.4
4.9
5.2
5.4
5.4
5.5
5.6
1.5
4.1
5.9
7.3
4.6
5.9
-3.8
5.3
6.5

Thus, the 3He nuclei energy would be


E(3, 1) = E(2, 1) +

E(2, 1 )* - E c
.
2

Now using the tabulated experimental data14 we find that the predicted
nuclear mass of the 3He should be
2.016000 - 5.49 x 10 -4
amu
2
= 3.020562 amu ,

E(3, 1) = 2.012836 amu +

compared to the tabulated value of 3.014848 amu. Similarly, the predicted


mass for 4Li should then be 4.028288 compared to the tabulated
4.025231. The difference between the predicted and the tabulated values
are 1.8 and 0.7 MeV/nuclear, respectively, for the 3He and 4Li nuclei.
Because the core energy and orbital energy levels should change
Table II. Energy-level average values.
Y
1
2
3
4
5
6

2E1+Ec
(amu)
2.011441
2.009952
2.009979
2.067271
2.071474
2.100007

2E1+6E2+Ec E1
(amu)
8.048629
8.012742
8.013231
8.000773
8.018938
8.032752

E2
(amu)
1.002976
------

(amu)
1.006198
1.000465
1.000542
0.988917
0.991244
0.988969

when the excess electron number of the core changes, we may construct
Table I, where selected nuclei are used to establish the core and shell
energy levels for different Y. Predictions of the mass of other nuclei are
made using the energy Eqn. (3.101) and assuming that the number of
protons in a full shell corresponds to the number of electrons in the atomic
shells, i.e., 2, 8, 18, .... For each Y, some of the experimental masses are
used to establish an energy value; therefore, the predicted value appears
the same as the experimental. In each case, the energy value established
by this data point appears in the appropriate column. The RMS error in
the predicted values of all 21 nuclei was 4.3 MeV, with an arbitrary
selection of which nuclei were used to establish an energy level.
A better way of approaching the establishment of the energies would
be to take the average value of all possible ways to find a particular energy.
Table II lists the energy-level average values needed in the total energy
equation for the same 21 nuclei. By using the average values from this
152

table in Eqn. (3.101) we may construct another table, which compares the
predicted masses with the experimental masses and also tabulates the predicted binding energy per nucleon (i.e., BE/A). The RMS error in the
predicted masses in Table III was 2.9 MeV. A comparison between the
M/A and BE/A will readily display the predicted error in the binding
energy per nucleon, because M/A is the error in mass per nucleon;
therefore, the sum of M/A and the predicted BE/A is the experimental
binding energy per nucleon.
In the development of the energy, Eqn. (3.101), we assumed that an
extra proton could be added in an orbit and the interaction between that
new proton and the other protons could be ignored.
This assumption was made even after we saw that any odd proton
sets up asymmetrical forces. Thus, errors could have been expected. Still,
the RMS errors from this crude averaging procedure do not appear too
inaccurate when even the best of the semiempirical nuclear mass formulas
does not address nuclei below a Z of 16 because of the large errors that
arise.15
To avoid the errors resulting from ignoring proton-proton interaction,
we must reconsider the simplest case, a single electron and three protons
in orbit around this electron. Proton-proton interaction suggests that the
protons will arrange themselves in a plane spaced on the points of an
equilateral triangle, as in Fig. 10. For this case, the force on a single
proton would be
F3=

where R =

3r

- | k |Y
r

1-

2|k |
R

cos

1-

397. Then the force becomes


F3=

|k |
r

1
1 - e e3
3r

3r

-Y 1-

(3.103)

This force differs from the two-proton force of Eqn. (3.99) by the coefficient
of the separation, r, in the proton-proton portion of the force. The
relativistic circular orbit equations of motion would be
2
2
-|k |
n (n )
=
3
2
mp r
r

1
p
1e
3
3r

-Y 1-

which differs from Eqn. (3.100) only be replacing R = 2r by R =


Then the energy is

153

(3.10

3r

400.

E 3 (n, Y) =

3|k |
R(n)

1 e
3

- Ye-

3r

3 m p c2
[R(n)]

(3.105)

+ Ec 4 .

Table III. Mass predictions for selected nuclei.

Y
1
1
1
1
2
2
2

Z
1
1
2
3
1
2
3

2
3
3
3
4
4
4
4
5
5
5
6
6
6

4
2
3
4
1
2
3
4
2
3
4
2
3
4

Expermental
Mass
2E1 + Ec
A (amu)
(amu)
2 2.012836 2.012836(2)
2 2.016000 Excited 2H
3 3.014848
(2)
4 4.025231
(2)
3 3.015484
(2)
4 4.001422
(2)
5 5.010676 1.992064(2)
(2)
6 6.016880 1.006204(4)
5 5.011046 2.004404(2)
6 6.013260
(2)
7 7.012129
(2)
5 5.035709
(2)
6 6.017709
(2)
7 7.013707 2.069947(2)
8 8.002459 2.056639(2)
7 7.026850 2.057980(2)
8 8.020624
(2)
9 9.009337
(2)
8 8.032752 2.079702(2)
9 9.024927 0.992175(7)
10 10.010689
(2)

Predicted
E2
Mass
(amu)
(amu)
2.012836
1.00773(2) 2.016000
1.001484(1) 3.020562
1.010179(2) 4.028288
(1) 2.998268
(2) 4.004472
(3) 5.010676

M
EP - EE
(MeV)/A
0
0
1.8

6.016880
(3) 5.011046
1.002214(4) 6.013260
(5) 7.015474
(3) 5.036203
(4) 6.024955
(5) 7.013707
0.988752(6) 8.002459
(5) 7.026850
0.993774(6) 8.020624
(7) 9.014398
(6) 8.032752
9.024245
(8) 10.017102

0
0
0
0.4
0.1
1.1
0
0
0
0
0.5
0
-0.1
0.6

-5.3
0.7
0

If we consider a possible symmetric orbit of four protons, as pictured in


Fig. 11, then the force on each proton would be

F4=

-Y | k |

1-

where R =

2
p

2 | k | cos
1 -

r
+

2r 402.

Figure 10. Four protons in symmetric orbit.


154

(3.106)

|k |
p
1e
2r
(2r )2

2r

Thus Eqn. (3.106) becomes

4
F =

-|k |
r

2
p
1e
2
2r

2r

2 |k|
1=
2 r2

2r

1
p
1e
4
2r
e

2r

2r

-Y 1-

(3.107)

+ F 2 (Y) ,

where F2 is the two-proton force of Eqn. (3.99).


circular motion equation of motion as
2
2
- 2 |k|
p
n (h )
=
1e
3
2
2r
mp r
3r

We can also write the

(3.108)

- F 2 (Y) ,

3r

and the energy equation as


E 4 (n, Y) =

4|k |
R(n)

2 e
2

2r

2 2 |k| =
e
R(n)

+
p

2r

1 e
4

2r

+ Ye-

e-

4 m p c2
[R(n)]

+ E c (Y)

(3.109)

+ 2 E 2 (n, y) - E c (Y) .

The foregoing discussion on a possible method of accounting for


proton-proton orbit interaction seems to imply that we must do every
possible nuclear configuration differently.
The final answer to this
question must await further research. The discussion was presented to
point out that ignoring the proton-proton interactions may be an error
source in the predicted masses. For instance, consider the force for the
symmetric four-proton orbit, F4, given by Eqn. (3.107). This force may be a
stronger attractive force for a given Y and r than F2, so long as r < p / 2 406,
but it is weaker for r > p / 2 407. Thus, for orbits where r < p / 2 408, four protons with n
= 1 can produce less ground-state energy from E4 than the energy expression in
Eqn. (3.99) does with two protons in the n = 1 orbit and two in the n = 2
orbit. The large error would thus diminish in the prediction of the 4He
nuclei mass.
The appearance of the exponential term requires the use of
numerical solutions, whether one tries the circular orbit approximation or
uses the quantum mechanical approach. This will be the subject of future
research. However, the masses predicted by the very simple assumptions
and the comparison of experimental and predicted binding energy per
mass number, plotted in Fig. 7, seem to imply that a more detailed
solution may prove very useful.
155

Figure 11. Binding energy per mass number versus mass number.

156

CHAPTER 5 - GRAVITATION
5.1 Charge-to-Mass Ratio and Magnetic Moments
In Chapter 1 a brief overview of the Dynamic Theory was presented.
The fundamental principles of the Dynamic Theory were presented in
Chapter 2. From these fundamental laws the constancy of the speed of
light was derived, the required geometry was obtained, classical and
special relativistic equations of motion were derived, and the conditions
requiring quantum mechanics were displayed. The requirements of the
fundamental laws were carried further in Chapter 3 by looking at the
gauge fields of the resulting five-dimensional geometry when mass is
considered as an independent variable. When quantization conditions are
considered in five-dimensions we found experimental features of particle
physics required by these new laws. For example, we saw that octets are
required fundamental states reminiscent of Gell-Man's eight-fold way; the
allowed fields for fundamental particles were shown to be quantitized in
electric charge; and the radial field dependence to display a short-range
non-singular behavior which allowed it to predict nuclear masses from it's
deviation from the Coulombic radial dependence and nuclear decay (beta
decay) from the asymmetry of while particle forces.
Thus, in chapters 2, 3, and 4 we have shown how the Dynamic
Theory reproduces, by using the appropriate restrictive assumptions, the
fundamentals of all the current branches of physics except gravitation. In
Chapter 4 the radial field dependence was derived and the long-range
dependence required that the new field components be interpreted as the
gravitational field and the gravitational potential. In this chapter we will
explore a few aspects of this interpretation. In particular, we will look at
some of the predictions of the Dynamic Theory in comparison with
Einstein's General Theory of Relativity.
Before we plunge into the derivation of a prediction to compare with
the General Theory of Relativity let us first consider a question which
arises from the necessity of keeping the units straight among the field
quantities E, B, V, and V4 when they are all to be considered as
components of the five-dimensional gauge field. By considering the units
of these field components it is soon found that a charge-to-mass ratio is
needed in order that the units of the gravitational field components may be
compared, or put in the same equation, with the electric and magnetic
components. Let us first see if we can determine this ratio.
In the Chapter 4 the derivation of the fields allowed for fundamental
particles was presented.
These field expressions give rise to the
specification of a charge-to-mass ratio which allows conversion of classical
gravitation field units to electromagnetic field units.
The gravitational field component in the system of field equations
with the electric and magnetic components brings up the requirement for a
157

gravitation-to-electromagnetic unit conversion. This need may be seen by


looking at the different field quantities. First consider the electric case.
The field units are given by [E] = volt/meter, while the expression for the
electric force density is Fe = E with units of newton/meter3.
For the gravitational field, in the Dynamic Theory, the units are [V] =
webers/meter squared, while the gravitational "current" density has units
given by [J4] = ampere/meter2. Thus, the gravitational force density is
given by Fa = (J4/c)V where again the units are newton/meter3.
In order to compare this system to the classical gravitational system
we need to be able to go from a gravitational field with units of
newton/kilogram to units of volt/meter. Now
nt volt - coul
volt
=
=
kg
m - kg
m

coul
kg

Thus, if is a quantity with units of coul/kg, then (1/) is the


conversion factor we seek. Similarly, we need to convert the gravitational
mass density, (J4/c), with units of coul/m3 to units of kg/m3. Obviously
will also be the conversion factor for this also. The question is; How do we
determine this charge-to-mass ratio and is it unique?
If we consider the fields the Dynamic Theory gives for fundamental
particles we may determine . Thus, let us look at the solution of the
gauge function for fundamental particles given previously, that is
1

ln F 2 = f r f f f f t .

We showed that this became


1

ln f 2 = f f t

(5.1)

for fundamental particles.


The functional dependence of the gauge
function upon time or mass density was not determined then.
Recalling from past reading that measurements of a time
dependence of the earth's gravitational field have been reported(). We may
proceed to make the simplest possible assumption about the functional
form for f and ft, namely linear dependence. If the functions have only
weak dependence upon time or mass density then this dependence could
easily be masked in experimentation not specifically designed to look for it.
Thus, lets' consider the form
1

ln f 2 = (a + bt) (s + w )

158

(5.2)

where a, b, s, and w are constants to be evaluated using known


information about the proton and time dependence measurements of the
gravitational field.
The gauge potentials are then
1

(ln f 2 )
N0
er
=
b (s + w )
0=
(ict)
ic
r
1

1=

(ln f 2 )
= - N 1 (a + bt) (s + w ) 1 dr
r
2

er
2
r

(5.3)

=0

and
1

4 = a0

(ln f 2 )

= a0 w(a + bt)

er
r

Using these potentials the field quantities become


Er =

( N1 - N0 )
b(s + w ) 1 c
r

er
2
r

E = E =0 ,
Br = B = B = 0 ,

(4.4)
-

V r = ( N 1 - N 4 ) a0 (a + bt) w 1 -

er
2
r

V =V = 0 ,

and
-

V4=

a0 ( N 4 - N 0 )
er
(bw)
c
r

In Eqns. (5.4) we may see the effect of the time and mass density
dependence of the fields. First, notice that the electric field, Er, vanishes as
the quantity b vanishes. From the expression for the radial component of
the gravitational field, Vr. we see that b is the time dependence of the
gravitational field. Thus, in order for an electric field to exist there must be
a time dependence of the gravitational field. Similarly, one may see that
the electric field must depend upon the mass density in order for there to
be a gravitational field. For a gravitational potential, V4, to exist not only
must the gravitational field depend upon time but also the electric field
must depend upon the mass density. This is a rather extraordinary
revelation!
159

Now we must check these field quantities in the eight field


equations. Because B = 0, then
B=0

is satisfied. Next
1 B
+
c t

x E=0

implies
x E=0

which is satisfied by the spherical symmetry of the E field. Then


xB-

E
=
t

4
c

J - a0

is satisfied if the current density vector, J, vanishes as it should for


particles.
Looking at
E=

V4

- a0

we find that this is satisfied by a spatial charge distribution of


( N 1 - N 0 )b(s + w )
=

cr

2-

(5.5)

The continuity equation


0=

J + a0

requires
J4

(5.6)

=0 .

Then
x V + a0

implies
160

=0

xV =0

which is also satisfied by the spherical symmetry of V.


When we consider the radial component of
1 V
E
+ V 4 = a0
c t

we find that it is satisfied if


(5.7)

N4= N0 ,

while the components in the and directions are satisfied identically.


The last equation is
V+

V4
4
=t
c

J4 .

This equation requires that


J4=-

a0 cw
( N 1 - N 0 )(a + bt)
4 r4

2-

(5.8)

Note that the expression for J4 satisfies Equation (5.6). We also find that
the radial dependence of J4 is identical to the radial dependence of .
Thus, we may rewrite Eqn (5.8) as
J4=

- a0 c 2 (a + bt)w
b(s + w )

T.C. Van Flandern, of the Naval Observatory, has reported a


measured very small time rate of decrease in the gravitational field given by
him to be approximately 6-parts in 1011 per year. If we designate this rate
of decrease by dG/dt, then
b
=G .
a

From the experimental measurement


G .6 x 10 -11 /yr

- 1.9 x 10 -18 sec-1 ,

then
a + bt

Thus, the non-zero field quantities are

161

a .

(5.9)

1E r = aGz(s + w )

1V r = Z a0 w(a + bt)

1Z a0 wa

r
r

J4=

r
r

(5.10)

- Z Ga(s + w )
=

2-

4 cr 4
- a0 c 2 aw
b(s + w )

where Z = N1 - N0.
We should note in Eqns. (5.10) that the gravitational field depends
upon the universal constant, ao, being non-zero. This implies a linkage
between the maximum mass conversion rate and the gravitational field in
somewhat a similar way as the electric field depends upon the speed of
light. Further, it should be noted that if the electric field does not depend
upon the mass then there could be no gravitational field.
Now the total charge is given by
q=

ds .
vol

Therefore, using Eqn. (5.5) and the spherical element of volume dv = r2sin
drd d , we have

q=
vol

- Z aG
4 c
- Z aG
4 c

(s +

) 2-

er 2
( r sin drd d ) .
r r4

r= R

(s + w ) 2 r =0

er
sin drd d
2
r

where R is the radius of physical extent of the particle. If the mass density
in the particle is a constant, 0, then the charge is given by
162

q=

r= R

-Z a G
c

(s + w )

er
dr
r r2

2-

r =0

Z aG
=(s + w
c

)1

(5.11)

eR.

Similarly, we may denote the gravitational "charge" as


M
J4
dvol .
=
c vol c 2

Then, using Eqn. (5.8), we have


M
c

- Z a0 wa

2-

er
ds
r

vol

(5.12)

Z a0 wa
1e r.
c
r

Now the electrons' force is given by


Z aG
F e = q Er =
c

s+w

1-

1-

ZaG
c

s+w

r
r

or

Fe

1-

2
2
2
Z Ga G
2
s
2
c

r
r

1-

since s >> 0. If we compare this with the classical expression for r >> ,
then we must have
ZaGs
c

1-

Z e
4

or
2 2
a s 1-

Thus,
163

e c
4

(5.13)

ec

as
4

1
2

G 1-

,
e

- r
2

which gives the product of a and s in terms of the experimental quantities


, c, e, and dG/dt.
Let us now turn to the gravitational force, which is given by
Fg= MVr
=

Z a0 wa

1-

e r (Z a0 wa) 1 -

(Z a0 wa )2

1-

er
r r2

-2

e r
.
2
r

By comparing this with the classical expression for the gravitational force
we find we must have
(Z a0 wa )2 1 -

e r=

GM2

or
aw =

m
Z a0

1-

(4.14)
e

If we consider the particle to be a proton then Z = 1, and the ratio of


the magnitude of the electric force to the gravitational force is given by
2

FR=

e
,
4 G m2

thus
G m2 =

e
4

FR

and
( a0 wa )2 =

from Eqn. (5.14). Then, we have

164

e
4

FR

aw =

e
a0

Fr

e
a0 c 4 F R

(5.15)

Now choose
a=

(5.16)

G ,

(5.17)

so that, by Eqn. (5.11),


e

b=

and, from Eqn. (5.15),


m

w=

Za o ce 1 -

1
2

(5.18)

.
e

2r

From Eqn. (5.13) we find


c

s=

4 G 1-

1
2

(5.19)
e

2r

These values point out the extremely weak time dependence of the
gravitational field and the very weak mass dependence of the electric field.
This mass dependence may be better seen if we write
aZG(s + w )
Er =
c

1-

r
r

then substitute for the field parameters to arrive at

165

ZeG
Er =
c

c
c
+
4 G a0

Ze
=
4

1+

G4

Ze
=
4

4 Gm
1+
a0 e

1-

1
4 Fr

1-

G m2

a0

1G

r
r

Now let us return to the search for the charge-to-mass ratio since we
have all the necessary information. The quantity we defined as the
gravitational "charge" is given by Eqn. (5.12). If we divide this by c we have
M Z a0 wa
=
1er ,
c
c
r

but, by Eqn. (5.14), this becomes


M
Z a0
=
c
c
=

m
Z a0

m
c

=m

G
1-

G 1-

1e

r
e

or, rewriting
M
=m
c

G 1-

e r.

(4.20)

Thus, the charge-to-mass ratio we seek is given by


=

M/c
=
m

G ,

or
=

G = 2.4296 x 10 -11 coul/kg .

166

(5.21)

The surprising, and pleasing, thing about this result is that it is


formed as the product of two known physical quantities rather than
depending upon new quantities, such as a0 and , whose values are not well
known. Further, in retrospect, it appears to be the simpliest, if not the
only, combination of an electromagnetic parameter and a gravitational
parameter whose units are coul/kg.
It is worthwhile to point out that the dependence of the fields upon
time and mass density (ie, b and w) is extremely small but is essential in
establishing , and the inductive coupling between the electromagnetic
and gravitational fields.
The charge-to-mass ratio brings up the notion that a rotating
gravitational, electrically neutral body should have a magnetic moment
stemming from the effective electric charge associated with the
gravitational mass. Given the charge-to-mass ratio we may quickly look at
its prediction for the earth's magnetic moment.
Using , the "effective" charge associated with a gravitational mass is
given by
q eff = M .

For the earth this effective charge would be


q eff = 1.454 x 1014 coul .

Thus, if the magnetic moment of the earth is given by


= q eff /2M A ,

where A is the earth's angular momentum then we have


=
=

q eff
2M

(1.454 x 1014 )(9.71 x 10 37 )(7.29 x 10 - 5 )


2(5.983 x 10 24 )

or
= 8.6 x 10 22 amp - m 2

This predicted value of the earth's magnetic moment compares very


well with the experimental value of 8.1 x 1022 amp- m2.
5.2 Perihelion Advance
No serious suggestion that the additional vector field in the fivedimensional gauge equations of the Dynamic Theory be the gravitational
field can be made without giving due consideration to the explanation of
the planetary perihelion advance provided by Einstein's General Theory of
Relativity. Though several attempts have been made to explain the
perihelion advance by other means none has succeeded in casting much
doubt on Einstein's explanation.
Let us recall some of the main features of the classical problem of
planetary orbits. Kepler's first law states that a planet describes a closed
elliptical orbit with the sun at a focal point. However, the presence of such
167

small influences as other planets moving in the suns' field causes a


perturbation in the motion of a given planet, and the resulting orbit is not
precisely elliptic. Indeed, one may think of the actual orbit as a slightly
bumpy ellipse which may precess in the plane of motion; that is, the
perihelion shifts about and does not always occur at the same angular
position.
The fact that the idealized classical orbit is a closed ellipse is a result
peculiar to the Newtonian inverse-square law; in fact, Newton himself
found that, if the force of gravity were proportional to 1/r(2+ ) instead of
1/r2, then a planetary orbit would not be closed and a perihelic shift of
order would occur. Indeed, this result was taken to indicate that, since
planetary orbits are very nearly closed, the Newtonian inverse-square law
must be very accurate, as in fact it is.
Let us now ask were may there be room for differences between the
predictions of classical celestal mechanics and the celestial mechanics of
the General Theory of Relativity or the Dynamic Theory presented here.
Since Kepler's first law is experimentally verified to be correct to a high
accuracy, we might expect that non-Newtonian Theories may merely add a
few bumps to the nearly elliptic orbits and contribute somewhat to
perihelic motion. Since angles are much more conveniently measured in
astronomy than are distances, it is natural to concentrate on perihelic
motion. Conveniently enough, there is, in fact, a well-known discrepancy
in classical mechanics concerning the perihelic motion of the planet
Mercury. Because of Mercury's high velocity and eccentric orbit, the
perihelion position can be accurately determined by observation. The
difference between the classically predicted perihelic shift (due to
perturbations by other planets) and the observed perihelic shift is 43
seconds of arc per century. Even though this is a very small difference, it
is about a hundred times the probable observational error and represents
a true discrepancy from the very precise predictions of celestial mechanics
which has bothered astronomers since the middle of the last century.
The first attempt to explain this discrepancy consisted in
hypothesizing the existence of a new planet, Vulcan, inside the orbit of
Mercury, and much theoretical work was done to predict the position of
Vulcan, using the known perturbation on Mercury's orbit. However,
careful observation failed to discover the hypothetical planet, and the
hypothesis was finally abandoned in 1915 when Einstein used general
relativity theory to explain the observed effect.
Now let us look at what the Dynamic Theory offers as an explanation
for the perihelic advance and then compare it to the predictions of the
general relativity theory.
The classical equations of motion are
mr - mr 2 = F(r)
(5.22)
mr + 2mR = 0 .
The second of Eqns. (5.22) has the solution
168

L
mr

(5.23)

where L is the angular momentum.


Using Eqn. (5.23) the first of Eqn.s (5.22) may be written
Mr = F(r) +

L
M r3

or
Mr

(5.24)

v (r) ,

where
v (r) = v(r) +

L
2 Mr 2

(5.25)

We are seeking the prediction of the Dynamic Theory with respect to


the perihelion advance. This may be found by comparing the frequency of
small radial oscillations about steady circular motion for the effective
potential given by Eqn. (5.25) for the non-singular potential of the Dynamic
Theory with the frequency of revolution.
By considering the non-singular potential of the Dynamic Theory,
Eqn. (5.25) becomes
2
K L
e r+
2 Mr 2
r

v (r) =

(5.26)

with K = -GMm, where G is the gravitation constant, M is the mass of the


sun, and m is the mass of the planet of interest.
Equation (5.23) gives the frequency of revolution. To determine the
frequency of small radial oscillations about steady circular motion we need
to evaluate the second derivative of the effective potential, v, the radius for
which the first derivative is zero.
The first derivative of the effective potential is obtained by
differentiating Eqn. (5.26) with respect to r. This may be found to be
r

v (r) =

-K
r

1-

er-

L
3
mr

(5.27)

The second order derivative of the effective potential may be found to


be approximately
2

v (r)

2K

1-

169

2
r

e r+

3 L2
mr

(5.28)

when terms involving 2/r2 are considered negligible with respect to terms
involving /r.
We may determine r0 from the condition
2
[v (r)]
GMm
L
=0=
1e r0 3
2
r
mr 0
r0
r0

(5.29)

The radius, r0 is the radius of near circular orbit and the effect of the
exponential factor and (1- /r) factor will be negligible for <<r. Thus, we
may approximate Eqn. (5.29) by
GMm

0=

L
3
mr 0

2
0

so that
2

L
2
GMm

r0

(5.30)

If we approximate the exponential factor in Eqn. (5.28) by its power


series expansion and retaining only those terms whose dependence upon
/r are linear or less, then Eqn. (5.28) is approximated by
2

v (r)

2K

1-

2K

2
r

1-

3
1r

3 L2
mr

3 L2

mr

(5.31)

Now the frequency of small radial oscillations about steady circular


motion may be found from
2

1
m

v (r)
r

r = r0 ,

Thus, we have
2

1
m

GMm
2
L

- 2GMm
=

- 2 G 4 M 4 m6
L

11-

6
4

G M m
6
L

3 GM m2
L

3 GM m 2
L

1+

6 GM m2
L

3 L2
m

3 G 4 M 4 m6
L

GMm
2
L

(5.32)

An approximation for the frequency of small radial oscillations about


steady circular motion may now be made by taking the square root of Eqn.
170

(5.32) and considering the second terms of the second factor as small
compared to one. Thus, we have
2

G M m
3
L

3 GM m 2

1+

(5.33)

The perihelion advance per revolution may now be found as the


difference between Eqns. (5.33) and Eqn. (5.23) evaluated at r0, divided by
the orbital frequency, or
-

=2
2

=2

G M m
3
L

1+

3 GM m2
L

G M m
3
L

L
2
2 3
G M m

so that
3 GMm2
2
L

The perihelion advance predicted by Einstein's General Theory of


Relativity is given by
GiR

3 G 2 M 2 m2

=2

c L

If
were to be such as to provide an identical prediction as the
General Theory then would have to satisfy
=

GM
c

For G = 6.7 x 10-8 gr-1cm3/sec2, M = 1.98 x 1033gr, and c = 3 x 1010


cm/sec,
=

(6.7 x 10 -8 )(1.98 x 10 33 )
cm = 1.47 x 1015 cm
(3 x 1010 )2

or
= 1.47 x 10 3 m .

This is an extremely small value compared to the radius of the sun


but is sufficient within the Dynamic Theory to provide the same prediction
of perihelion advance as the General Theory of Relativity.
171

5.3 Redshifts
Einstein's General Theory of Relativity predicted the advance of the
perihelion of planetary orbits by using the full effect of the geometrical
equations. We saw in the previous section that the Dynamic Theory
predicts a similar planet orbit perihelion advance.
Another of the
predictions of Einstein's Theory concerns the redshifts associated with
light received from distant light emitting objects or when light travels
through a changing gravitational field.
The Dynamic Theory should also predict frequency shifts that are
experimentally measurable. If it does not then it doesn't have the same
strength of predictability as Einstein's General Theory.
There are two types of redshifts resulting from the theoretical
approach of the Dynamic Theory. First, there is an expansion red shift due
to the increasing "entropy" of the universe. Secondly, there is a frequency
shift caused by a difference in the gravitational strength between the point
of emission of the light and the point of its reception. Both of these types
of frequency shifts are the result of a difference in the effective unit of
action at the emission point and the reception point.
Both of the above types of frequency shifts may be referred to as of
geometrical in origin in that they both come from the gauge function.
However, each originates from a different variable change in the gauge
function. For instance the expansion shift involves considering the
universe as an isolated system resulting in the entropy principle requiring
small frequency shifts toward the red. This comes from the gauge function
being dependant upon time. The second type of frequency shift comes
from the gauge functions dependence upon space and mass.
We may first consider these types of frequency shifts to be
independent and look at each in turn. Then we shall consider them
together. First, we will need to consider the local systems where a photon
is first emitted and then where it is received. In both systems the energy of
a photon is given by hv, where v is the frequency and h is the effective unit
of action. The effective unit of action is the product of the gauge function
and Planck's constant if a locally flat metric is considered.
At the heart of both types of frequency shifts is the gauge function
which has previously been given as
1

(5.34)

ln f 2 = f r f t f f f

and found to be

ln f 2 = f t f

172

er
r

where ft and f indicates functions of time and mass density.


We need to determine more about the gauge function than we
previously have. The square of the arc lengths differ by the multiplicative
gauge function as
( dq0 )2 = f(d

or
( dq0 )2 = exp 2 f f

er
r

2
) .

(d

From this we see that the differential change in entropy is given by


0
dq = exp f t f

er
r

(5.35)

Recalling that there can be no decrease in the entropy for an isolated


system we must then consider the possible effect of the entropy principle
upon the universe as an isolated system. We can see from this line of
thinking that ft in Eqn. (5.35) is the one to focus on for the moment. The
simplest function is of course the linear function and this linear
dependence appeared previously in section 5.1.
Suppose then, we consider the effective unit of action for an isolated
universe at some time which we will set at t = 0. We find
1

h0 = h f 0 ,

which can be written


(5.36)

h0 = h exp[0]

at t = 0. Here the value of the effective unit of action corresponds


identically with Planck's constant, h.
At some later time t = T the effective unit of action would be given by
1

hT = H exp[AT] ,

where A is a constant. Let us further consider a change of variables using


the distance light will travel in free space instead of the time, T. Since T =
L/C, L is the distance variable we seek. We now have
1

hT = H exp

AL
c

(5.37)

If a photon were emitted at t = 0 it would have been emitted with an


effective unit of action given by Eqn. (5.36). If that photon is received at
173

the later time, T, at a distance of L, then the universe's effective unit of


action would be given by Eqn. (5.37) at reception. If the energy of the
photon when emitted is given by h10 e 85, then no loss of energy by the photon until
reception would require that
1

(5.38)

h0

= hT

Substituting from Eqn. (5.36) and Eqn. (5.37) into Eqn. (5.38) we find
e

exp

AL
c

(5.39)

The frequency shift would be given by


r

1- e

AL
c

AL
c

or
= e-

AL
c

(5.40)

-1 .

The question arises whether or not the frequency shift given by Eqn.
(5.40) is red or blue? From Eqn. (5.40) it may be seen that the frequency
shift is negative if A > 0, thus the shift is red or blue as A is positive or
negative. Going back to Eqn. (5.35) and using the gauge function of Eqn.
(5.37) we see that
0
dq = e

AL
c

This indicates that a given element of arc length, d , yields a larger change
in entropy, dqo, at the time T = L/C than before at time t = 0. Thus, the
entropy change is increasing and our universe is expanding.
The expansion red shift given by Eqn. (5.40) may also be expressed
in terms of wavelength as
=
e

- 1= e

AL
c

-1 .

1 AL
3! c

(5.41)

Equation (5.41) may be expanded as


=
e

AL
1 AL
+
c
2! c

which may be approximated by


174

+ ...

(5.42)

AL
c

(5.43)

when AL << c.
Experimentally it has been determined that

HL
,
c

where H is the Hubble constant which is given by


17
-1
H = (5.6 + _ 0.6) x 10 sec .

Then we find the predicted frequency shift given by Eqn. (5.42) is seen in
nature when the constant, A, is taken as the Hubble constant.
From Eqn. (5.41) we see that experimentally found red shifts can be
used as astronomical markers from the expression
c
ln 1 +
H

L=

exp

(5.44)

For small experimental red shifts, compared to one, the astronomical


distances L, given by Eqn. (5.44), are not significantly different from the
linear markers given by the approximation of Eqn. (5.43). However, as the
red shift begins approaching unity the difference becomes significant.
Turning to the second type of frequency shift, and returning to Eqn.
(5.34) to again write
-

er
r

ln f 2 = f f

(5.45)

The effective unit of action is expressed by


h

= h f g kl
jk

+
ie

j s
x
sl

(5.46)

But when in a locally Euclidean geometry the gkl =


becomes
-

h = h f = h exp 2 f t f

From the expressions for the gauge potentials

175

er
r

kl

then Eqn (5.46)

(5.47)

j=

(ln f 2 )
1 (lnf)
=
3
2 xj
x

(5.48)

and
F ij =

i, j

(5.49)

j,i

Using Eqn. (5.49), F14 gives the radial component of the gravitational field.
From Eqn. (5.48) we find
1

(ln f 2 )
=- f t f
1=
r

er
2
r

1-

and
F 14 = - a o f t

df

1-

er
r

(5.50)

The expression for the gravitational field given in Eqn. (5.50) is in


terms of a field density. By integration over the volume occupied by the
gravitational mass density we have the gravitational field
V r = - ao f t

df M
dM

1-

er
r

(5.51)

where M is the total gravitating mass.


For any weak time variation in the field we can ignore the time
dependence. Thus, we have
V r = - ao a

df M
dM

1-

er
r

(5.52)

to be compared with the experimental field


V r exp =

GM
r

Certainly for r>>


Vr in Eqn. (5.52) is approximated by the
experimental expression of Eqn. (5.53) if A = -G and
df M
=M .
dM

From Eqn. (5.54) we find that


176

(5.53)

1 2
M + constant.
2

fM=

The conversion to a field density may be done by dividing by the mass, M,


so that the gauge function in Eqn (5.47) becomes
1 2
M + constant
2
M c2

- 2G
1

h = H exp

er
r

or
-

- GM

h = h exp

er
r

(5.55)

where c2 has been used to obtain a unitless quantity, which must be the
case for f.
Now, using the unit of action given by Eqn. (5.55), suppose a photon
is emitted from one body with a gravitational field
GM 1 - 1
e R1 ,
2
c R1

and is received in another gravitational field


GM2

c R2

R2

the conservation of photon energy would then require


he

= hr

or
e

- GM 1

exp

c R1

R1

exp

- GM 2
2

c R2

R2

so that
r

exp

- GM 1
2

c R1

The frequency shift would then be


177

R1

GM 2 - 2
e R2
2
c R2

(5.56)

=
e

For R2 >>

GM 2 - 2 GM 1 - 1
e R2 - 2
e R1
2
c R2
c R1

= exp

(5.57)

and R1 >> 1, then Eqn. (5.57) may be approximated by


e

G
c

-1 .

(5.58)

The approximation in Eqn. (5.58) shows that if M1/R1 > M2/R2 then this
frequency shift given by Eqn. (5.57) is negative, or towards the red end of
the spectrum.
We can make the further simplification of assuming that both
GM 1
GM
< < 1 and 2 2 < < 1 ,
2
c R1
c R2

then we would have the approximation that


_
e

GM 2 GM 1
- 2
.
2
c R 2 c R1

(5.59)

In terms of wavelengths we have


=
e

= exp

-1

GM 1 - 1 GM 2 - 2
e R1 - 2
e R2
2
c R1
c R2

(5.60)

with the above assumptions Eqn. (5.60) is approximated by


_
e

GM 1 GM 2
- 2
.
2
c R1 c R 2

Suppose we look at this red shift for a photon emitted from the
surface of the sun and received at the earth's surface. The needed
numbers are:
G = 6.67x10-11nt-m2/kg2 Msun = 329,390(5.983x1024kg)
Mearth = 5.983x1024kg
Rsun = 6.953x108m
Rearth = 6.371x106
c = 3x108 m/sec
Thus, from Eqn. (5.61),
178

(5.61)

2.1 x 10 -6 ,
e

or 0.63 in km/sec. This is the same as predicted by Einstein's General


Theory of Relativity.
For a terrestrial test of the red shift the prediction would be
_
e

G
c

M
M
R
R+ R

GM
2

c R

R .

If R = 72 ft = 21.95m, then / e 2.4x10-15.


Since the approximation given by Eqn. (5.59) may be expressed as
d

where
= G[(Me/Re)-(Mr/Re)], then it may be seen that the red shift given
by Eqns. (5.57) and (5.60) produce the red shifts predicted by Einstein's
General Theory of Relativity if R1>> 1, R2>> 2, GM1<<c2R1, and GM2<<
c2R2. However, if these conditions of approximation are not met then one
must resort back to Eqn.'s (5.57) and (5.60) for the predicted red shifts.
Suppose one considered a photon which may have been emitted on a
dense star such that GM1/C2R1 is too large to allow a simplification of the
exponential expression. If this photon were received on earth then, by
Eqn. (5.60), we would have
G

exp
e

exp

M e M earth
R e R earth

-1 .

Because of the small value of (GMearth)/(C2Rearth) 7 x 10-10, then the


approximation becomes
_e

GM 2
c2 r 2

(5.62)

-1 ,

exp

where GM/c2R is the gravitational field at photon emission.


From Eqn. (5.62) we may learn something of a stars' density by the
red shift in the light received from it. For instance, Eqn. (5.62) has
GM
_ ln 1 +
2
c R

.
e

exp

Notice that even the large red shifts displayed by quasars are allowed
by Eqn. (5.63) without requiring them to be at the far reaches of our
universe.
179

(5.63)

We have looked at the predicted frequency shifts as if they were


independent phenomena. In the sense that they stem from the same
gauge function, then both types of redshifts may be present in any light
received. Thus, we should look at the expression for the entire red shift.
Suppose we let
1

ln f 2 = f r f t f = (s + w + z

k eR
r

) (a + bt)

where s, w, z, a, b, and k are to be evaluated. The effective unit of action


becomes
1

h0 = h exp

k e r (s + w + z
r

)(a + bt)

(5.64)

for t = 0. From the above, let us take s = w = 0, then Eqn. (5.65) becomes
1

he = h exp

kz(a + bt e ) M e
Re

(5.65)

Re

when integrated over the entire gravitating mass as before and the
subscript, e, refers to the unit of action at the place and time of emission of
the photon. A similar expression is found at the place and time of
reception, or
kz(a + bt r ) M r

h r = h exp

Rr

Rr

(5.66)

Equations (5.65) and (5.66) may be used to express the frequency shift
required to conserve photon energy, since
1

he

= hr

Thus, we have
r

exp

zkM e (a + bt e ) - e zkM r (a + bt r ) - r
e Re e Rr
Re
Rr

then

180

= exp zk
e

M e (a + bt e ) - e M r (a + bt r - r
e Re e Rr
Re
Rr

Similarly, the expression for the wavelength shift becomes


= exp zk
e

M r (a + bt r ) - r M e (a + bt e )
e Rr Rr
Re

-1 .

(5.67)

If we write Eqn. (5.67) as a power series and make the approximation of


keeping only the first term, we find
zk
e

M r (a + bt r ) e
Rr

Rr

M e (a) e
Re

Re

(5.68)

where we've also let te = 0. By letting tr = L/c Eqn. (5.68) becomes


bL - r
e Rr
c

M r a+
zk

Rr

Meae
Re

Re

(5.69)

We want to evaluate the unknowns a, k, z, and b in terms of the


previously determined quantities such as G, c, and H. Therefore, suppose
that the gravitational field at the time of emission of the photon is the same
as here on earth at its reception, then we find Eqn. (5.69) becomes
zkbLM earth
.
CR earth

(5.70)

Experimentally, we have found the expansion red shift is given by


=
e

exp

HL
,
c

thus, we should have


H=

zkbM earth
.
R earth

181

(5.71)

On the other hand if the time between photon emission and


reception is sufficiently close then our approximation to Eqn. (5.69) can be
written as
Mr Me
R r Re

_ zka
e

where Rr>>
set

(5.72)

and Re>> e. Thus, in order to compare with experiment, we


G

zka = -

(5.73)

If we now substitute Eqns. (5.71) and (5.73) into Eqn. (5.68), we have
e

G
c

Mre
Rr

Rr

Mee
Re

Re

HL
c

Mr
Rr

R - r
e Rr ,
M

(5.74)

where R/M is the radius and mass of the earth as in Eqn. (5.71). While
the full expression, Eqn. (5.67) becomes
= exp
e

-G
c

Mre
Rr

Rr

Mee
Re

Re

HL
c

Mr
Rr

R - r
e R R - 1.
M

From Eqn. (5.75) it may be seen that, if the receiving location is the
earth, then the time dependence in Eqn. (5.75) is given by
HL
= Ht
c

so that b, of Eqn. (5.67) is given by -H. Thus, the gauge function


dependence upon time is also given by b= -H. Recall that a time
dependence of the gravitational field has been reported by T. Van Flandern
with a reported value of b = -1.9 x 10-18sec-1. The measured Hubble
constant is H-1 = (5.6 0.6) x 1017 sec so that 2.00 x 10-18 sec H 1.61 x
10-18 sec.
Thus, we see that the Dynamic Theory predicts that the expansion
redshift and a time decrease in the gravitational field strength are both the
result of the time variation of the gauge function. Further, it seems
amazing that two such different and difficult type measurements have
such a good agreement!
Returning to the wavelength shift given by Eqn. (5.75), it may be
seen that the contribution of expansion, or gravitational potential, to the
total red shift is contained within this equation. Equation (5.75) has three
unknowns: the astronomical distance L, the mass of the emitting star (or
object), Me and the size of the emitting star, Re. Given only two pieces of
182

(5.75)

experimental data such as the redshift and the apparent luminosity, we


can determine the astronomical distance, L, and the gravitational density,
Me/Re by assuming a mechanism for the light production (ie, sun-like).
Given another data point such as light fluctuation periodicity then a
limiting size might be obtained.
5.4

"Fifth" Force

Is a "fifth" force really necessary? Obviously, a new force is much


more exciting than finding an explanation for the measured effect within
another force description. A new force may be more tenable because it
may appear not to compete with existing forces. A correction to existing
forces may lack excitement and must certainly be shown to be compatible
with existing forces where they are measured to great accuracy. A
correction to an existing force is usually difficult to find and may go against
the preference of many. But to assume there is an additional force is to
assume its independence and would then necessitate yet another force to
be "unified".
Accurate measurements show that the gravitational force of the
Earth differs from Newton's Law at close range. More specifically, the
difference in the Earth's gravitational field over a difference of height in a
deep well is not the same as predicted by Newton's Law. This leads to a
simple choice; either Newton's Law of gravity needs to have a correction or
an independent fifth force is needed to explain the difference. In the past
when we found that the proton-proton scattering data differed from the
coulombic predictions we opted for an independent force and have had the
fun of searching for a method of unifying electromagnetism and the strong
force every since. We could make that same choice here, or we could
investigate the difference in the prediction of the non-singular potential
and Newton's law of gravity. To do this we need to look at the gravitational
attraction on a mass in a well deep down from the surface of the Earth.
Freshman texts on physics typically show how to calculate the
gravitational influence of a thin spherical shell on a mass both inside and
outside of the shell. This is the procedure we need here because a mass in
a deep well feel the influence of both the mass of the Earth interior to it
and in the shell of the Earth exterior to it where the shell thickness is the
depth of the well. If we recall the procedure for this using the Newtonian
potential, then we remember that, for the 1/r potential, all of the mass
interior to the test mass attracts it as if the mass were located at the center
of the Earth. On the other hand we recall that there is no gravitational
influence due to the mass in the outer shell which is exterior to the test
mass in the deep well.
For a potential which differs from the 1/r Newtonian potential these
conclusions may not be true. Indeed, ones first suspicion is that they are
183

not the correct conclusions. What we now need to do is to calculate the


influence on a test mass both outside and inside of gravitating mass.
First, suppose we calculate the gravitational influence on a mass
exterior to a thin spherical shell. Using the neo-Newtonian potential we
find the integral to be
R+ r

F=

2
2
R - r +1 1e x dx
2
x
x

- G rtm
R

x= R - r

2
R+ r

F=
x= R - r

d
dx

f(x) e x dx
x

Figure 12. Neo-Newtonian Potential.


Figure 13. Neo-Newtonian Force.
Figure 14.
Gravitational attraction of a section dS of a
spherical shell of matter on m.
where by making use of the integral tables and a lot of algebra we arrive at
the solution
F is an improper integral for R=r. This means that terms in the series have
denominators which tend to zero as R tends to r. We must show that the
series converges because this is the case when our test mass is at the
bottom of a deep well. It would then be at the outer surface of the inner
mass. It is easy to see that as tends to zero the solution tends to the
classical solution.
If we now consider our test mass to be inside a thin shell and look at
the force.
- G rtm

F=
=

- G rm
R

R+ r

d
dx
x= -(R - r)

f(x) e x dx
x

2r e- R+r - e- -(R - r) + 2 log

r-R
R+r

1
1
(- )N
N
N.N!
(R
+
r
(r
R )N
)
N =1

+2

as

0, F
0 which is the classical result.
Suppose now we look at a couple of approximations which may give
some insight into the influence of the exponential term in the potential.
First, let us consider a mass outside another mass for which R-r>> , then
we would have the approximation
F

- GMm
R

1-

184

2R - r
2
2
R -r

; if R - r > > .

This result shows that the force of attraction on a test mass outside
another mass is reduced by the second term in the square brackets. This
is, of course what we should have expected from a potential which deviates
from the Newtonian potential by turning around and going back to zero.
The first deviation from Newtonian-like character would be to become
weaker.
The other approximation to consider is that for the expression for the
force an the test mass inside the shell. For this we find that inside the
shell if r-R>> , then
F =

- GMm
R

(- )

r + rR - R
r( r 2 - R 2 )

; if r - R > > .

This is a force away from the center of the shell and toward the inside of
the shell.
The Big Question is: What is the force when the test mass is
on the immediate exterior, or interior, of a shell? That is, do we have
convergence of the infinite series in the solutions for both the inside and
outside forces on the test mass.
To address this consider the absolute value of the ratio of the n+1
and the nth terms in the force for a mass outside of a shell of finite
thickness t, then from our previous results we found that
F=

- G trm
R

+2

R-r
R+r

2r e- R+r + e- R - r + 2 log

1
1
(- )N
N
N.N!
(R
+
r
(R
+
r )N
)
N =1

The question arises whether or not F is finite for t > 0? If R=r+t/2


then R+r=2r+ t/2 and R-r = t/2, or r = R - t/2 R+r = 2R - t/2 ; R-r = t/2
- G tr 2 m
F=
2
R

2 e

2R -

t
2

(- )N
t
N.N!
R - N =1
2
2

if we have 2R>>t/2 and 2R>>


approximation,

+e

t
2

2
R-

1
2R -

t
2

log

t
2

t
2
2R -

t
2

1
t
2

then we may make the following

185

- GMm
R

t
1 2
log
e 2R + e t +
2
2R
2R
2

N
1
(- )N
2
2R N =1 N.N! (2R )N (t )N

Now look at the ratio of two successive terms in the series

a n+1
=
aN

N +1
(- )N +1
1
2
(N + 1)(N + 1)! (2R )N +1 t N +1
N
(- )N
1
2
N.N! (2R )N t N

(2R )
=

(N + 1 )2 tR

N +1

t
2

(2R )N -

t
2

N +1

But R>>t/2 so that


lim
N

a N +1
aN

lim
N

2
(N + 1 )2 t

Therefore, the series converges absolutely!


Now how about an expression for the force for a large number of
shells of thickness t sufficient to make up a sphere? If there are M shells
making up the sphere then the thickness t = Re/M. While if we want to
monitor the force on a test mass at the surface of the sphere then the
limits on x in the integration varies with each shell.
For instance for the outer shell the limits on x would be
R-r=

t
2

x R + r = 2R -

t
R
but R = R e and t = e 154 or the limits on x would be
2
M
Re
2M

x 2 Re -

3
2

1
Re
= Re 2 2M
2M

For the next shell R - r = t x R + r = 2R -

186

3t
156 or
2

3 Re
2M

3 Re
3
= Re 2 2M
2M

x 2 Re -

Thus, for the Nth shell we have limits given by


2N - 1
2

Re
N

x Re 2 -

(2N - 1
2N

Then we have
F (N) =
+

+
N =1

- GM (N)m
R

1 e
2 Re

2
e

log

(2N - 1)
Re 2 N
(- )N
N - N! R eN

2-

(2N -1)
N

+ e-

Re

2N -1
2N

2N - 1
4N - (2N - 1)

(2N )N
(2N )N
N
(2N - 1 )N
4N - 2N + 1

Now, looking at
exp1 =

(2N - 1)
Re 2 N

if N = 0 _ exp1 =

Re 2 -

if N = N _ exp1 =

1
N

Re

on the other hand for


exp2 =

R e 2N - 1
N
2

if N = 1 _ exp2 =
-

if N = N _ exp2 =

Re 1 -

Further, since r = R e 1 -

1
2N

(2N - 1)
162
2M
M (m) = 4

Re
M

1-

(2N - 1)
2M

If we choose M such that R e > > 164 then we may write


M

187

-2
MRe

1
N

F (m) =

- Gm
2

Re

Re
M

1-

(2m - 1)
2M

1
22

log

(2m - 1)
M

Re 2 +

(2m - 1)
Re 2 M

Re

2m - 1
M

2m - 1
4M - (2m - 1)

(2m - 1 )N - 1
(- )N (2M )N
N.N! R eN
(2m - 1 )N 4M - 2m - 1 N
N =1

Suppose we look at three shells such that M = 3 then we find

F1=

- GM
R

2
e

F1=

log

5
3

Re
=

Re
3

- GM 1 m
R

1-

Re

1
6

12 Re 2 -

1
3

2 Re

1
3

1
(- )N 6 N 10 - 1
+
12 - 1 N =1 N.N! R eN [11 ] N
1-

2
e

- GM 1 m

1-

3
3
2.398)3
2 5 Re 2 Re
5

4.198
Re

if M 1 = 4

Re
3

5
6

or
F out =

- GMm
2

Re

1 - 3.548

Re

Now we can do the same sort of thing for the "inside the shell" force.
This case has the lower limits on x changed so that
F in =

- G rtm
R

R+ r

2
x= r - R

d
dx

f(x) e- x
dx , where x = r - R
x

with
188

t
2

f(x) - ( R2 1 r 2
+ x e- x + 2
e x=
x
x

log

1
(- )N
+
N
x
N =1 N.N! x

Then for a shell with Re as the inner radius and a thickness of t then
r=Re=t/2 and we desire to know what the gravitational effort for R=Re.
Then we have R+r=2Re+ t/2 and r-R = t/2 then
- G

Re +

F in =

t
tm
2

- R e2 - R e +

t
2

t
2 Re +
2

2 Re +

t
2

t t e 2 Re + t - e t
2
2 2
2

e 2 + 2 Re +

t
2

+2

t
2

R - Re +

t
2 Re +
2

Re

2
e

t
2

(- )N
N.N!

1
t
2 Re +
2

t
2

if >>2Re, Re>>t/2, then we have the approximation

F in

2 Re

- GMm

t
4

Re

log

t
4 Re

1-

2 Re

2 Re

(- )N
2 R e N =1 N.N!

-2
t

2e
+
t

1
2 Re

2
N
t

Example: Compare the change in gravitational field as one goes from the
Earth's surface down a deep well to a depth of d. The Newtonian
gravitational strength at the earth's surface is
FN=

-G M e m
2

Re

For the neo-Newtonian force we must see the same force, thus, we must
set
F NN =

-G Mem
R

2
e

- GMm
2

Re

189

1 - 3.55

Re

or
Me

M=

1 - 3.55

Re

Now the Newtonian gravitation at a depth of d.


F N (d) =

- GM e m
Re - d

- GM e m
2
2
R e - 2d R e - d

or
- GM e m
2d
2
Re 1 Re

F N (-d)

The variation in force from the surface to the depth is then


- GM e m

FN

Re

1 - 1+

GMem
GM e m
+
=
2
2d
Re
2
Re 1 Re

GM e m
2
Re

1-

2d

Re
2d
Re

GM e m
2
Re

or
FN

2d
Re

GM e m
.
2
Re

On the other hand, the gravitation force in the well with the neo-Newtonian
potential would be given by
F NN (-d) =

- GM e m
R

2
e

1+

M
M e 4 Re

2
t
M
M
log
1 - e- t 4 Re
2 M e Re
2Me

(-2 )N
2 M e R e N =1 N.N!
M

190

1
4 Re

1
t

then FNN = FNN(Re) - FNN (Re-d)


F NN =

- GM e m
R

2
e

GM e m
2
Re - 4

but

M
Me

1
1 - 3.55

1+

M
M e 4 Re - d

(-2 )N
2 R e - d N =1 N.N!

2
1
d
log
1 - e- d 2
2 Re
4 Re

1
4 Re

1
d

181 so that for d = 5,500 ft = 1,676m and Re = 6.4 x 106m then


Re
F NN

2d

GM e m
2
Re

Re

- 1.492 x 10 - 4

1+

M
Me

(-2 )N
N =1 N.N!

- 954 1 - e- d + 1.512 x 10 - 3
1
7 N

2.56x 10

1
(1676 )N

Now

(-2 )N
N =1 N.N!

1
7 N

2.56x 10

1
(1676 )N

+2
5.966 x 10 - 4 1.193 x 10 - 3
1
+4 2
(-3.56 x 10 -7 - 3.56 x 10 -7 2
2.2
=
-8 3
(-2.124 x 10 -10 + 9.44 x 10 -11
3 7 3
+ 16 4
(-1.267 x 10 -13 2.11 x 10 -14
4 2 3 4

Thus, we have
F NN
F NN

2d
Re

GM e m
2
Re

2d

1-

3.548
1676

3.548
GM e m
12
d
Re
Re
- GM e m
F N (-d) =
2d
2
Re 1 Re

191

But the neo-Newtonian gravitational force at the bottom of a well of depth d


is given by
F NN (-d) =

- GMm
2d
2
Re 1 Re

- GM e m
2d
2
Re 1 Re

F NN (-d) =

1 - 3.548

Re

1 - 3.548

- GM e m
2d
2
Re 1 Re

Re

1 - 7.096

Re

which may be written


F NN (-d) = F N (-d) 1 - 7.096

Re

This gives the first order approximation of the deviation from


Newtonian gravitation predicted by the neo-Newtonian potential and shows
that the predicted gravitational force of the Earth decreases more rapidly
than Newtonian gravitation does.
5.5

Inertial and Gravitational Mass and their Equivalence

There are three ways in which mass appears in Newton's Second


Law when gravitational forces are considered. Consider his gravitational
force law which may be written
F = -

Gm1 m2
r
3
r

and his Second Law which is


F =

2
d
d r
(mv ) = m 2 .
dt
dt

In these equations there are the inertial mass, m, and two


gravitational masses, m1 and m2.
The force equation comes from
considering the force on m2 due to the the gravitational field of m1. In this
case m1 is usually referred to as the active gravitational mass while m2 is
the passive gravitational mass. Classically Newton's Third Law is imposed
in order to show that the ratios of active and passive gravitational masses
must be equal. Consider
192

F1 = - F 2

so that
Gm1a m 2p = Gm2a m1p

where the subscripts a and p refer to the mass's role as either an active or
passive gravitational mass. This leads us to the equation
m1p
m1a
=
m 2a
m 2p

which means that since the ratios must be equal the ma and mp may be
made equal.
The equality of inertial and gravitational mass is not predictable by
Newton's laws. Rather, it is taken as an assumption. This assumption
has been subjected to increasingly accurate experimentation by Eotvos in
the 1880's, by Dicke in 1964, and by Braginski in 1971. The present limit
of comparision between gravitational and inertial mass in about one part in
1012.
Now let's consider these same three mass concepts in the context of
the Dynamic Theory. First, there is the inertial mass density. It makes its
appearance in Section 3.1 when we impose the principle of increasing
entropy as a variational principle. The metric element is given in terms of
the specific entropy while the entropy principle is in terms of the entropy
density. The effect of this is to introduce the mass density as a product of
the acceleration into the equations of the force densities (see Eqn. (3.5)).
The same inertial mass concept leads to the Einstein energy and mass
relation in Section 3.2.
The other two mass concepts enter first through the field equations
given by Eqn.s (3.15) and from them the force densities in Eqn.s (3.17). In
Section 5.1 we went through the field equations to determine the chargeto-mass conversion needed to keep the units consistant. Here we found
that the passive gravitational mass given by Eqn. (5.12) was
M

mp

(5.76)
while the gravitational field associated with a gravitational mass is given by
Eqn. (5.10), when the evaluated parameters are used, as

193

V r = ma

11 + Gt

r
r

(5.77)
where the mass in the gravitational field equation is to be considered the
acitve mass and therefore we've used the subscript a to denote this. The
gravitational force due to the passive mass M being in the gravitational
field Vr is then
1F 12 = - Gm1p m 2a (1 + Gt)

2a

r
r

2a

r 12 .

(5.78)

By looking at Eqn. (5.78) we may see that, first, it is the active


gravitational mass that has the time dependence and not the passive
gravitational mass. Further, only when the active gravitational masses are
identical with their 's the same will the active and passive gravitational
masses be equal. We've used the subscripts 12 in the force equation to
denote the force on mass 1 in the pressence of the field of mass 2. We may
consider the force on mass 1 when in the field of mass 2 and we find
1a

1-

F 21 = Gm2p m1a (1 + Gt)

1a

r 12 .

(5.79)

If we form the ratio of Eqn.s (5.78) and (5.79) we find that


F 12 =
- F 21

m1p m2a

1-

m 2p m1a 1 -

1a

r
2a

1a

(5.80)

2a

Further, only for identical gravitational masses will Newton's Third Law be
satisfied within the Dynamic Theory.
5.6

Cosmology

The hot big bang model of the Universe is the model which is in
vogue now. Virtually all the journals print numerous articles relating to
some aspect of the hot big bang model. The model is based upon the
194

Newtonian Gravitational and the notion of a scale of the universe that is


changing with time. This notion is borrowed from Einstein's General
Theory of Relativity, however Einstein's theory is not used in the hot big
bang model itself. It would seem a shame to discuss a new gravitational
potential such as presented in this book without some discussion of its
possible impact upon the hot big bang model. It would have been
preferable to wait until the entire solution could be presented. However,
this is not possible now so this presentation will include a discussion of
how one might expect the new potential to impact the hot big bang model
and the problems that render the solution illusive.
The development of the standard big bang model begins with
considering a spherical piece of the universe with an observer at the
center. This sphere is considered to be filled with "dust" of density (t) with
a galaxy of interest placed at the outer boundary of the sphere which has a
radius denoted by x. When Gauss's law and Newton's laws of motion and
gravitation is used one arrives at
mg

2
4 x
d x
=2
3
dt

(t) Gm g
x

=-

4
G (t) m g x.
3

(5.81)

But consider what happens if one wishes to compare this with the nonsingular potential of the Dynamic Theory. Then Eqn. (5.81) becomes
2
4 x
d x
mg 2 = 3
dt

(t)G m g
x

1-

e x=-

4
G (t)x 1 ex .
3
x

Now let us replace x with the comoving coordinate x=R(t)r where R(t) is the
scale factor of the universe and r is the comoving distance coordinate as is
done in the standard model. When we also normalize the density to its
value at the present epoch, o , by (t)= oR-3(t) we obtain
2
4 G
d R
=2
3
dt

1-

e Rr
.
Rr R 2

We can begin to see the trend to be expected from the universe from
Eqn. (5.83) by noting that should we look back in time to the point when
R=r/ then we would have a point in time, say T1 when the acceleration of
the universe would have been zero. At times before T1 there would have
been an acceleration outward while for times after T1,such as the current
epoch, the rate of the expansion of the universe is slowing down. This is a
very different story than is told from by the standard model. But how is it
different? It is the same as the standard model in that from Eqn. (5.83)
one sees that the universe was forced into expansion at early times and is
now slowing down its rate of expansion. One big difference between the
story to be told by Eqn. (5.83) and the standard model is that Eqn. (5.83)
gives the reason for the initial expansion and it denies that the universe
195

(5.83)

was ever collapsed to a singular point as supposed by the standard model.


To better see the first contention we should proceed a little further.
If we multiply Eqn. (5.83) by dR/dt and integrate with respect to
2
R =

8 G
3

(t) R 2 e- Rr +

(t) R 2
- kc 2 .
2 c2

(5.84)

time we find
In Eqn. (5.84) we have included the term for the radiation for
completeness. Now let us evaluate the constant of integration, k, by
setting the values of R, dR/dt, , and at their present day values of 1, Ho,
o, and o. Then Eqn. (5.84) becomes
2

Ho=

e r+

4 G o
- kc 2 .
3 c2

(5.85)

If we now make the definitions


c

3 H o2
and
8 G

Eqn. (5.85) may be put into Eqn. (5.84) to obtain


2
2
R = Ho

Re

Rr

+1-

Ho
2 c2

-1 .

(5.86)

We may now take a look at some of the implied dynamics from Eqn. (5.86).
First look at the dynamics as R tends to infinity and there is no
radiation. For this case we would have
2
2
R = H o (1 -

),

which is the same as in the standard model.


Now suppose we look backward in time to the time when dR/dt was
zero? Then Eqn. (5.86) becomes
0=

Re- Rr -

2 R e r+

2
R +

2 c2

1 - R2 .

This is a transcendental equation which could be solved for R if we knew ,


r and the density of the dust and radiation at the current epoch. It may be
seen from Eqn. (5.87) that if there is no radiation and R does not equal
zero then

196

(5.87)

R=
-

Rr

er-

.
c
o

There is also a trivial solution at R=0 in Eqn. (5.87) but for this case the
acceleration is also zero and therefore no dynamics are allowed.
While at first glance it may appear that we have as good a developed
solution as is arrived at in the standard model, consider the following
points. Our galaxy was considered to be on the outside limit of a sphere of
dust. For the current epoch the density of the dust is very small locally to
the galaxy and Gauss's law for considering the total mass of the sphere of
dust to be placed at the center should hold very well. But what about
when we are looking back in time when the density was a lot greater. At
some density we are no longer able to approximate result used in Eqn.
(5.82) but must use the solution developed in the discussion of the Fifth
Force. Then our conclusions arrived at above are only good in a general
sense and are not quantitatively accurate.
A second point concerns the fact that we have developed the gauge
function in prior sections. If this is the scale of the universe as the gauge
function is supposed to be, then why are we again trying to solve for it
here? If the scale of the universe is given by the gauge function then the
dynamics may be over specified if we put the radiation into the equation for
the acceleration such as Eqn. (5.83). On the other hand, what is the
source for the radiation? If there is no hot big bang for the radiation to
come from where might it originate? The Dynamic Theory displays an
inductive coupling between the electromagnetic and the gravitational fields.
could the radiation be due to the expansion of the gravitating mass of the
universe? If so then a knowledge of the gauge function might turn the
equation of motion for our galaxy into a prediction of the radiation required
at the present time. This prediction might then be compared to the
measured radiation. However, there is the necessity to have a for the
universe. It may be obtained from the gauge function also as GM/c2. But
how is M determined?
Perhaps this is sufficient to point out that the overall picture of
cosmology to be given by the Dynamic Theory is not yet complete but in
any event will likely be very different from the hot big bang model of the
universe. Will it allow for high temperatures needed for accounting for the
abundances of the elements? Since it allow for the universe to be much
smaller in the past it would have the associated high temperatures. Yet it
should not have the infinite temperatures associated with a singular
universe.

197

(5.88)

Chapter 6 Electromagnetogravitic Waves


Given the system of equations, Eqn. (3.15), and the interpretations
that; E is the electric field, B is the magnetic field, V is the gravitational
field, and V4 is the gravitational potential, then the question arises as to
how the electromagnetogravitic waves may propagate.
6.1

Wave Equations

The usual assumptions such as / = / = 0 and = 0 may be


used to derive the wave equations for the four field quantities. Other
assumptions used are that the media is isotropic and that J = E and J4 =
4V4. The resulting wave equations are
2

V-

E-

B-

V
- 2
t
c
4
c

t
E
t

V
2

+ a02

B
- 2
t
c

a0 4
c
2

+ a02

+ a02

( E)

E
2

B
2

=0 ,

=0 ,

E
4

, (6.1)

(6.2)

(6.3)

and
2

V4-

V4
- 2
t
c

V4
+ a02
2
t

V4
2

=0 .

(6.4)

The inhomogeneous term in Eqn. (6.3) displays an interconnection


between the electric and gravitational waves. Further, this term produces
the question of whether the propagation vector for the gravitational wave
can be the same as the propagation vector for the electric wave. In
Maxwellian electromagnetism it may be shown that the propagation vector
for the electric wave must be the same as the propagation vector for the
magnetic wave. However, this is not true, in general, for this system of
equations.
6.2

Wave Solutions

Given that the propagation vectors may be different a trial solution


may be sought during which the conditions for identical propagation
vectors may be exposed.
If the waves are considered to be propagating in the positive xdirection, then the trial solutions may be taken to be of the form

198

E = E 0 exp - i t - k e x - k 4e

B = BO exp - i t - k b x - k 4b

V = V O exp - i t - k v x - k 4v

and
, (6.5)

V 4 = V 40 exp - i t - k 4 x - k 44

By making the definitions


Ae = a0 ck 4e
Ab = a0 ck 4b
Av = a0 ck 4v
A4 = a0 ck 44

,
,
,
,

(6.6)

and substituting the trial solutions, Eqn. (6.5) into the wave equations,
Eqn.s (6.1)- (6.4), we obtain the indicial relations:
2
k v0 c =

+ i4

- Av2 + i4

Ae

2
e

2
b

+ i4

k e c = (kc ) - A

k b c = (kc ) - A

i
k 4e

E
V

(6.7)

and
2
k4 c =

- A42 ,

where

2 + i4
= 1, 2, 3, and (kc)2 =
.
Substituting the trial solutions into the continuity equation of Eqns.
(3.14), we find

-i

k 44

(6.8)

The ratio (E /V ) appearing in Eqn. (6.7) indicates that we need to


know the relationship between the components of the V field and the
components of the E field.
These relationships may be found by
substituting the trial solution into each of Eqn. (3.15). In this substitution,
the further limiting assumption that the electric field may be polarized so
the Ez = 0 and Ex = 0 is made in order to simplify the solution. It should be
pointed out though that, in contradistinction with Maxwellian
electrodynamics, Ex is not required to be zero by the differential equations.
The differential equations require the following relationship among
the non-zero components, given the trial solutions and the imposed
restrictive assumptions:
Bz =

ke c

E y , (6.9)

and
V y=

- Ae

Ey=

Ab

ke
Ey .
kv

Thus, the imposed assumptions reduce the solution to only three non-zero
components, Ey, Bz, and Vy.
If we consider the different expressions, from Eqn. (6.9), for Bz, and
take the partial derivative with respect to the mass density we find
199

ke c
Bz

Ey

requires that
1

ke

ke

(6.10)

Ab - Ae .

a0 c

From Eqn. (6.9) we also find


(6.11)
while differentiating with respect to -produces
Ab k e = Ae k v ,

Ae

Ae

(6.12)

Av - Ae .

a0 c

With the assumption that there is no longitudinal field component


the surviving system of equations is
2
2
2
k e c = (kc ) - Ae ,

(6.13)

2
2
2
k b c = (kc ) - Ab ,

2
2
2
k v c = (kc ) - Av +

ke

ke
1
Ae

Ae

Ae

Ab
ke ,
Ae

kv =
1

a0 c

=
=

(6.14)

Ab - Ae

a0 c
i

Av - Ae

a0 c

where the definition


(kc )2 =

(6.15)

+ i4

has been used.


The partial derivative, / , appearing in Eqn. (6.14) may be found
from experiment in the following manner; because
T

where T is the temperature. The conductivity, , is the reciprocal of the


resistivity, , whose linear dependence upon the temperature is given by
r = r0 1+

T -T0

Then
T

1
r

=-

r0

On the other hand, the coefficient of volume expansion is defined as


=

1
(vol.)
,
(vol.)
T

but
200

mass
,
vol.

Therefore,
=

Thus,
=

r0

. (6.15)

The solution of Eqn. (6.15) is


r0

=
ln

where

(6.16)
0

is the mass density evaluated at the temperature T0.


Given the expression for
/ , it may be seen that the system of
Eqns. (6.13) and (6.14) represent six-equations in the six unknowns, Ae,
Ab, Av, ke, kb, and kv. The system may be solved as shown in the following.
The unknown, kb, appears only in one equation; therefore, this
equation may be considered to determine kb once the solution for the other
unknowns are determined.
Eqn. (6.11) may be used to eliminate Ab, in Eqn. (6.10) leaving us
with four equations in four unknowns. Eqn. (6.10) now looks like
0

kv = ke -

ia0 c
Ae

ke

(6.17)

Eqn. (6.13) may be used to determine Av by rewriting it as


Av = Ae -

ia0 c
Ae

Ae

(6.18)

This may be used to eliminate Av from the third of Eqn.s (6.7) leaving three
equations in three unknowns.
By differentiating the first of Eqn. (6.7) with respect to the mass
density, it becomes
2 k e c2

ke

= i4

- 2 Ae

Ae

(6.19)

Substituting Eqns. (6.17) and (6.18) into the third Eqn. (6.7) and using
Eqn. (6.19) results, after some manipulation, in
2

(kc )

1
Ae

Ae

- 2 i2

Ae

Ae

This is a quadratic equation in Ae/


Ae

+ i2

Ae

(kc )2

= 0.

whose solutions are

i2
=

Ae + _

or
201

Ae - (kc )

(6.20)

Ae

i2
(kc )2

Ae + _

Ae - (kc )

Therefore we have
dAe
Ae + _

Ae - (kc )

d(kc)
.
(kc)

(6.21)
But from the definition of (kc)2 we find that
d(kc )2 = i4

Thus,
d(kc )2 2(kc)d(kc) d(kc)
i2
d
=
=
=
.
ki
(kc )2
2(ki )2
2(kc )2

Eqn. (6.21) now becomes


d Ae
Ae + _

Ae - (kc )

d(kc)
.
(kc)

(6.22)
By using the method of substitution, recognizing that it may be put
into a homogeneous form, and realizing the solution may be complex, we
arrive at the solution of Eqn. (6.22) as
Ae =

1
(kc )2 c 22 - 1
2 c2

(6.23)
where c2 is a constant of integration such that
c2

=0 ,

that is, c2 may depend upon , , and but not .


Eqn. (6.23) is a quadratic equation in c2 and may be solved yielding

202

Ae + _

c2 =

2
2
AE + (kc )
.
(kc )2

(6.24)
Because c2 does not depend upon
Aeo = Ae ( = 0) then
c2 =

it is unaffected by setting
2

Aeo + _

= 0, so if

Aeo +

(6.25)
By substituting Eqn. (6.25) back into Eqn. (6.23) we find the sign before
the radical must be taken to be positive. Then the expression for Ae
becomes
Ae = Aeo + i4

h .

(6.26)
where
h

Aeo
2

1
2
2 Aeo + Aeo
+

Using Eqn. (6.26) in the system of equations, Eqn. (6.14), we may,


after a great deal of algebra, write the total solution as:
Ae = Aeo + i4

h ,

(6.27)
where
h

Aeo
2

1
2 Aeo + Aeo +

Now we may write


ke=

with

203

+i

4 ~
1+ ~

2
~

Aeo

1-

+1

4 ~
1+ ~

~=

1
2

~=

-1

h )2

(4

+
2

1
2

(6.28)

1 - 2 hAeo .

Now we have
Av =

av

+i

av

where
= Aeo (1 + f) ,
h (1 - f) ,
av = 4
av

[ a0 c 4
f=

(6.29)

h
.

h )2

Aeo + (4

The propagation vector for the v component may be written as


kv =

with

1+

c
1=

2
av

1+

2
-

1-

2
av
2

+1

,
1
2

Aeo f

2 a av
4

1
2

av

-f

h
.

(6.30)

204

Now we have
Ab =

ab

+i

ab

with
= Aeo D - 4
ab = Aeo F + 4

hF ,
hD ,

ab

D=

F=

e
2
e

v
2
e

v
e

v
2
e

Then
kb =

+i

(6.31)
where

1+

4
1+

2
=

11-

2
4

2
av

av

2
av
2

av

1
2

+1

,
1
2

-1

-f

The system of equations, Eqns. (6.26) through (6.31), representing


the solution is an extremely complicated system and should be put on a
computer in order to fully explore all the ramifications of this solution.
6.3

Non-thermal Transmission through Media

It is the intent of this section to briefly show the effect of the solution
given above and discuss how this solution may be useful in modeling
electromagnetic interactions with biological systems. Therefore, consider
the question of component attenuation, or how the different components of
the electromagnetogravitic wave may be attenuated? A simpler question
would be, "For what frequencies will the components not be attenuated at
all?"
205

From Eqn. (6.28) we find that the electric component will pass
unattenuated if =0. This is satisfied by two conditions. The first
condition is that
= 0 which is the classical condition of a perfect
dielectric. The other condition is that
h=

1
2 Aeo

(6.32)

Substituting the definition for h into Eqn. (6.32) we find, after some
manipulation, that this is satisfied if
2

Aeo

Aeo

Aeo

- 3= 0 .

(6.33)
which has only one real solution
2
c

= 1.7971 A2ae .

The complex solutions are:


2

that

= (-0.8985 + _ i 1.0434) Aeo .

Considering the real solution and assuming Aae2 to be real, we find

1.7971

Aeo .

We do not yet know the dependence of Aeo upon , , . The assumption


that Aeo is linear in
would mean that the relative strength of the
gravitational component compared with the electric component, given by
V y=-

Ae

Ey ,

does not depend upon frequency in free space.


assumption that
Aeo =

In Eqn. (6.33), the

(6.34)
206

implies that there are no frequencies for which


consistent with classical theory.
If now we look at the frequencies for which
(6.29), that v = 0 when ' = 0, or
1-

2
4

av

av

= 0, and this would be

e
v

= 0, we find, from Eqn.

=f .

(6.35)
Substituting for the defined quantities in Eqn. (6.35), assuming 2<< ,
and disregarding negative frequencies, we find two possible frequencies for
which v = 0, or
- 16

1
c1

1
2

a0 c
2

(6.36)
and

1
2

4 ao c

c2

1-

2+

1-

The magnetic component is unattenuated when


4

=2

ab

ab

= 0, or when

(6.37)
The condition specified by Eqn. (6.37) represents a seventh order
polynomial in , therefore, the roots of this polynomial have not been
sought. It may be noted though that there are up to seven possible
frequencies for which the magnetic component is unattenuated.
Thus, for frequencies satisfying the conditions of Eqns. (6.35) and
(6.37), the gravitational or the magnetic component respectively will
experience no attenuation. Because these conditions result in polynomials
in , then there must be frequency regions where either v < 0 or b < 0, or
both v and b are negative. In these regions the gravitational and/or the
magnetic component will experience an amplitude growth.
On the other hand, from Eqn. (6.32), we found that there were no
frequencies for which e < 0 for > 0. This then leads to the possibility
207

that the growth in the gravitational and/or magnetic component is at the


expense of the electric component.
If then, non-thermal transmission is defined to be transmission
during which none of the wave energy is deposited in the media, we find
that our simple solution will support non-thermal transmission for
frequencies satisfying Eqn. (6.33). For this type of transmission the energy
originally in the electric component experiences an attenuation and is
transferred to the gravitational and/or the magnetic components which
experience a gain. The net result is the transmission of energy through the
media without loss, only a change in form.
We have seen that the three fundamental postulates of the Dynamic
Theory have led to the use of mass density as a fifth dimension,
fundamentally independent of space and time. The five-dimensionality of
the theory produced the eight differential equations describing the allowed
interrelationships between the five dimensional gauge fields.
Other
investigations in fields allowed for fundamental particles produces the
interpretation of the V field as the gravitational field.
With the interpretation of the V field came the question of how these
waves might propagate given their specified interrelationships. In answer
to this question we've shown that for simplified, continuous media there
exist frequencies for which the electric component is attenuated while gain
in the gravitational and/or magnetic components is experienced. This
gives rise to the possibility of transmitting energy through the conductors
where no such energy transmission is allowed by Maxwellian
electromagnetism.
Even though biological systems are complex structures, is it not
possible that the five-dimensional fields of the Dynamic Theory have
applicability in describing radiation interaction with these systems? Is it
possible that a description of non-thermal effects of radiation on biological
systems may be aided through the use of the non-thermal transmission
effects discussed here? A great deal of discussion these days concentrates
on nonlinear approaches. The five dimensional waves provide a linear
description of effects which in four dimensions would appear as nonlinear.
Thus, it would seem possible to replace nonlinear four-dimensional
problems with five dimensional linear ones.
6.4

Boundary Conditions

Classical work on boundary conditions of field vectors generally


starts with Coulombs' Law. Here polarization of materials enter the
picture. Polarization can be caused by either alignment of molecules or
induced asymmetry. From these considerations the electric dipole moment
is defined as
p = ql

208

where p is the electric dipole moment and l is a vector from -q to +q. The
net dipole moment per unit volume is the polarization, P, of the medium.
From this we get
P dvol = q
vol

where q is the net polarization charge within the volume. If the density of
the polarization charge is 'l then we have
q =-

dvol
vol

where the minus sign arises since by definition, the direction of the
polarization vector is from negative to positive, whereas the electric field is
directed from positive to negative. Thus, we arrive at
P= -

Now in order to write an equation like


E=4

that is valid in a dielectric medium and account for both free charges and
polarization, and ' respectively we must write
E=4( +

) ,

(6.38)
or using Equation (3.53)
( E + 4 P )= 4

Maxwell named
displacement, or

the

quantity

in

parenthesis

the

dielectric

D= E +4 P .

Therefore, the equation for a dielectric media becomes


D= 4

From experiment it is found that a large class of media exhibit P


proportional to E, for field strengths not too great. Thus,
P=

(6.40)
209

where xe is the electric susceptibility of the medium.


Then
D = (1 + 4

)E .

The proportionality factor between D and E is called the dielectric constant


and
= 1+ 4

Therefore,
D= E .

Thus, Equation (6.39) may be written as


( E )= 4

(6.41)
Consider now that in the Dynamic Theory we derived the equation
( E )= 4

- ao

( V4)

(6.42)
Thus it may be seen by comparing Equation (6.41) with Equation (6.42),
that the second term on the right hand side plays a role of gravitational
polarization charge density, or
ao

( V4)

Pg .

(6.43)
Then we would have
( E + 4 P g )= 4

or
(1 + 4

)E + 4 P g = 4

or
E + 4 P + 4 Pg = 4

(6.44)
210

Thus, in order to include the gravitational polarization the dielectric


displacement must be given by
D = E + 4 P + 4 Pg .

Now the dielectric polarization, P, is an averaging over a finite


volume, thus when using the Gaussian pillbox in looking at boundary
conditions it is assumed that the pillbox may contain free charges but not
polarization charges. Thus, when considering boundary conditions we
must look at the displacement vector D not the field strength E.
Therefore, consider the usual "Gaussian pillbox" of cross-sectional
area S and thickness t. Let n be the unit normal to the surface S. The
pillbox volume is V = S t and is assumed to contain free charge but no
polarization charge, nor gravitational polarization (we may want to rethink
this assumption concerning gravitational polarization when it is better
understood.) If we integrate over the volume V we have
D dv = 4

dv

or, by using the divergence theorem,


D da= 4
s

dv .
v

The left hand side may be integrated by noting that since the normal
component of D is involved there is no contribution from the sides. Thus,
since the volume V can be made sufficiently small, we have
D 2 n 2 + D1 n 1 = ( D 2 n 2 - D1 n 1 )S = 4

t .

If
lim
(
t 0

t) = s 57; or free surface charge density, then


( D 2 - D1 ) n = 4

(6.45)
relates the charge in the normal component of D across a boundary to the
surface density of free charge on that boundary. If s = 0 then the normal
component D is continuous across the boundary. Equation (6.41) may
also be written as
(

E 2 + 4 P g2 -

E 1 - 4 P g1 ) - n = 4

211

s .

This points out the need to consider the physical meaning of the
gravitational polarization but we won't go into that at this time.
The next condition that must be fulfilled at the boundary comes
from
xE = O

for static fields so that B/ t = 0. (This may safely be assumed since even
for non-magnetostatic field the contribution by the B/ t term vanishes in
using Stokes theorem).
Now construct a rectangular path which has sides 1 width t, and
for which the sides parallel a segment of the bounding surface. Then by
Stokes' theorem
E .d l = ( .E ). n o da = 0 .
s

Thus,
E .d l = ( E 1 . n 1 ) 1 + ( E 2 . n 2 ) 1 + contribution from ends.

Now,
n 1 = - n 2 , 63 therefore
( E 2 - E 1 ). n 2 1 + (ends) = 0 .

Since the contribution from the ends is proportional to t, the second term
vanishes in the limit as t
0. Thus we have
( E 2 - E 1 ) - n 2 = 0 ,
n0 .( E 2 E 1 ) x n = 0 .

Now
n 2 = n 0 x n

so that
( E 2 - E 1 ).( n 1 x n ) = 0

or
n0 .( E 2 E 1 )xn = 0 .

212

But the orientation of the rectangle is arbitrary so that


( E 2 - E 1 ) x n = 0 .

(6.46)
Eqn. (6.42) implies that the tangential component of E must be
continuous.
For the magnetic induction field, since here
B=0

which is the same as in the classical case we would have


( B 2 - B1 ) n = 0

or the normal component of B is continuous across the boundary.


For the tangential component we must consider the possibility of
another source term as we did for
E . 73 For an Amperian loop current I, a directed loop area S, the magnetic dipole
moment is defined by
m=

IS
.
c

Averaging over a volume we obtain the magnetization, M, which is the net


dipole moment per unit volume,
M=

dm
.
dv

From this the Amperian current density becomes


J =c

xM .

Now, in the classical case, we have, for electrostatic fields,


x B=

4
(J+J )
c

where J is the Amperian current density. Thus, we could write


x ( B - 4 M )=

4
J .
c

We can then define the magnetic intensity vector as


H =B-4

213

M .

Again, experimentally, numerous materials are found such that


M=

is a good approximation for the magnetization for small fields. xm is called


the magnetic susceptibility. Thus we would have
B = (1 + 4

where

)H= H

is the permeability of the material.


Now in the Dynamic Theory we have, when
x ( B/ ) =

/ t = 0,

4
(V / )
J - a0
.
c

(6.47)
If we define a gravitational magnetization by
4

Mg

- a0 V

then Eqn. (6.43) becomes


x

-4

Mg

4
J
c

and the magnetic intensity vector should be defined by


H=B-4 Mg

or
H =B-4 (M +Mg ) .

(6.48)
It may be seen from Eqn. (6.44) that the gravitational magnetization
adds to the Amperian magnetization and could lead to misinterpretations.
Now, in an analogous fashion to the dielectric displacement vector,
the boundary condition becomes
( H 2 - H 1 ) x n =

4
c

where K is the surface current density according to

214

K = lim ( J t)
t 0
J

Thus, if there is no surface current density, then


( H 2 - H 1 ) x n = 0

or the tangential component of H must be continuous across the


boundary.
Now we need to do a similar thing for V and V4. Starting with the
equation
.

1
c

( V4)
4
=J4
t
c

we will look at the gravitational field defined at


G=

where

is the gravitational charge-to-mass ratio. Thus, we have


V4
.G +

1
c

=-

4
c

J4

where the quantity (J4/ c) is the free gravitational mass density.


Now by using the divergence theorem we have
V4
1
G .d a +
c
s

dv = - 4
v

J4
dv .
c

By using a Gaussian pillbox again, then for a sufficiently small box


( G 2 . n 2 + G 1 . n1 ) +

S t V4
=-4
c
t

J4
S t .
c

(6.49)
For free gravitational mass density to exist on the boundary, the product J4
t must remain finite as t
0. Therefore, in the limit
lim J 4 t
= ms .
t 0
c

215

Then Eqn. (6.45) becomes


( G 2 - G 1 ).n = - 4 m s .

Thus, if there is no free gravitational mass on the surface, the normal


component of the gravitational field must be continuous.
Question: The relationship between G and V is similar to that between H
and B. Is it fair then to consider a similar behavior between them? By this
I mean since
B= H = H +4

H = H +4

G=G+4

then can
V

= G=G+4

where g may be called the gravitational susceptibility and N is yet


unnamed?
Now lets look at
x V = - ao

For this we construct the closed rectangular path across the boundary.
Then we would have
V .d l = (

x v ). n0 d a = ao

( B. n0 )da
s

Where S is the rectangular area 1 T and no is the unit vector normal to


the rectangle and lies along the boundary of the surface. Performing the
line integral, we obtain
( V .d l ) = ( V 1 . n 1 ) 1 + ( V 2 . n 2 ) 1
+ (contribution from ends)
= - ao

. n 0 1 t

But since
n 1 = - n 2 102 and the contribution from the ends is proportional to t we have, as t
( V 2 - V 2 ).( n 0 x n 1 ) = 0

or
n 0 . ( V 2 - V 1 ) x n = 0 .

216

0.

But, since the orientation of the rectangle is arbitrary, then


( V 2 - V 1 ) x n = 0

This states that the tangential component of V, the gravitational induction,


must be continuous.
The boundary condition for the scalar, V4 is that V4 must be
continuous because it is a scalar.
It perhaps should be noted that the new physical notions that
appear in the foregoing could prove extremely important should one
consider going into materials development.
6.5

Reflection and Refraction


First we shall consider normal incidence as shown in Figure (15)

Figure 1. EMG wave propagating in the z-direction.


have:

Applying the boundary conditions on the tangential components we


1

0
0
0
E 0 eik 4eo - E 1 eik 4 E1 = E 2 eik 4e2
o

0
0
H 0 eik 4bo + H 1 E ik 4b1 = H 20 eik 4b2
o

and
0
0
0
V 0 eik 4vo o - V 1 eik 4v1 1 = V 2 eik 4v2

or by using

1= o

we have
0

E 0 - E 1 = E 2 e k 4e2 2 k 4e1 1
0
0
0 i
H 0 - H 1 = H 2 e k 4b2 2 k 4b1 1

(6.50)
and
0

V 0 -V 1 =V 2 e

i k 4v2

2-

k 4v1

But
H=

B 4 Mg
,

and
217

By =

ke c

Ex .

Therefore, if we assume for the moment that gravitational magnetization is


zero, or
M g = 0, 112 then
0

H0=

k e1 c

0
E0 .

Note: This assumption places some, perhaps severe, restrictions upon


V / 114 and we will have to come back and look at these, but for now it seems like a
reasonable assumption to allow us to proceed with reasonable simplicity.
Thus, from the first two of Eqn. (6.50) we have
0

E0 - E1 = E 2 e

(6.51)
and
o

E o + E1 =

ke 2 o i
E2 e
ke1

- k 4e1

Where
k 4e2

By adding Eqn. (6.51) we find, after rearranging the terms,


0

E2 =

2 k e1 e-i e 0
E0 .
k e + k e2

(6.52)
On the other, by subtracting Eqn. (6.51) and rearranging, we have
0

E1 =

k e2 - k e1
k e1 + k e2

0
E0 .

For the solution sought in the non-thermal biological section


Ez=V4=Vz=0. Thus, if we stay with that solution we need only look at Vx,
but
V x+

- a0 ck 4 e

Defining
218

Ex .

Ae = a0 ck 4e ,

then
V x=

- Ae

Ex .

Thus, the last of Eqn. (6.50) becomes


0

E1 - E0 =

- Ae
Ae

0
E2 .

(6.53)
Using Eqn. (6.52) this becomes
0
E1 = 1 -

2 k e1
Ae2
0
E0
Ae1 ( k e1 + k e2 )

(6.54)
Comparing Eqn. (6.54) with Eqn. (6.53) we find
- k e1 = k e1 -

Ae2
Ae1

2 k e1

which is only satisfied only if


Ae1 = Ae2 .

(6.55)
Equation (6.55) implies that the dependence of the electric field upon
mass density is not influenced by the type of material there. This is a
result that is a direct consequence of the assumption previously made and
is further evidence that we must return to that assumption soon. For now
we shall forge ahead.
We shall now consider the case where the incident wave impinges
upon the boundary interface at an oblique angle o. The wave is polarized
so that the electric component is parallel with the interface.
For the incident wave we have

219

- i(wt - k eo r - k 4eo

- i(wt - k b0 r - k 4bo

E0 = E0 e

H0= H0 e
c

1)
1)

k eo x E 0

1
o

and V 0 = V o e-i(wt - k vo r - k 4vo


- Aeo
=
E0 .

For the reflected waves


0

E1 = E1 e

- i(wt - k e1 r - k 4e1

H=

1)

k e1 x E 1

and V =

- Ae1

E1

(6.56)
The refracted waves are given by
0

E 2 = E1 e
H2=

- i(wt - k e2 .r - k 4e2

k e2 x E 2

and V 2 =

- Ae

(6.57)
The tangential components of , H, and V, can be continuous across
the boundary only if the phases of the field vectors are all equal at the
interface.
k eo r + k 4eo
k bo

r + k 4bo

k vo r + k 4vo

= k e1 r + k 4e1

1 = k b1 r + k 4b1
1

= k v1 r + k 4v1

= k e2 r + k 4e2

1 = k b2 r + k 4b2

= k v2 r + k 4v2

(6.58)
For each component the propagation vectors keo, ke1, and ke2 are
coplanar, so if r is chosen to lie in the interface and in the plane of the
propagation vectors, then we have,
220

k eo sin

eo

+ k 4eo

= k e1 sin

e1

+ k 4e1

= k e2 sin

e2

+ k 4e2

For keo-ke1 we find


sin

= sin

e1

eo

k eo

k 4e1 - k 4eo .

From this we find that sin eo = sin e1 if 1 = 0 or if


k 4e1 = k 4eo . 133 Not wanting to restrict ourselves un-necessarily by assumptions, lets
continue.
For other components we have
sin

b1

= sin

bo

sin

= sin

V1

k bo

k 4b1 - k 4bo

and
1

k vo

k 4 v1 - k 4 v 2 .

Again using Eqn. (6.58) we find


k e1 sin

e1

- k e2 sin

e2

= k 4 e2 2 - k 4 e1

(6.59)
and
k eo sin

eo

- k e2 sin

e2

= k 4 e2 2 - k 4eo

(6.60)
However, for keo = ke1, subtracting Eqn. (6.60) from Eqn. (6.59) yields
1

( k 4eo - k 4e1 ) = 0

so that we have as a required result


k 4eo = k 4e1 .

(6.61)
In a similar fashion we have
k 4bo = k 4b1

and
k 4vo = k 4v1

221

With this result Eqn. (6.59) becomes


sin

e2

k e1
sin
k e2

- ( k 4e2
e1

- k 4e1

k e2

(6.62)
Similarly, for the other components we have
sin

b2

k b1
sin
k b2

- ( k 4b2
b1

- k 4b1

k b2

(6.63)
and
sin

v2

k v1
sin
k v2

- ( k 4v2
v1

- k 4v1

k V2

(6.64)
Because of Eqn. (6.57), we must have
eo

e1

b1

b1

vo

v1

Now the tangential components of , H, and V must be continuous


at the interface. Therefore,
( E o + E 1 ) x n = E 2 xn ,
( H o + H 1 ) x n = H 2 x n ,

(6.65)
and
( V o + V 1 ) x n = V 2 xn .

Eqn. (6.65) may be written in terms of the electric component then we


would have
( E o + E 1 )xn = E 2 xn ,
( k eo x E o ) + ( k e1 x E 1 ) xn =

( k e2 x E 2 )xn ,

(6.66)
and
( Aeo E o + Ae1 E 1 )xn = Ae2 ( E 2 xn ).

222

From the first and last of Eqn. (6.66) we have


( Aeo E o - Ae 1 E 1 )xn = Ae2 ( E o + E 1 )xn

or
( Aeo - Ae2 ) E o + ( Ae1 - Ae2 ) E 1 x n = 0.

(6.67)
Since both o and
Eqn. (6.66) requires that

are perpendicular to the plane of incidence then


( Aeo - Ae2 ) E o = - ( Ae1 - Ae2 ) E 1 .

But, since Aeo = Ae1, this can be satisfied only if


Ae1 = Ae2 .

We've seen this result before.


Now suppose we expand the triple cross products in Eqn. (6.62),
then
(n k eo ) E o - (n E o ) k eo + (n k e1 ) E 1 - (n E 1 ) k e1 =
(n k e2 ) E 2 - (n E 2 ) k e2 .

1
2

All the products


(n E ; ) 155 vanish, and
n k ej = (-1)j k ej cos ej ; 156 j = 0, 1, 2, so that
k eo cos

eo

E o - k e1 cos

e1

E1 =

k e2 E 2 .

If we use the fact that

eo

e1,

and rearrange the terms we arrive at

( E oo - E 1o ) cos

eo

k eo
cos
k e1

e2

i( k 4e2

2 - k 4e1 1 )

(6.68)
Since the electric vectors are all parallel to the boundary surface, we
must have
O

O i( k 4e2

EO + E1 = E 2 e

2 - k 4e1 1 )

(6.69)

223

We may combine Eqn. (6.68) and (6.69) by subtraction to eliminate E2o and
obtain
cos

E1 =

cos

eo

eo

k e2 cos

e2

k e2
cos
2 k e1

0
E0 .

e2

(6.70)
Eliminating EO we have
0

E2 =

2 cos
cos

eo

eo

i( k 4e1 1 - k 4e2

k e2
cos
2 k e1

e2

(6.71)
Equations (6.62) and (6.63) may be used to determine the refracted
angles for each component while Eqns. (6.69) and (6.70) determine the
magnitude of the reflected and refracted electric field components. The
magnitudes of the reflected and refracted magnetic and gravitational
components may be found using Eqns. (6.56) and (6.57).
6.6

Complex Refraction Angles

In order to solve the refraction problem when one, or both, mediums


at an interface are conductors then we must have a computer code capable
of solving the complex angle problem. Thus, we must learn how to
interpret the complex angle of refraction then how to compute it.
We may start with the equations already derived for the sine of the
refraction angles, which are
sin

2=

k e1
sin
k e2

e1

( k 4e2

- k 4e1

k e2

(6.62)
sin

b2

k b1
sin
k b2

b1

( k 4b2

- k 4b1

k b2

(6.63)
and
sin

v2

k v1
sin
k v2

v1

( k 4v2

- k 4v1

k v2

(6.64)
224

With the realization that one or more of the k's in Eqns. (6.62),
(6.63), or (6.64) may be complex then one must consider the right-handside to be complex. Thus, we have the situation that sin 2 for each
component, would be complex and, therefore, we must consider 2 to be
complex. Thus consider the case where
sin

= x + iy

(6.72)
with
2

+i

Then, by trigometric identity, we may write


sin

= sin

cosh

+ i cosa

sin h

(6.73)
Equating Eqns. (6.72) and (6.73), we find
sin

cosh

+ i cos

sin h

= x + iy .

(6.74)
From Eqn. (6.74) we find that
sin

x
cosh

sinh

and
y = cos

(6.75)
Now 2 is a real angle and the expression for sin 2 in Eqn. (6.75)
reduces to the usual expression for the sine of the refraction angle. Thus,
we shall take 2 to be the angle of refraction and it is given by Eqn. (6.75).
We must now learn how to find 2 and 2 given x and y. We may
start by rewriting Eqn. (6.75) as
sin

x
cos h

(6.76)
and
225

cos

y
sin h

.
2

Now using Eqn. (6.76) the trigometric identity


1 = sin 2 + cos 2

becomes
2

x
1=
cosh

y
+
sin h

or
2

2x

1=

e +e

2y

e -e

(6.77)
when cosh 2 and sinh 2 are written in terms of exponentials.
Now suppose we define
w = cosh (2

2 )=

+ e-2
2

(6.78)
where w

1 always. Then Eqn. (6.77) becomes


1=

4 x2
2

( e2 + e- 2 + 2)

4 y2
2

( e2 + e- 2 - 2)

or
1=

2 y2
2 x2
+
w+1
w-1

This may be rewritten as


2
2
2
2
2
w - 2w ( x + y ) + 2( x - y ) - 1 = 0 .

(6.79)
Equation (6.79) is a quadratic equation in w which has the solutions

226

w = ( x 2 + y 2 ) + _ ( x 2 + y 2 - 1 )2 + 4 y 2 .

(6.80)
The expression under the radical is non-negative (as may be shown by a
lengthy procedure) so w is real, which must be the case from Eqn. (6.78).
Consider three cases:
Case A: x2 + y2 > 1
Let x2 + y2 = 1 + where
0, then Eqn. (6.80) becomes
w= 1 - + _

+ 4 y2 .

If y = 0, then because w 1, the positive sign must be used. If y = 0, then


there is no need for Eqn. (6.80).
Case B: x2 + y2 1
Let x2 + y2 - 1 + where
0, then Eqn. (6.80) becomes
w= 1 - + _

+ 4 y2 .

Again, because w 1, clearly the upper sign must be chosen.


Case C: x2 + y2 = 1
Equation (6.76) now becomes
w = 1 + _ 2y

If y 0, the upper sign must be used. If y 0, there is no need for Eqn.


(6.80).
From Eqn. (6.80) and the above logic on how to choose the proper
sign we may obtain w from x and y. Then from w we find 2 using Eqn.
(6.78) or
2

arc coshw
2

(6.81)
then by Eqn. (6.72) we may find

from

= arc sin

x
cosh

.
2

(6.82)
227

Thus Eqn.s (6.80), (6.81) and (6.82) give us 2 and 2 for any x and
y. The 2 and 2 must be checked against Eqn.s (6.75) and (6.76) to
resolve any ambiguities.
6.7

Assumptions and Wave Solutions

In order to try to bring to light the significance of the variation of


material properties with respect to changes in the mass density in the
attempt to obtain wave solutions let us consider each assumption that
must be made. To do this let us begin with the form of the trial solution.
The classically assumed form for a trial solution is
exp - i(wt - kx)

for plane wave propagation in the x-direction. In as far as it may be


possible we would like to stay with a similar form. Thus suppose we try
the form
exp - i(wt - kx - k 4 )

Once the form of the trial solution is chosen then one can look at the
eigenvalues of the differential operations since the trial form is the
exponential form. Consider the partial derivative with respect to x
x
= i[

{ exp[-i(wt - kx - k 4 )]}

(kx - k 4 )] exp[-i(wt - kx - k 4 )] ,

if we assume that
w
=0
x

Since we desire to consider how waves of a certain frequency propagate it


seems appropriate to adopt this assumption.
Next, consider the eigenvalue of the differention
i (kx + k 4 )
x
x
k
k4
= i{x + k +
+ k4
}
x
x
x

There is nothing thus far in our considerations that forces us to


consider only those cases for which the phase is strictly linear in x, or
where the density constant cannot depend upon space. In the classical
case I believe one can use the Maxwell equations, or the wave equations
coming from them, to show that the phase must be linear in x. The same
may prove true for these five-dimensional waves but it has not yet been
228

proven. For the sake of simplicity let's make the same assumption here
and hold in abeyance any attempt to prove a linear relationship.
Therefore, lets assume
k
=0
x

With regard to k4 we have no precedence set by classical theory and


with no real feeling for the physical interpretation of k4 we are left to our
own devices. For the moment suppose we make no assumptions regarding
the dependence of k4 upon space, but we might assume isotropy is mass
density. Thus, our eigenvalue for the space differential operator we have
x

k4
)
x

i(k +

Now let's look at the mass density differential operator, or


{ exp [-i( t - kx - k 4 )]}
=i

(kx + k 4 ) exp [-i( t - kx - k 4 )]

if we assume
=0 .

(6.83)
The assumption that the frequency should not depend upon the
mass density seems justifiable since we want to determine how a wave of a
certain frequency will propagate. Therefore, we want to control the
frequency. We should not, however, allow this desire to lock us into this
assumption.
Given the assumption Eqn. (6.83) we have
(kx + k 4 ) =

x k

k4

+ k4 .

By analogy with the classical result that the phase of the wave depends
only linearly upon x it seems a fair assumption that we may simplify some
by assuming that
k4

=0

We have no real justification for this assumption at the moment though.


However, with this assumption our mass density operator becomes
229

i(x

+ k4 ) .

If we consider the potential variation of k with respect to mass


density then we run into the dependence of k upon and and whether
these material properties depend upon mass density. If they do, as
experiment tends to show then
k

Since we desire to retain the correspondence to experiment we shall make


no assumptions concerning the dependency of k upon mass density.
The remaining operator is the time operator. For this we have
t

{ exp | -i( t - kx - k 4 ) |} = - i exp | -i( t - kx - k 4 ) |

if we assume that
k
=0
t

and
k4
=0 .
t

From the point of view that both k and k4 are determined by material
properties then these assumptions appear appropriate for static materials.
Thus the eigenvalue of the time differential operator is
t

-i

We have now chosen a general form for the solution we will seek.
But there are several potential components to the wave. For example,
there are the electric transverse, magnetic transverse, gravitational
transverse, electric longitudinal, and the gravitational longitudinal
components.
In addition there is the scalar wave component, the
gravitational potential.
In the classical case it may be shown that the propagation constant
may be the same for both the electric and magnetic components. That is
not so with these more complex five-dimensional waves. Thus, we should
allow for the possibility that each component may have a different
propagator. With this in mind we will try to find wave solutions with the
following trial forms.
230

E = E O exp[-i(wt - k e x - k 4e )]
B = B0 exp[-i(wt - k b x - k 4b )]
V = V O exp[-i(wt - k v x - k 4v )]
V 4 = V 4O exp[-i(wt - k 4 x - k 44 )] .

By using these four forms which constitute our trial solution we may
find that two, or all, of the k's must be the same but we aren't forcing them
to equality prematurely.
Now lets put our trial solutions into the field equations. Lets start
with the field equation
1
c

B
+ xE = 0 .
t

(3.15b)
The x and y components of this equation require that
Bx = B y = 0 .

For the z component we have, after simplifying,


c
w

Bz =

ke
Ey .
x

ke +

(6.84)
The second field equation is
xV + aO

=0 .

(3.15f)
The x and y components require that
V z=0 .

However, the z component gives us


- a0 c
V y=
w

k 4b + x

kb

ke +
k +

k 4e
x
Ey .
k 4V
x

(6.85)
Now look at the field equation
231

( E )= 4

- aO

( V4)

(3.15d)
We see that we face more assumptions. The first one concerns whether or
not varies with space. The classical assumption seems appropriate here
also. That is, if the medium is isotropic then
x

=0 .

The second assumption is that there are no free charges so that


=0 .

The third place for a possible assumption resides in the possible


dependence of upon mass density. However, experiment indicates that
does vary with changes in mass density. Therefore, it seems inappropriate
to assume differently.
Thus, Eqn. (3.14d) requires that
V4 =

( ke+

-1

aO - 1
i

k 4e )Ex
x
. (6.86)
+ k 4 + k 44 ]

From Eqn. (6.82) we see that the gravitational potential, V4, depends only
upon the longitudinal electric field component.
The fourth field equation is
1
c

V
E
+ V 4 = a0
.
t

(3.15g)
The x component of this vector equation requires
( k4+
-c
{
V x=
w
a O | k 44 x

k 44 )( +
k 4e )
ke
x
x
k4
k4 i
|

+ aO ( k 4e + x

ke

} Ex .

(6.88)
232

From the y component of Eqn. (3.15g) we find


V y=

- a0 c
k
( k 4e + x e ) Ey .
w

(6.88)
If we compare Eqn. (6.88) with Eqn. (6.85) we must have
kv+

k 4v
x

ke

k 4e + x

k 4e
x

= ke+

k 4b +

kb

(6.89)
The z component of Eqn. (3.15g) is an identity, thus we can
turn to the fifth field equation, which is
V+

V4 -4
=
t
c

J4 .

(3.15h)
The usual conductivity assumptions seem appropriate here and are taken
as
J = E , and J 4 =

V4

(6.90)
Therefore, Eqn. (3.15h) becomes the indicial relation
2

c kv+
k 44 + x

k 4v
x
k4

k4+
i
e

k 44
x
=(

k 4e
k
+ a02 k 4e + x e
x
x

ke+
2

+ i4

4) k e +

k 4e
.
x

(6.91)
when Eqns. (6.86) and (6.88) are used.
The sixth field equation is

xB -

E
4
=
t
c

J - aO

(3.15c)
233

The x component of this equation produces the indicial relation

+ i4

= c 2 k 4v + x

k 44
x

k4 +

kv

k 44 + x
ke

+ a O2 k 4e + x

ke+

k 4e
x

k4 - i

(6.92)
The z component of Eqn. (3.15c) is an identity but the y component
gives us another indicial relation in
2
w + i4

k 4e
)( k b +
x

w =( ke+

+ a O2 c 2 ( k 4 + x

)( k 4e + x

k 4b
)
x
ke

(6.93)
The seventh field equation
.B = 0 .

(3.15a)
is an identity since Bx = 0. However, the last remaining field equation
0=

J + aO

J4

(3.15e)
is not satisfied identically. Rather, it requires

234

(
Ex
= - aO
x

V4)

(6.94)
if we assume
x

=0 .

We are once more in a dilemma created by our ignorance. We don't


know what 4 is other than a "gravitational conductivity". Thus, it would
appear that we cannot assume that it is independent of the mass density.
Thus Eqn. (6.94) may be put into the form
(

) k 44 + x

k4

=i

(6.95)

This is an extremely curious equation. It relates the electrical


conductivity to the gravitational conductivity and includes in this relation
the variation of the dielectric constant with respect to changes in the mass
density.
In the absence of any knowledge about 4, the gravitational
conductivity, let us assume that there is a linear relationship between it
and the mass density. The only justification for this assumption comes
from the association of gravitational mass to gravitational conductivity and
also to inertial mass and hence mass density. Regardless of our lack of
knowledge lets assume
4

Further, in attempting further simplification lets assume


=

Using these assumptions we may rewrite Eqn. (6.95) as


k 44 + x

k4

i
l-

ln

(6.96)
235

The right hand side of Eqn. (6.96) is in terms of quantities which


should be available through experimentation, thereby giving us a
differential equation in the two unknowns k44 and k4.
We now have a system of equations given by the following numbered
equations: (6.84), (6.86), (6.87), (6.88), (6.89), (6.91), (6.92), (6.93), and
(6.96). These equations give the field components in terms of Ey and Ex.
However, the eight field equations cannot determine a relationship between
the longitudinal and transverse components. This is left for the energymomentum tensor to determine.
In order to attempt a reduction of these equations we might try the
simplifying assumptions:
k 4e
=0 ,
x
k 4b
=0 ,
x

and
k 4v
=0 .
x

Using these assumptions let us now look at the requirements that


come from the interchangeability of substitution and differentiation. For
instance,
k c
Bz
= ik b e E y ,
x

when differentiation is taken first.


substitution first then we get

On the other hand if we take the

k c
Bz
= i ke e E y .
x

By comparing these two we see that


kb = ke .

When we compare the two expressions for the partial derivative of Bz


with respect to the mass density we find
( lnk e )

= i( k 4b - k 4e ) .

Turning next to the transverse gravitational component, the partial


derivative with respect to x requires that

236

ke

=0 .

(6.97)
This sets up something of a dilemma since ke should depend upon
both and . These in turn depend upon . Therefore, Eqn. (6.97) is a
result that does not seem to correspond to experiment. Thus, our choice of
simplifying assumptions appears too restrictive. But which assumption is
the one that must be relaxed? An investigation into the necessity of the
assumptions made seems required prior to making advancement toward
the solutions of the electromagnetogravitic wave equations.

237

References
W.R. Adey, Tissue Interactions with Nonionizing Electromagnetic Fields,
Physiological Reviews, Vol. 61, No. 2, April 1981.
D. ter Haar and H. Wergelande, Elements of Thermodynamics, 1966.
H. Weyl, Space, Time, and Matter, Dover, 1922.
R. Adler, M. Bazin, and M. Schiffer, Introduction to General Relativity,
1965.
Th. Kaluza, Sitzungsber. d. Preuss. Akad. d. Wiss., 1921, p. 966.
O. Veblen, Projektive Relativitatstheorie, Berlin, springer, 1933.
W. Pauli, Ann D. Physik, 19, 305 (1933); 18, 337 (1933).
A. Einstein and Mayer, Berl. Ber., 1931, p. 541; Berl. Ber. 1932, p. 130.
A. Einstein and P. Bergmann, Ann. of. Math., 39, 683 (1938).
Einstein, V. Bargmann, and P.G. Bergmann, Theodore von Karmaan
Anniversary Volume, Pasadena, 1941, p. 212.

238

Chapter 7 Hydrodynamic Systems


The equation of motion for the fifth dimension, mass density,
appears as a generalization of the principle of the conservation of mass.
Further, in classical hydrodynamic systems five equations in five
unknowns are used. It seems logical then to expect the five equations of
motion appearing in the five-dimensional Dynamic Theory to be
generalizations of the classical equations.
An added incentive to
investigate the possibilities of this generalization is gained when
electromagnetically contained ionized plasmas with mass conversion are
considered. For if the five equations are generalizations of the classical
hydrodynamic equations, then the use of the five-dimensional fields
allowing mass conversion should provide an entirely new viewpoint of a
controlled fusion reactor.
Since it is suspected that the five equations of motion resulting from
the application of the principle of increasing entropy to a
thermo-mechanical system are generalizations of the classical equations, it
then becomes necessary to show that this is indeed the case. This seems
possible by restricting the system so that it corresponds to the usual
system considered.
First, from the Dynamic Theory approach, the manifold required for
a description of the system is the five-dimensional manifold of space, time,
and mass density. Within this manifold the continuity equation of mass
no longer holds for the general system. We can, however, restrict our
system by first requiring that the system remain on a hypersurface within
the five-dimensional manifold. For a system so restricted, any of the five
dimensions may be considered as functions of the other four.
In
particular, since by custom in hydrodynamics the mass density is
considered to be a function of space and time, we may consider the mass
density to be the variable chosen to be function of the others or
=

0
1
2
3
x ,x ,x ,x

so that
d =

dx

Such a system will be constrained to be on a hypersurface


embedded within the five-dimensional manifold of space, time, and mass
density as shown and upon this hypersurface will be described in a
four-dimensional manifold of space and time.
If we further restrict our system by requiring that the total derivative
of the mass density to be zero or

239

d =0=

d
dx

dx

then
d
=0=
+ 1 v1 + 2 v 2 + 3 v 3
dt
t
x
x
dx

or
t

+ grad

v=0 ,

which is the usual continuity equation. Thus, by restricting the system to


this particular hypersurface we have constrained the system to obey the
continuity equation as does a usual hydrodynamic system.
Not only does this restriction place our system within the space-time
manifold where we may compare the resulting four equations of motion
with the equations of motion in relativistic theories but, since the seven
gauge field equations must hold in the five-dimensional manifold they
must also hold on the hypersurface. This allows the new field quantities to
be expressed as functions of the , B fields and the partial derivatives of
the mass densities. Further, it appears that the additional B field
equations may be used to determine a dependence of the E and B fields
upon the mass density and/or its changes.
Then by comparing the equations of motion obtained here for the
system restricted to the mass conservation hypersurface with the
relativistic Navier-Stokes equations it should be possible to identify the
viscous coefficients with the field quantities and perhaps see how the
viscosity depends upon these fields as I feel it does.
Since we have restricted the system to a hypersurface where the
mass density is a function of space and time, then the surface is defined by
five equations of the type
i
i
0
1
2
3
x = x ( u ,u ,u ,u ) .

(7.1)
Further, since x4 = /a0 and x4 = x4(x0, x1, x2, x3), then Eqn. (7.1) becomes
0
0
1
1
2
2
3
3
x = u ,x = u ,x = u ,x = u

and
4
0
1
2
3
x = f ( u ,u ,u ,u ) .

240

Since u0, u1, u2, and u3 are independent variables, the locus defined
by Eqn. (7.1) is four-dimensional, and these equations give the coordinates
xi of a point on the hypersurface when u0, u1, u2, and u3 are assigned
particular values. This point of view leads one to consider the surface as a
four-dimensional manifold S embedded in a five-dimensional enveloping
space. We can also study surfaces without reference to the surrounding
space, and consider parameters u0, u1, u2, and u3 as coordinates of points
in the surface.
If we assign to u0 in Eqn. (7.1) some fixed value u0 = u0, we obtain a
three-dimensional manifold
i
i
0
1
2
3
x = x ( u , u , u , u ), (i = 0, 1, 2, 3, 4)

which is a three-dimensional manifold lying on the hypersurface S defined


by Eqn. (7.1). By assigning fixed values for any three of the four
hypersurface variables we obtain a net of curves, on the hypersurface,
which may be called coordinate curves.
Obviously the parametric representation of a hypersurface in the
form of Eqn. (7.1) is not unique, and there are infinitely many curvilinear
coordinate systems which can be used to locate points on a given
hypersurface S. Thus, if one introduces a transformation
0
0
-0
-1
-2
-3
u = u ( u ,u ,u ,u ) ,
1
1
-0
-1
-2
-3
u = u ( u ,u ,u ,u ) ,
2
2
-0
-1
-2
-3
u = u ( u ,u ,u ,u ) ,

and
3
-0
-1
-2
-3
u3 = u ( u , u , u , u ) ,

(7.2)
where the u (u-0, u-1, u-2, u-3) are of class C1 and are such that the
Jacobian
J=

( u O , u1 , u 2 , u 3 )
( u -0 , u -1 , u - 2 , u - 3 )

does not vanish in some region of the variables u, then one can insert the
values from Eqn. (7.2) in Eqn. (7.1) and obtain a different set of parametric
equations
i
i
-0
-1
-2
-3
x = f ( u ,u ,u ,u )

(7.3)
defining the hypersurface S. Equation (7.2) can be looked upon as
representing a transformation of coordinates in the hypersurface.
241

7.1 First Fundamental Quadratic Form


The properties of hypersurfaces that can be described without
reference to the space in which the hypersurface is embedded are termed
"intrinsic" properties. A study of intrinsic properties is made to depend on
a certain quadratic differential form describing the metric character of the
hypersurface. We proceed to derive this quadratic form for our restricted
system.
It will be convenient to adopt certain conventions concerning the
meaning of indices to be used. We will be dealing with two distinct sets of
variables: those referring to the five-dimensional space in which the
hypersurface is embedded (these are five in number) and with four
coordinates u0, u1, u2, and u3 referring to the four-dimensional manifold S.
In order not to confuse these sets of variables we shall use Latin letters for
the indices referring to the space variables and Greek letters for the
hypersurface variables. Thus, Latin indices will assume values 0, 1, 2, 3, 4
and Greek indices will have the range of values 0, 1, 2, 3.
A
transformation T of space coordinates from one system X to another X will
be written as

a transformation of Gaussian hypersurface coordinates, such as described


by Eqn. (7.2) will be denoted by

A repeated Greek index in any term denotes the summation from 0 to 3; a


repeated Latin index represents the sum from 0 to 4. Unless a statement
to the contrary is made, we shall suppose that all functions appearing in
the discussion are of class C2 in the regions of their definitions.
Consider the hypersurface S defined by
i
i
0
1
2
3
x = x ( u ,u ,u ,u ) ,

(7.4)
where the xi are coordinates covering the five-dimensional space in which
the hypersurface S is embedded, and a curve C on S defined by
u =u ( ) ,

(7.5)
where the u 's are the Gaussian coordinates covering S. Viewed from the
surrounding space, the curve defined by Eqn. (7.4) is a curve in a
five-dimensional manifold, which we shall assume, for the present, is
242

Riemannian entropy manifold of the Dynamic Theory, and its element of


arc is given by the formula
( dq0 )2 = g ij dxi dx j

(7.6)
From Eqn. (7.4) we have
i

x
du
u

i
dx =

(7.7)
where, as is clear from (7.5),

du =

du
d
d

Substituting from Eqn. (7.6) and Eqn. (7.7), we get


dq

0 2

x x
du du
u u
= A du du ,

= g ij

where
x
u

g ij

x
u

(7.8)
The expression for (dq0)2, namely
dq

0 2

= A du du

is the square of the linear element of C lying on the hypersurface S, and


the right hand member of (7.8) can be called the First Fundamental
quadratic form of the hypersurface. The length of arc of the curve is given
by
2

_A u b d

q 2 - q1 =

where
u =

du
25 and q0 is the specific entropy.
d

along the curve C would then be


243

The total change in the entropy

0
0
q 2 - q1 =

_A u u d

(7.10)
Consider a transformation of surface coordinates
0

u = u ( u ,u ,u ,u )

(7.11)
with a non-vanishing Jacobian
u

J=

It follows from Eqn. (7.11) that


u

du =

du

and hence Eqn. (7.9) yields


dq

0 2

=A

du du

If we set
A =A

we see that the set of quantities A represents a symmetric covariant


tensor of rank two with respect to the admissible transformations Eqn.
(7.11) of hypersurface coordinates. The fact that the A are components of
a tensor is also evident from Eqn. (7.9), since (dq0)2 is an invariant and the
quantities A are symmetric. The tensor A is called the covariant metric
tensor of the hypersurface.
Since the form Eqn. (7.9) is positive definite, the determinant
A= A

>0

and we can define the reciprocal tensor A by the formula A A =


The properties of surfaces concerning the study of the first
fundamental quadratic form
dq

0 2

= A du du

244

constitute a body of what is known as the 'intrinsic geometry of surfaces.'


They take no account of the distinguishing characteristics of surfaces as
they might appear to observer located in the surrounding space. Two
surfaces, a cylinder and a cone, for example, appear to be entirely different
when viewed from the enveloping space, and yet their intrinsic geometries
are completely indistinguishable since the metric properties of cylinders
and cones can be described by the identical expressions for square of the
element of arc. If a coordinate system exists on each of the two surfaces
such that the linear elements on them are characterized by the same
metric coefficients A , the surfaces are called "isometric."
Thus, if our description of the restricted system is done only in terms
of the intrinsic geometry of the hypersurface we may lose sight of features
which may characterize our system when viewed from the enveloping
space. Therefore, in order to characterize the shape of the surface we must
develop a view which involves the enveloping space.
7.2 Second Fundamental Quadratic Form
An entity that provides a characteristic of the shape of the surface as
it appears from the enveloping space is the normal line to the surface. The
behavior of the normal line as its foot is displaced along the surface
depends on the shape of the surface, and it occurred to Gauss to describe
certain properties of surfaces with the aid of a quadratic form that depends
in a fundamental way on the behavior of the normal line. Before we introduce this new quadratic form let us recall the definition Eqn. (7.8),
A

g ij

x
u

x
(i, j = 0, 1, 2, 3, 4) ( , = 0, 1, 2, 3) .
u

We note that the foregoing formulas depend on both the Latin and
Greek indices, and we recall that the Latin indices run from 0 to 4 and
refer to the surrounding space, whereas the Greek indices assume values
0, 1, 2, and 3 and are associated with the embedded hypersurface.
Furthermore, the dxi and gij's are tensors with respect to the
transformations induced on the space variables xi, whereas such
quantities du and A are tensors with respect to the transformation of
Gaussian surface coordinates u . Equation (7.8) is a curious one since it
contains partial derivatives
i

x
35 depending on both Latin and Greek indices. Since both A
u

and gij in Eqn (7.8)

are tensors, this formula suggests that


i

x
36 can be regarded either as a contravariant space vector or as a covariant surface
u

vector. Let us investigate this set of quantities more closely.

245

Let us take a small displacement on the hypersurface S, specified by


the surface vector du . The same displacement, as is clear from Eqn. (7.7),
is described by the space vector with components
i
dx =

x
du
u

(7.12)

The left-hand member of this expression is independent of the Greek


indices, and hence it is invariant relative to a change of the surface
coordinates u . Since du is an arbitrary surface vector, we conclude that
x
u

(7.13)
is a covariant surface vector. On the other hand, if we change the space
coordinates, the du , being a surface vector, is invariant relative to this
change, so the Eqn. (7.13) must be a contravariant space vector. Hence we
can write Eqn. (7.13) as
x

x
u

(7.14)
where the indices properly describe the tensor character of this set of
quantities.
Let A and B be a pair of surface vectors drawn from one point P of S.
FIG HERE
Then using Eqn. (7.14) they can be represented in the form
i
i
i
i
A = x A and B = x B

(7.15)
The five-dimensional vector product, defined by
k
N =

kij

Ai B j ,

(7.16)
246

is the vector normal to the tangent plane determined by the vectors A and
B, and the unit vector n perpendicular to the tangent plane, so oriented
that A, B, and n form a right-handed system, is
kij

n=

Ai B j
A B

(7.17)
We call the vector n the unit normal vector to the hypersurface S at P.
Clearly, n is a function of coordinates (u0, u1, u2, u3), and as the point P(u0,
u1, u2, u3) is displaced to a new position P(u0 + du0, u1 + du1, u2 + du2, u3 +
du3), the vector n undergoes a change
dn=

n
u

du

(7.18)
whereas the position vector r is changed by the amount
dr=

du

Let us form the scalar product


dn dr=

du du

(7.19)
If we define
b =

1
2

so that Eqn. (7.19) reads


d n d r = - b du du

(6.20)
the left-hand member of Eqn. (7.20), being the scalar product of two
vectors in a Riemannian space by being in the entropy manifold, is an
invariant; moreover, from symmetry with respect to
and , it is clear
that the coefficients du du in the right-hand member of Eqn. (7.20) define
a covariant tensor of rank two. The quadratic form
247

B b du du

(7.21)
called the second fundamental quadratic form of the hypersurface, will be
shown to play an essential part in the study of hypersurfaces when they
are viewed from the surrounding space, just as the first fundamental
quadratic form
A d r d r 49 or
A = A du du

did in the study of intrinsic properties of a hypersurface.


We can rewrite the formula Eqn. (7.17) in terms of the components
x i of the base vectors a . We denote the covariant components of n by ni
and observe that its covariant components ni are given by
ni =

A B
AB sin
ijk

(7.22)
and
A B sin =

A A

(7.23)
Substituting in Eqn. (7.22) from Eqn. (7.15) and Eqn. (7.23), we get
-

ni

ijk

x x

A B =0

and, since this relation is valid for all surface vectors, we conclude that
=

ni

ijk

x x

(7.24)
Multiplying Eqn. (7.24) by
result

, and noting that

ni =

1
2

ijk

x x

=2, we get the desired

(7.25)

248

It is clear from the structure of this formula that ni is a space vector


which does not depend on the choice of surface coordinates. This fact is
also obvious from purely geometric considerations.
7.3. Tensor Derivatives
We wish to reduce the second fundamental quadratic form eqn.
(7.21) analytically by the operation of tensor differentiation of tensor fields
which are functions of both surface and space coordinates. To do this we
shall first present the concept of tensor differentiation introduced by A. J.
McConnell*.
Let us consider a curve C lying on a given hypersurface S and a
vector Ai defined along C. If is a parameter along C, we can compute
the intrinsic derivative
i

56 of Ai, namely,
i

i
k
dA i
j dx
+g
,
A
jk
dt
d

(7.26)
In formula eqn. (7.26) the Christoffel symbols
g

i
58 refer to the space coordinates xi and are formed from the metric
jk

coefficients gij. This is indicated by the prefix


g 59 on the symbol. On the other hand, if we consider a surface vector A defined along the
same curve C, we can form the intrinsic derivative with respect to the surface variables,
namely,
S

dA
+a
d

du
d

(7.27)
In this expression the Christoffel symbols
61 are formed from the metric coefficients a

associated with the Gaussian

hypersurface coordinates u . A geometric interpretation of these formulas


is at hand when the fields Ai and A are such that
i

A =0
62 and
A =0
63. In the first equation the vectors Ai form a parallel field with respect to

C, considered as a space curve, whereas the equation


249

= 0 64 defines a parallel field with respect to C regarded as a surface curve.

The

corresponding formulas for the intrinsic derivatives of the covariant vectors Ai and A are
Ai

j
k
d Ai
dx
- g
Ak
ij
d
d

(7.28)
and
A

dA
-a
d

du
d

(7.29)
Consider next a tensor field
i
T 67, which is a contravariant vector with respect to a transformation of space coordinate
xi and a covariant vector relative to a transformation of surface coordinates
u . An example of a field of this type is the tensor
I
x =

x
68 introduced earlier. If
n

i
T 69 is defined over a surface curve C, and the parameter along C is , then

. We introduce a parallel vector field Ai along C,


regarded as a space curve, and a parallel vector field B along C, viewed as
a surface curve, and form an invariant
i
T 70 is a function of

( ) = T i Ai B

The derivative of
expression

( )

with respect to the parameter

i
i
d
dT
dB
i dA
i
=
Ai B + T
B + T Ai
d
d
d
d

is given by the

(7.30)
which is obviously an invariant relative to both the space and surface
coordinates. But, since the fields Ai( ) and B ( ) are parallel,
j
k
d Ai
dx
dB
= g
and
=a
Ak
ij
d
d
d

and eqn. (7.30) becomes


250

du
d

i
k
i
d
dT
j dx
=
+ g
-a
T
jk
d
d
d

du
d

Ai B

(7.31)
Since this is invariant for an arbitrary choice of parallel fields Ai and B ,
the quotient law guarantees that the expression in the brackets of Eqn.
(7.31) is a tensor of the same character as
i
T 75. We call this tensor the intrinsic tensor derivative of
i
T 76 with respect to the parameter , and write
k
i
i
dTi
T
j dx
=
+ g
=a
T
jk
t
d
d

du
d

If the field
i
T 78 is defined over the entire hypersurface S, we can argue that, since
T

i
i
T
j k
+g
T x -a
jk
u

du
d

is a tensor field and


du
80 is an arbitrary surface vector (for C is arbitrary), the expression in the bracket is a
d

tensor of the type


i
T 81. We write
T

i
i
T
j k
+ g
T x -a
jk
u

(7.32)
and call
i
T , 83 the tensor derivative of
i
T 84 with respect to u .
The extension of this definition to more complicated tensors is
obvious from the structure of Eqn. (7.32). Thus the tensor derivative of
i
T 85 with respect to u is given by
T

i
i
T
k
i
+g
T x -a
jk
u

-a

(7.33)
If the surface coordinates at any point P or S are geodesic, and the
space coordinates are orthogonal Cartesian, we see that at that point the
tensor derivatives reduce to the ordinary derivatives. This leads us to
251

conclude that the operations of tensor differentiations of products and


sums follow the usual rules and that the tensor derivatives of gij, A , Gijk,
and their associated tensors vanish. Accordingly, they behave as
constants in the tensor differentiation.
The apparatus developed in the preceding section permits us to
obtain easily and in the most general form an important set of formulas
due to Gauss. We will also deduce with its aid the second fundamental
quadratic form of a surface already encountered.
We begin by calculating the tensor derivative of the tensor
i
87,
representing
the components of the surface base vectors a . We have
x
i
x =

+ g {

i
} x j x k - a{
jk

} xi ,

from which we deduce that


x

= xi ,

(7.34)
Since the tensor derivative of a
the relation

vanishes, we obtain, upon differentiating


i
j
A = g ij x x ,

g ij xi , x j + g ij xi x j , = 0 .

(7.35)
Interchanging , ,

cyclically leads to two formulas:


g ij xi

j
i
j
x + g ij x x , = 0

(7.36)
and
g ij xi , x j + g ij xi x J , = 0 .

(7.37)
If we add Eqn. (7.36) and Eqn. (7.37), subtract Eqn. (7.35), and take into
account the symmetry relation Eqn. (7.34), we obtain
g ij xi , x j = 0 .

This is the orthogonality relation which states that


i
x , 94 is a space vector normal to the surface, and hence it is directed along the unit normal
ni. Consequently, there exists a set of functions b such that
252

= b ni .

The quantities b are the components of a symmetric surface tensor, and


the differential quadratic form
B b du du

is the desired second fundamental form.


Now since
i
i
=
njn
j 97, and
j
ni = g ij n 98, then
i
j
b = g ij x , n ;

but since
ni =

1
2

ijk

j k
x x 100, then

b =

1
2

ijk

x x

(7.38)
We now have, in Eqn.s (7.8) and (7.38), the formulas necessary to
determine the first and second fundamental quadratic forms for our
system constrained to a four-dimensional hypersurface. Our objective is to
show that by appropriately constraining our system we arrive at the
Navier-Stokes equations. Let us determine the first fundamental quadratic
form.
First recall that our system was restricted so that x4 = x4(xo, x1, x2,
x3) or the mass density is a function of space and time; then we have the
relations
0
0
x =u ,
1
1
x =u ,
2
2
x =u ,
3
3
x =u ,

and
4
0
1
2
3
0
1
2
3
x = f ( x , x , x , x )= f ( u , u , u , u ) .

Since Eqn. (7.8) is


A = g ij

x
u

253

x
= g ij xi x j ,
u

then
2

A00 = g 00 = 2 g 04 f 0 + g 44 ( f 0 )

where
f

f0

In a similar fashion we may determine the remaining coefficients and find


that
A = g + h

(7.39)
where
h = 2 g

f + g 44 f f

, = 0, 1, 2, 3 ,

(7.40)
where the h are functions of the partial derivatives of the mass density
with respect to space and time in addition to space and time from the
g i4 109 where i = 0, 1, 2, 3, 4.
Though we may use Eqn. (7.38) to determine the metric coefficients
for the second fundamental quadratic form, it is not necessary for the
current presentation.
The hypersurface which is embedded in the five-dimensional space
is a four-dimensional curvilinear space-time manifold. Thus the relativistic
hydrodynamic equations are applicable here so long as the metric
coefficients are determined as coefficients of the hypersurface quadratic
form.
The complete energy-momentum tensor for a fluid in a flat
Riemannian space-time manifold is given by
T

= u u +

P
c

(u u - g

(7.41)
where
du
111, s is the arc length. Then based upon this energy momentum tensor the flow
u
ds
of a fluid under the effect of its own internal pressure force is given by setting the
divergence of Eqn. (7.41) equal to zero, or
254

, =0 .

(7.42)
If we reduce Eqn. (7.41) to the non-relativistic limit, the use of Eqn. (7.42)
gives us
g

, , = 1, 2, 3 ,

where

= Pg is the three-dimensional stress tensor of an ideal fluid.


If in Eqn. (7.41) we use the fact that the metric coefficients for the
hypersurface may be written as the sum of Eqn. (7.40), then we have
T

= u u +

P
c

u U - g

-h

(7.43)
where it must be remembered that the
u u 115 are also dependent upon this same sum. In the non-relativistic limit the effects of
this sum of metric tensors appear as a sum in the stress tensor
= - P g

- Ph

, = 1, 2, 3 .

(7.44)
Recall that the
g 117 refer to the three-dimensional space viewed from the five-dimensional manifold.
The h , however, contain the information about the surface embedded in
the five-dimensional space. If we then associate the tensor
- Ph

(7.45)
with the viscous stresses, we are saying that the viscous stresses depend
upon the geometric character of the hypersurface.
In the limit of small displacements we write the strain velocity tensor
as
e =

1
v
2

+v

Then the first order coefficients of viscosity are related to the strain velocity
tensor and viscous stresses according to

255

t =

,n

+ v n,

(7.46)

If we then use Eqn. (7.45) in Eqn. (7.46), we find that the relationship between the geometric character of the hypersurface and the viscous
coefficients is given by
- Ph =

,n

+ v n,

(7.47)
Equation (7.47) then expresses the functional dependence of the
viscous coefficients upon the strain velocities, pressure, mass density, and
their derivatives.
7.4. Relativistic Hydrodynamics.
By viewing classical hydrodynamics to be given by the embedding of
a four-dimensional hypersurface within a five-dimensional manifold, the
association Eqn. (7.47) between the geometric properties of the
hypersurface and the viscous coefficients could be tentatively made. We
may now go back and develop this relationship more completely.
The hypersurface, which becomes embedded in the five-dimensional
manifold by the restriction that x4 = x4(x0, x1, x2, x3) is a four-dimensional
relativistic manifold. Thus, for the surface we may use the relativistic
energy-momentum tensor, which is
T

= u u+

P
c

u u -g

(7.48)
where
u

dx
123 and v = 0, 1, 2, 3. The divergence of Eqn. (7.48) yields the flow
0
dx

equations for a fluid under the effects of its own internal pressure.
However, from the viewpoint of the Dynamic Theory, the surface
metric coefficients may be written in terms of the metric coefficients of the
first four space coordinates as given by Eqn.s (7.40) and (7.41), or
A = g

256

, = 0, 1, 2, 3 .

Thus, the square of the arc length for the entropy manifold may be written
as
dq

0 2

= A dx dx = g dx dx + h dx dx

or, if
dx
126, then
0
dq

1 = A u u = g u u + h u U

Then on the hypersurface the energy momentum tensor would become


T

= u u +

P
c

u u -A

or
T

= u u +

P
c

u u - g

-h

(7.49)
Since the surface coordinates, x , are the same as the first four coordinates
of the surrounding space, the velocities u are the same whether
considered as surface or space vectors. The difference between the surface
view and a four-dimensional space view appears in the metric coefficients.
Thus, while the square of the arc element on the surface is unity, the
square of the arc element in the surrounding space is not, or
1= A u u

but
g u u = 1 - h u u

7.5. Classical Hydrodynamics.

Suppose we consider the metric given by


132 to be a flat space then because of Eqn. (7.49) we may write

= u u +

P
c

If we then form the space divergence


257

u u - g

P
c

0
T , =

1
c t

P
c

00

P u
u
+ 2
- H0
c c c

=0 ,

this may be written as


1
c

1 ( Ph00 ) 1
1
+ 2 (P v ) - 2
2
t
c
c
c

( v) -

( Ph-0 ) = 0 ,

where h0 has components h0 , = 1, 2, 3. Therefore,


t

( v )=

1
c

(P v ) +

( Ph00
+
t

1
c

so that if hv is a four-vector with components h0


then
t

( v )= -

( Ph0 )

h 137,

1
P
1
P v - v , + 2 ( Ph ),
c
c
c

The remaining components of the divergence are given by


T
+

, =

1
c t

u
Pv P
+ 3 - 2h
c
c
c

P v v
v v
+ 2
+
2
2
c
c
c

Ph
=0 ,
2
c

which may be rearranged to read


v
+v
t

v =-

x
-

-v

1
c

( v) +

( Pv )
+
t

(P v v )

1 ( Ph 0 )
( Ph )
+
c
t
x
.

If we look at the non-relativistic limit, then, by neglecting the terms P(v/c),


we get
t

( v )=

1 1 ( Ph00 )
+
2
t
c c

( Ph-0 )

The multiplicative factor 1/c2 on the right-hand side suggests that


t

( v) 0 ,

which is the assumption we chose to place our system on a particular


surface. This corresponds to a classical system where conservation of
mass is assumed. Therefore, on the surface of a curve specified by
258

, =0 ,

we must then have


v
+v
t

v =-

1 ( Ph ) )
( Ph )
+
c
t
x

or
a =-

1 ( Ph 0 )
( Ph )
+
c
t
x

x
1 ( Ph 0 )
=+
+ ( Ph ),
c
t
x
P

= 1, 2, 3 .

Thus, we may write


a =

where
= - P g + Ph = - P( g

-h ) .

(7.50)
The term
1 ( Ph 0 )
148 has been neglected in Eqn. (7.50).
c
t

Thus we see that the geometric character of the hypersurface,


contained in the term Ph , behaves as if it were a viscous effect to be
added to the normal viscous effects. Recalling Eqn. (7.41), it may be seen
that the viscous-like effects of the geometry of the hypersurface depend
upon the density gradient. If these terms exist, they must be very small in
everyday phenomena. Yet if we consider phenomenon which involve very
large density gradients, these terms could become large enough to see.
7.6. Shock Waves.
One field of physical phenomena that displays large density
gradients is shock waves. Therefore let us take a quick look at the effect of
these additional terms on the description of a shock front for a steady,
one-dimensional shock.
The total stress in a steady, one-dimensional shock would be given
by
= P 1-

1
a

2
O

259

du
dx

when g11 = 1 and h11 is evaluated using Eqn. (7.41). However, for a steady
shock we also have the jump conditions
u = k1 ,
u
k1 + = k 2 ,

and
2

k1 E -

= k3 .

These equations represent the conservation of mass, momentum, and


energy. By using the conservation of mass relation we may write the total
stress as
= P+

eff

du
dx

where
eff

P k 12 du
2 4
a0 u dx

(7.51)
may be called the effective viscous coefficient. Since within the shock front
the velocity gradient du/dx is negative, we see that the effective viscous
coefficient acts so as to thicken the shock front when compared to the
classical viscous coefficient .
Using the second jump condition, an expression for the velocity
gradient is
du a02 u 4
4 Pk 12
[P - k 2 + k 1 u] , (7.52)
1
1
+
=
2 4 2
dx 2 Pk 12
a0 u

which may be approximated by


du
1
P k2
= [P - k 2 + k 1 u] 1 - 2 4 1 2 [P - k 2 + k 1 u] .
dx
a0 u

(7.53)
The effect of the correction term on the velocity gradient is seen in
Eqn. (7.53), because the multiplicative factor outside the brackets is the
classical expression for the negative velocity gradient. The effect of the
260

correction term lessens the negative velocity gradient and extends the
shock front.
The effect of the correction term in Eqn. (7.51) is estimated by
considering the strong shock dependence of pressure upon shock
velocities. For instance, the shock pressure, from the jump conditions, is
P=

U up .

(7.54)
If the shock velocity is related linearly to the particle velocity as the
assumed solid equation of state, U = co + sup, then Eqn. (7.54) becomes
P=

U
(U - c o ) .
s

Thus, for strong shocks, P varies approximately as the square of the shock
velocity.
Consider Eqn. (7.52) or (7.53). From either of these equations, the
velocity gradient varies as the square of the shock velcoity. Using these
two conclusions in Eqn. (7.51) for the total viscosity eff and remembering
that the integration contant k1 is given by - oU, the effective viscosity
varies approximately as the square of the shock velocity or, essently , as
the pressure.
The conclusion is that if the effective viscosity varies with the
pressure, an increase by the same factor of 103 must be accompanied by a
viscosity increase by the same factor of 103. This explains the apparent
discrepancy between the low and high pressure aluminum viscous effects.
For instance, the Asay-Bertholf limits are:
P = 25 GPa
P = 36 GPa

> 40 poise
< 2,500 poise .

Another experiment places an upper limit of 103 poise for a shock pressure
of 40 GPa. If 102 poise is considered representative of the viscosity when P
10 GPa, then from Eqn. (7.51), a pressure of 103-104 GPa must be
accompanied by a viscous effect of 104-105 poise.
This total viscosity estimate is supported by numerical integration
across the shock front using the Tillotson equation of state for aluminum.
The classically predicted rise times for shocks of 40 GPa with = 575 p
and 5x103 GPa for
= 5x104p are duplicated by using the effective
viscosity experssion in Eqn. (7.51) with = 1.0 p and a0 365 g/cm4.
Thus, the Dynamic Theory correlates these data points that appear
contradictory by classical theory. Further, these data points provide an
estimate of the new universal constant appearing in the Dynamic Theory.
261

This value of a0 provides an estimate of other predictions of the theory in


fields other than shock waves.
7.7. Mass Conservative Electrodynamics
One of the incentives for seeking to determine whether the five
equations of motion were generalizations of the classical hydrodynamic
equations was the possibility of shedding new light upon fusion plasmas.
Now before mass conversion is accomplished the plasma must reach
certain conditions.
The attainment of these conditions involve
electromagnetic fields not encountered in usual circumstances on earth. If
the Dynamic Theory is to be believed, then perhaps it may provide new
insight into the attainment of the appropriate conditions before mass
conversion begins.
The following development still assumes conservation of mass in
order to see the geometry of the hypersurface for a system under the
influence of electromagnetic fields.
Suppose we now describe the behavior of charged matter under the
influence of an electromagnetic field from the viewpoint of the Dynamic
Theory. From this viewpoint the conservation of mass has the effect of
restricting our system to a four-dimensional hypersurface which is
embedded in the five-dimensional manifold of space, time, and mass
density.
Since we desire to consider the effects of an electromagnetic field we
must consider a gauge function. When a gauge function exists, the square
of the arc length in the entropy space is related to the square of the arc
length in the sigma space by
dq

0 2

= g ij dxi dx j =

h00

g ij dxi dx j =

1
h00

When the system is restricted to a hypersurface by the relation x4 = x4(x0,


x1, x2, x3), then the entropy surface may be written as
dq

0 2

= a du du

where

a = g ij

x
u

xi
= g ij xi x j .
u

Likewise for the sigma surface


(d

2
) = a du du

262

where
i
j
a = g ij x x .

Thus, we have
a =

1
h00

The principle of increasing entropy requires that the equations of


motion be geodesics in the entropy space but they will appear as equations
involving forces in the sigma space. We desire to expose these forces and
therefore should work in the sigma space. Our objective then is to
determine the effect of embedding a four-dimensional surface given by x4 =
x4(x0, x1, x2, x3) in the sigma space and thus obtain a sigma surface
describing a system subjected to the classical conservation of mass restriction.
Having previously determined the metric coefficients for the entropy
space by Eqns. (7.39) and (7.40) we may write the coefficients for the sigma
surface as
a = h00 a = h00 g + h

However by considering the effects of the electromagnetic field as a


force we must first consider the space field tensor:
0 E1 E 2 E3 V 0
- E 1 0 B3 - B 2 V 1
F ij = - E 2 - B 3 0 B1 V 2
- E 3 B 2 - B1 0 V 3
-V 0 -V 1 -V 2 -V 3 0

If we restrict ourselves to the classical field quantities E and B and


for the moment assume that the field quantities V4 and V are zero or
negligible, then we obtain only the effects of the hypersurface viewpoint.
This assumption seems reasonable considering the interpretation of the
new field quantities are gravitational effects. Under this assumption our
field tensor becomes
0 E1 E 2 E3 0
- E 1 0 B3 - B 2 0
F ij = - E 2 - B 3 0 B1 0
- E 3 B 2 - B1 0 0
00000

263

We can now use this space field tensor to determine the appearance
of the fields when viewed from the surface. The surface field tensor will be
given by
i
j
F = F ij x x .

But since x i = i for i, = 0, 1, 2, 3 and x 4 = f , the surface field tensor of


a purely electromagnetic space field tensor is only the four-dimensional
portion of the space field tensor since Fi4 = 0 for i = 0, 1, 2, 3, 4.
Thus when we use the relativistic energy-momentum tensor for the
surface, we have
T

= u u +

1
c

1
F + a F
4

(7.55)
which is the relativistic energy-momentum tensor for matter under the
influence of electromagnetic fields. But since
a = g + h 169, then Eqn. (7.55) becomes
T

= u u +

1
c

F +

1
g + h
4

or
= T rel + T geo

(7.56)
where
T rel

u u +

1
c

F +

1
g
4

is the four-dimensional space relativistic energy momentum tensor and


T geo

1
h
4 c2

is the portion of the energy-momentum tensor which contains the


geometrical properties of the hypersurface.
From Eqn. (7.56) we can say that the Dynamic Theory has the
appearance of adding a term to the relativistic energy-momentum tensor.
This term contains the geometrical character of the surface and represents
the difference between the appearance of the energy-momentum tensor
when viewed from the surrounding space as compared to the view from the
hypersurface.
264

If we take the divergence of the energy-momentum tensor Eqn.


(7.56), we have
T

, = T rel , + T geo , .

The additional force terms from the surface geometry are given by
1
4 c2

, =F

But if we define
- 16

(7.57)
as the electromagnetic energy density, where
=

1
2
2
E +B
8

then the geometric energy-momentum tensor becomes


-4 h
2
c

T geo =

and the additional forces are given by


F =-

4
c

We may also look at the radiation pressure predicted by the


Dynamic Theory to see how the surface restriction affects the relativistic
prediction of radiation pressure.
The relativistic radiation pressure is taken as one third of the
three-dimensional Maxwell stress tensor which is the space portion of the
energy-momentum tensor, or
T

1
E E +B B 4

where , = 1, 2, 3.
To get the equivalent stress tensor for the Dynamic radiation
pressure we must add the space portion of Eqn. (7.57) so that the total
stress tensor becomes
=

T
=

1
E E +B B 4

1
E E +B B 4

265

-4h
+h

We can then obtain the negative of the trace by


- {T} = -

1
2
2
E + B - 3 - h11 + h 22 + h33
4
= - - - h11 + h 22 + h33
.

The radiation pressure is then given by


P=

[1 + h11 + h 22 + h33 ] .

(7.58)
The first term in Eqn. (7.58) is the classical radiation pressure in
electrodynamics. The remaining three terms give the difference between
the pressure predicted by the Dynamic Theory and the classical prediction.
To determine what this difference is let us restrict our system to again be
very near equilibrium so that the g 4 = 0 for
= 0, 1, 2, 3 and g44 =
constant. Thus we have a flat space. For this space the
2

g 44

h =

a0

from Eqn. (7.58) and g44 = -1. Thus


h11 + h 22 + h33 = -

1
2

a0

(7.59)
By substituting Eqn. (7.59) into Eqn. (7.58) the pressure becomes
P=

1-

1
2

a0

However, since the classical pressure is given by


Pc =

187, then the pressure predicted by the Dynamic Theory becomes

P D = Pc 1 -

1
2

a0

We see then that the Dynamic Theory predicts a decrease in the


radiation pressure as a result of viewing the system to be restricted to a
four-dimensional hypersurface embedded in a five-dimensional space. The
amount of this decrease in pressure depends upon the gradient of the
266

mass density and the constant a0. Once the constant a0 is determined,
then the deviation in predicted pressures can be specified.
This prediction should appear in attempts to use electrodynamic
forces to control ionized plasmas and perhaps there are large enough
density gradients for these predictions to show up in cosmological events.
References:
*A.

J. McConnell, Absolute Differential Calculus, London, 1931, Chapters


IV - XVI.

267

Chapter 8 Experimental Tests


Every new theory should possess some feature that can be checked experimentally, for
the objective in the creation of a new physical theory should be a better understanding of
physical phenomena. The following suggested experiments represent but a few possible
tests of the five-dimensionality of the Dynamic Theory, for each depends upon the fivedimensional fields.
8.1 Speed-of-Light Measurements
The various speeds associated with the five-dimensional plane wave have been studied in
detail.17 Here, for simplicity, we will limit our discussion to phase velocity, defined as
that velocity at which the wave phase remains constant. The trial solution used in the
wave equation was
exp[ i ( t kx K 4 )];
(8.1)
therefore, the wave phase
t kx k 4 .
(8.2)
If we set
d
0,
dt
we find
d (k 4 )
vp
k
kdt
Substituting for k4 from Chapter 6, the phase velocity becomes
C3
vp
, (8.3)
k a0 c 2 k
where it is assumed that k and k4 are independent of time.
Now if mass is conserved, then
u j , j , (8.4)
where the uj are the components of the medium flow where vp is defined. By substituting
Eqn. (8.4) into Eqn. (8.5), the phase velocity becomes
C3
vp
u j , j.
(8.5)
k
a0 c k
In classical electromagnetism any uniform motion of the medium is not reflected in the
speed-of-light measurements. The same thing is true of the phase velocity given by Eqn.
(8.5). On the other hand, a divergence in the flow of the medium will affect the phase
velocity. The velocity change, owing to a divergence in the flow, is inversely
proportional to the density of the medium. Because of the change also is proportional to
parameter C3, which has not yet been completely determined by the wave solution, it
cannot be seen yet whether any envisioned experiment could measure the predicted
change in velocity. To do so would require completing the wave solution to determine C3.
One suggested experiment might be measuring the phase velocity in the divergent
flow coming out of a nozzle in a hypervelocity wind tunnel. Although such an

268

experiment might not be sensitive enough to detect the predicted velocity change
particularly, and would change because of the change in mass density.
Another possibility, which was suggested by Bobby G. Craig, is to measure the
travel time of a strong beam of gamma rays through a divergent flow of gas created by
explosives. This may create the largest divergent flow possible, but whether or not other
experimental difficulties could be surmounted to make reliable measurements is
unknown.
8.2 Index of Refraction
The change in the parameters and was mentioned in the speed-of-light experiment
discussion. From the plane wave solutions, we found that
i
(C1 C 3 )
0
a0 c
and
0
2
0

C 2 C3

Classically, the index of refraction for dielectrics is given by ()1/2. However, given the
wave solution, we must consider the boundary conditions as the wave passes through a
boundary between two media, determine the energy transmission and reflection
coefficients, and then find the index of refraction from a modification of Snell's law. That
is, the index of refraction should indicate the angle of the refracted wave with respect to
the incident wave.
A cursory look at a five-dimensional wave incidence upon a boundary produces
the relation
sin 0 k 2 ( 0 k 40
2 k 42 )
.
(8.6)
sin 2 k 0
k 0 sin 2
But from Chapter 6 we find
k

C3
C2

(C1

C2 )

1
2

if y=0. Also,
k4
Then we have
sin 0
sin 2
where

C3
.
a0 c

C4

2
0

[(a 0 c(C 40

C 42 ) (C 30 0
0 a 0 sin 2

C3
(C1 C 3 ) .
C2
Then, if the frequency is high enough,
1

269

C 32

)]

, (8.7)

C 42 ) (C 30 0 C 32 2 )]
sin 2
0
0 a0
We would be tempted to define as the index of refraction. On the other hand, the classical
notion of the index of refraction involves the ratio of the sin of the incident and refracted
waves. In Eqn. (8.7), the appearance of sin 2 in the right-hand side makes matters more
difficult. However, C45 is a phase angel we may set at zero, and we may choose the
reference medium, 0, to be free space for which 0 = 0; then Eqn. (8.7) can be written as
1
[C 32 2 a 0 cC 42 ]
2
a0 sin 2
sin 0
.
sin 2
0
Then we may define
(C 3
a 0 cC 4 )
,
(8.8)
a 0 sin
so that Eqn. (8.7) becomes
sin 0
2
,
sin 2
0
so long as the reference medium is free space. If we call the index of refraction, we find
that
2

[a 0 c(C 40

C3
C2

(C1

C3 )

(C 3
a 0 cC 4 )
,
a 0 sin

depends upon both the frequency and the mass density.


A possible experimental test may be obtained by applying rigorous boundary
conditions to the five-dimensional wave incident upon a boundary. This would verify or
correct the modified Snell's law given by Eqn. (8.7). The frequency dependence of the
refracted wave angle that was determined experimentally may be compared with the
predicted angle. Another comparison may be done by considering the density dependence
of the refracted wave.
8.3 Neutron Interferometer
A neutron interferometer can detect extremely small differences in forces upon each of
two neutron beans by using interference techniques. It can detect the difference in the
earth's gravitational field that is due to a height change of only 2 cm near sea level with
some thirty fringe shifts.
If the long-range character of the V field, as seen from the radial dependence
required for a fundamental particle,
Wg

e r
r
r
requires that the V field is to be interpreted as the gravitational field, then the force law,
J4
K
E
V,
c
Vr

270

would require that J4/c be interpreted as the gravitational mass density. This would
require
J4
mg
dV
c
v
to be the gravitational mass of a particle contained within the volume V. The gravitational
force on a particle in a gravitational, or V, field would be given by
F m gV .
This implies that the transverse V field accompanying the and B component in the
electromagnetic wave would apply a force on a neutron through an interaction with its
gravitational mass. Therefore, a beam of neutrons passing through a polarized layer
beam should be slightly deflected owing to the gravitational field component. This effect
would be most easily detected if, through the use of some appropriate mirror, a standing
optical wave could be created using a polarized laser beam. Then a neutron beam passing
through an appropriate part of the gravitational component of this standing wave would
have all the neutrons deflected in the same direction.
The sensitivity of the neutron interferometer may be such that, if one neutron
beam passes through a standing optical wave created by a laser of appropriate frequency,
very minuscule deflections could be detected. The appropriate laser frequency should be
chosen to maximize the predicted deflection. This, of course, requires that the wave
solution be completed so that the relative strength of the transverse gravitational
component is known; that is, because
C3
Vy
Ey ,
we must know C3 before we can choose the best laser frequency and power.
If a deflection is detected and has the predicted dependence upon laser frequency
and power, then the electromagnetic wave must be accompanied by a gravitational
component.
8.4 Nuclear Mass
We infer from the neo-coulombic electrostatic force that the nucleus may be made up of
complex orbits of electrons and protons, plus possibly positrons, as discussed in Chapter
4. The transcendental nature of the forces involved requires the use of a computer in
solving the equations of motion. Computer solutions may be obtained and the masses
predicted; then these predicted masses may be compared with the existing experimental
masses. A good comparison between the predicted and experimental masses, accounting
for possible errors introduced by any assumptions made to obtain a solution, would
increase the theory believability.
8.5 Gravitational Rotor
The continuity equation in the Dynamic Theory is
J
0
J a0 4
t

271

If we consider only steady state conditions such that


0
t
then we have
J
a0 4

J.

(8.9)

Equation (8.9) states that if one can create a non-zero divergence in the current density
then one creates a particular variation between the gravitational mass density and the
inertial mass density. This is in violation of both the classical conservation of charge and
Einstein's assumed equivalence principle.
Suppose we consider what happens when we pass a current into the apex of a
cone, as shown in Figure (16).
FIGURE 1: Current into the apex of a cone
Any position on the exterior of the cone is given by
y d fx
where the height of the cone is
d
,
h
f
therefore f = d/h.
If a steady current, I, is flowing into the apex of the cone, then at any x the current
density is given by
1
J
cos x y .
area
But the area is given by
Area
(2 y )t ,
while the
h
cos
d 2 h2
thus our current density vector becomes
J

I
2 yt

x oy .

(8.10)
d
h
Now we may form the divergence of this current density, noting that the above assumes
that the current density is a constant throughout the thickness t.
2

272

Jx
x

Jy
y

Jx
x

I cos
x 2 t[ d ( d / h ) x
hI cos
2 td

(8.11)

1
x (h

x)

Ih cos
1
2 dt
(h x) 2
Substituting the result, Eqn. (8.11), into equation (8.9), we have
J
Ih cos
a0 4
(8.12)
2 dt (h x) 2
This is a differential equation whose solution is
Ih cos
J4
J0
(8.13)
2 a 0 dt (h x) 2
Thus, the gravitational charge density is given by
J 40
J4
Ih cos
, (8.14)
2
c
c
2 a 0 cdt (h x)
Suppose we now consider two cones joined at their bases as shown in Figure (17).
FIGURE
The element of torque about the point A experienced due to the presence of the earth's
gravitational field V is found from the relation, torque = force x distance. Therefore,
2 J 40
Ih cos
d
2
( ty )Vxdx
( tyVx) (8.15)
2
c
2 a 0 cdt (h x)
The effect of the constant of integration term with J40 is to predict a constant torque
without current flow. Since this should have been noticed we shall take J40 = 0. Thus
Ih d (1 x / h)V
d
xdx cos
a 0 cd (h x) 2
(8.16)
I vx
d
dx cos
a 0 c(h x)
We obtain the total torque by integrating from x given by
dx
D d
h
or
x0 h( D / d 1)
to
x 0.
Thus, we have

273

torque

d
x h ( D / d 1)

I v xdx
.
a 0 c (h x)

Then the torque is determined by


I v
torque
x h ln(h
a0 c
I v
a0 c
I v
a0 c

h ln(h)
h
(d
d

x) | xx

0
h( D d )

cos

h
( D d ) h ln[h h( D / d 1)] cos
d

D) h ln[ D / d ] cos

or
I vh
[(1 D / d ) ln( D / d )] cos
(8.17)
a0 c
Suppose we pick some parameters; such as I = 10 amps, h = d = 0.1 m, D = 0.01 m.
With these parameters
1 0.01
(0.01)
(0.1) 2
ln
m
h cos
.1
0.1
[(1 D / d ) ln( D / d )]
c
3 10 8 (m / sec) (0.1) 2 (0.1) 2
torque

3.3059 10

10

sec 1 .

Thus, we have
v

torque

sec

) amps.
a0
We now need to choose a material to obtain the mass density, determine the gravitational
field V, and obtain a value for a0. Let's do them in the reverse order. Shock wave physics
investigations produced an extremely rough estimate of a0 which was
a 0 ~ 4 10 7 kg / m 4 .
The earths gravitational field strength, at sea level, is given by
volt coul 1
v
9.8
m kg
9.8

(3.3059 10

volt coul
m kg

2.4296 1011 (coul / kg )


4.0336 1011 volt / m.
If we choose aluminum as our material, then = 2.7 x 103 kg/m3 and our torque becomes

274

torque

(2.7 10 3 kg / m 3 )( 4.0336 10 11 volt / m)


7

(4 10 kg / m )

(3.3059 10

coul )

9.0 10 2 volt coul


9.0 10

nt

0.73757 ft. lb
nt

so that
torque 6.638 10 2 ft. lb.
This is not a very large torque, however, different cone parameters could be chosen to
optimize the torque.
There is another aspect which I don't yet know how to approach. In electric
motors there is a phenomena known as armature reaction which tends to limit armature
current far more than the armature resistance does. I suspect there is a
Figure 2. Upright cone, powered from within.
somewhat analogous reaction here that may further reduce the torque but it would take
time to investigate this possibility.
One final point on the creation of a gravitational field. A long held desire of
mankind is to be able to exert some control over the grip the earth's gravitational field has
over him. A slight variation of the above torque device might allow the generation of this
control.
Suppose we look at a single cone set upright as shown in Figure (18). From the
equations generated before for the torque we see that the lift force generated by this
simple device would be given by integrating the element of force
I V cos
dF
dx
a 0 c(h x)
or
I v cos
Force
{ln(h x) | 0x h ( D / d 1)
a0 c
I v cos
{ln[1 D / d 1]}
a0 c
I v cos
ln ( d / D )
a0 c
Thus, since V = 4.0336 x 1011 volt/m, the force becomes
I ln(d / D)h
Force
(4.0336 1011 volt / m)
2
2
a0 c d
h
Obviously, other physical shapes may achieve similar results; perhaps with an even
greater levitation force than the simple cone.

275

8.6 Nuclear Lamb Shifts


The atomic Lamb shift experiments were some of the best experiments in science
because of the comparison between predictions and experimental data. We now should
have the capability of doing gamma ray spectroscopy. Then preditions of the nuclear
energy levels as given by the nuclear model and potential given in Chapter 4 may be
checked experimentally.

276

Chapter 9 Epilogue
There are brief summaries at the end of Chapters 1 and 2, which give some measure of
summation of the Dynamic Theory. Here I wish to provide a little further discussion in
three areas. First, what is really new in the Dynamic Theory? Secondly, how might it help
us teach science and physics? The third topic is where might the theory lead? This does
not mean that there are not other new things presented in the preceding chapters. For
instance, it is not new to state that the Unit of Action appearing in the derivation of
Heisenberg's Uncertainty Principle depends upon the geometry. Anyone who has gone
through the derivation considering covariant differentiation has ended up with this
conclusion. Indeed, this is necessary in order to get the correct predictions of atomic
states. However, when one considered Einstein's vector curvature and the atom or
nucleus it was easy to argue that the vector curvature was so small that on the order of the
nucleus or the atom the curvature could not influence the unit of action to any meaningful
extent. This statement is perfectly true and easily supportable so long as your description
of the phenomenon does not involve a gauge function. What is new in the Dynamic
Theory is the appearance of a full gauge function. Now we can no longer assume that
geometry can be assumed away on the scale of the nucleus. Indeed we have shown that
all of the observation laws and experimental data are satisfied if the neutron is a proton in
nuclear orbit around and electron.
What is really new?
9.1 Only three basic assumptions.
The Dynamic Theory is based upon only three fundamental assumptions as stated in
Chapter 2. When I tried to count the needed fundamental assumptions in all of our
current branches of physics I came up with something more than twenty. I am aware that
one can generate considerable discussion about whether or not specific one of the twenty
plus assumptions were really fundamental or not. However, the criteria I used was
whether or not I knew of a method by which it could be derived from another
assumption. If the assumption could not be derived, then it must be fundamental.
Though I've been told many times that you can't do it, I see very little logic
restraining one from deriving Quantum Mechanics from a continuum theory. For
example, if one jiggles a guitar string that is tied down on only one end, there is a
continuum of solutions possible. On the other hand, if both the ends of the string are tied
down, and this represents an additional restriction, then only certain, and quantified
solutions are possible. Why not the same thing in the more broader sense of physics?
Table IV shows the necessary restrictive assumptions that must be made in order to start
with the

277

Table I. Restrictive Assumptions to Reach the Branches of Physics.


Branch of Physics Restrictive Assumptions
Classical Thermodynamics i. only a pdv work term
Special Theory of Relativity

Group A assumptions
i.
ii.
iii.
iv.

Newtonian Mechanics Group A assumptions, plus


(4-dimensional)
v.

only three spatial work terms

Electromagneto-Gravitic Fields
Maxwellian EM Fields

Group C assumptions
i.
ii.
Group C assumptions, plus
iii.

isolated system, dE=0


only 3 spatial work terms
near equilibrium
( ij=constant)
variation of paths

Quantized Gauge Potentials [e(-

/r)

/r]

isolated system, dE=0


gauge field equations

only three spatial work


terms
Group C assumptions, plus
iv.
j independent of path
v.
isentropic states

Strong Nuclear Force


i.

like particle forces,

i.

unlike particle forces,

i.
ii.

4-D Quantum Mechanics


r >> max

vi.

quantized gauge potentials

i.
ii.

quantized gauge potentials


geometrical unit of action

Weak Nuclear Force


Atom Physics (Classical)
Perihelion Advance

Group A assumptions, plus

1= 2
1

Redshifts

fundamental assumptions of the Dynamic Theory to the foundations of various branches


of physics and theories.
One example of the ability of the Dynamic Theory to derive the fundamental
assumptions of the various branches of physics which is not specifically pointed out in
Table IV is that of deriving Einstein's postulate concerning the limiting aspect of the
speed of light. We saw, in Chapter 2, that this is a direct result of the Second Law and
has, therefore, the same place in the theory as the limiting temperature in
thermodynamics.

278

9.2 Geometry is specified.


Scientists have searched for Lagrangians in their attempts to find a way of unifying the
fields of nature. One reason for doing so is because given a variational principle, such as
the Principle of Least Action, they can arrive at equations of motion and then field
equations. One problem with this approach is that they must make an assumption with
respect to the geometry of their space when they employ the Principle of Least Action.
This was the same necessity that Newton faced. Newton had no choice of geometries at
his disposal. There was only Euclidean geometry to be had. Einstein chose the
Riemannian geometry for use in his relativistic theories but rejected the more general
Weyl geometry when it was proposed. The Dynamic Theory leaves us no choice.
The fundamental laws specify what the geometry must be. To me this is extremely
satisfying from the sense that I feel that properly chosen laws should do just that; they
should leave no choice as to geometry.
A good many physicists know that one may choose a Lagrangian and use the
Principle of Least Action to arrive at equations of motion. Few, if any, know that by
doing so they must assume a type of geometry in this process and, thereby, have
interjected a restriction into the process. Therefore, the more power to a process that
does not allow this unwitting interjection of a restriction.
Another extremely important point with regards to the geometry is that within the
Dynamic Theory there are two geometries to be considered. This is as fundamental
within the Dynamic Theory as the fact that the concepts of heat and entropy are two
different things is fundamental in thermodynamics. The temperature is the integrating
factor in thermodynamics and, as such, plays a pivotal role between the heat and the
entropy. Within the Dynamic Theory the gauge function was shown to be a geometrical
integrating factor between the two geometries.
9.3 The Arrow of Time
The notion of irreversibility is embodied in the laws of thermodynamics, but is not in the
Newton and Einstein laws of motion. Yet mankind has been constantly aware of the
relentless march of history that Omar Khayym expressed in his:
The moving finger writes and having writ, moves on; nor all your Piety nor Wit,
Shall lure it back to cancel half a Line, Nor all your tears wash out a Word of it.
Many articles and books have been written on the subject of time symmetry, but
still questions remain. I first wrote on the Arrow of Time in 1981. The recent book, "The
Arrow of Time" by Peter Coveney and Roger Highfield is an excellent one. It sets forth
the problem of time symmetry in the mechanical theories in a very clear fashion.
The Dynamic Theory adopts generalizations of the classical laws of
thermodynamics as the basis for a new view of all physical phenomena. It was shown in
Chapter 2 how the adoption of these laws led to an integrating factor for purely
mechanical systems and that this integrating factor was strictly a function of velocity. In
Chapter 3 we saw the scope of the entropy increase further, though it still retained the
connection to the energy exchange by the integrating factor. Further, we
saw that for the isolated systems the Second Law produced an Entropy Principle that the
entropy could never decrease, or dS0. From this we derived equations of motion using the

279

expression for the square of the differential of the entropy. The square masks the fact
from the Second Law we obtained dS0. Further, in Chapter 2 we discussed the fact that
we could have obtained our equations of motion as third order equations in time, but
chose not to in order to have second order equations.
All of this means that the Arrow of Time is part of the Dynamic Theory from the
point of adoption of the Second Law.
9.4 Mass as a coordinate
Hermann Weyl titled his 1918 book "Space-Time-Matter" which somewhat implies that
matter is considered on the same footing as space and time. This, of course, is not
supported by the contents of the book. In the book he treats space and time as coordinates
while he leaves matter in its usual place in mechanics as either being the inertial or
gravitating mass. Further, the fact that the Dynamic Theory goes into five-dimensions
presents no new factor on that basis alone. Many other researchers have looked into five
dimensions in order to try to obtain the necessary degrees of freedom with which to build
into their theory the various fields thought to be needed to describe
the universe. What is different about the Dynamic Theory is that it treats mass on an
equal footing as space and time. This means that it is treated as a coordinate the same as
space and time. No other researcher, looking into five-dimensional systems, allowed any
physical significance to the fifth dimension. That is to say that they wished to have the
added freedom of the five dimensions but did not wish to allow the fifth dimension to
play a physical role as space and time were allowed to do.
9.5 Non-singular gauge potential
There are two aspects of the non-singular potential which makes it's appearance
something that is really different. The first is the fact that the maximum absolute value of
the potential is different for different particles. This is the extraordinary feature, which
leads to a description of phenomena usually reserved for the "Weak Force." The second
new aspect of the non-singular potential is that when applied to the planetary orbits the
potential produces the correct variation of perihelion advance as a function of orbit size
by itself. Numerous gravitational potentials were guessed and tried in attempts to obtain
an alternative to Einstein's General Theory of Relativity. One of the most severe tests for
these candidate potentials is the variation of the predicted advance as a function of the
orbit size. None of them could pass this test. The prediction of the perihelion advance
within the Dynamic Theory depends upon the orbital parameters in the same fashion as
the General Theory of Relativity.
A further utility of having a non-singular potential is that there is no need to
renormalize any functions including the gauge potential or any of its derivatives. In
Quantum Mechanics renormalization has always created problems and/or discussions.
With nothing to renormalize the problems and the need for discussions go away.

280

9.6 Unification of the Branches of Physics


Numerous researchers have worked on the problem and an untold amount of time has
gone into the attempts to find a unified field theory. Still a unified field theory is but a
promise of the future. The promises still hold out hope of a theory in the "near" future.
These promises remind one of the carrot on the pole. The more the poor horse tries to
reach the carrot the more the pole moves the carrot ahead. Nowhere in my reading of
books written on physics and unified field theories do I recall reading anything about an
attempt to unify the various branches of physics except in my own hand or word
processor.
The unification of the various branches of physics is the result of the
generalizations of the laws of classical thermodynamics and seeking how to obtain
equations of motion from them. Once the Entropy Principle was seen to provide a
variational principle from which equations of motion could be obtained the method of
unifying the branches of physics became visible.
While it may seem rather anticlimactic to be so brief about a point which has as
much significance as this point, there seems to be little that needs to be added to the
seven preceding chapters.
9.7 The pedagogical aspect of the Dynamic Theory
The precept that all current branches of physics (i.e. classical thermodynamics,
Newtonian mechanics, Special Relativistic mechanics, and Quantum mechanics) plus all
the forces in nature (i.e. electromagnetic, gravitational, weak, and strong nuclear) stem
from a single, simple set of fundamental laws may now be used to teach each of these
branches and forces by the application of a different set of restrictive assumptions. The
logic and rigor underlying this ability comes from the Dynamic Theory, which shows that
three fundamental laws may be used with restrictive assumptions to derive the
fundamentals of the various branches and forces. This not only displays the
interrelationships of the different branches and forces, but also sets up the excellent
teaching situation where different restrictive assumptions are the only difference between
the very dissimilar branches.
Students may be excused for some level of confusion during their advancement
through a school system wherein they are confronted with additional fundamental
assumptions as they encounter new branches of physics. This confusion may be
somewhat enhanced by the necessity of learning increasing skills in mathematics.
However, it is the sheer number of fundamental assumptions currently perceived
necessary for providing the basis for the existing branches of physics that is the source of
the confusion.
The notion that the different forces in nature might somehow be tied together is
the impetus behind the unified field theory hunt. Indeed this notion was behind the first
formal theory, which attempted to unify the electromagnetic and gravitational fields as
far back as 1836. Since then innumerable scientists have conducted investigations into
the unification of the forces, or fields, in nature. However, no theory has yet been
suggested that has gained undeniable experimental verification. Theoretical physicists

281

are still at work trying to find a theory that will ultimately unify the forces of nature. Such
is the belief in the unity of nature.
On the other hand the unification of the branches of physics has not enjoyed the
same level of attention. Indeed, relatively speaking, there is very little discussion in the
literature of this concept. This concept was the motivation for the development of the
Dynamic Theory. The unification of the forces comes as an additional feature.
To based a study of science and physics upon the three fundamental laws and then
lead to the various branches of physics by restricting our attention by specific
assumptions would seem to be a very logical way to learn about our universe.
9.8 Where to from here?
This might be one of my favorite topics. I seldom get the chance to do much work in this
area anymore and certainly have few with which to discuss the topic. One of the things
that I disliked about the course of instruction that I received was that the prevailing
attitude from virtually all of my professors was that "We now knew where advancement
could be made." If this were true then where was there any room for new work? Why
should I study a subject for which there was nothing new to be learned? This was a
terrible turn-off.
However, I didn't believe them then and don't believe them now.
From the Dynamic Theory's point of view there is a great deal of things to be
learned! For example, almost everything in the preceding chapters refers to systems
which are isolated and for which the Entropy Principle was employed. What do the nonisolated equations of motion look like? Wouldn't they describe a particle's transition from
one stable state to another?
We know that when one has a non-isolated thermodynamic system and are
pumping heat energy in or out of the system we must then minimize the free energy to
determine what happens. The same logic would apply here. If we wish to seek solutions
for non-isolated systems we need to minimize he free energy to obtain the non-isolated
equations of motion. This should give us the ability to better describe what happens in
these systems.
Notice, though, what this implies. We need to be vary careful when we are
considering mechanical systems to classify them as isolated or non-isolated. This is
something we never had to worry about before. It is also a way of seeing that there may
be a lot more to be learned if we do things differently.
The new things to be learned are not limited to non-isolated systems. Consider the
simpler question, can something go faster than the speed of light? From the relativistic
point of view one must answer that something going slower than the speed of light now
must forever remain slower than the speed of light. Similarly, things faster must remain
faster. But is the same conclusion true in the Dynamic Theory? The answer is no. From
the five-dimensional point of view the limiting aspect comes from
dq 0
v 2 g 44 2
1 2
,
dt
c
a 02 c 2
when time rate of change in entropy goes to zero. Should d/dt0 and g44<0 then v may be
allowed to be greater than c. What does this mean?

282

In nuclear weapons and reactors, mass is converted into energy. However,


Einstein's theory, which predicts the energy released in this conversion, says nothing
about the rate at which this conversion can or does proceed. On the other hand, the
Dynamic Theory not only provides an additional equation of motion that can be solved to
find the mass conversion rate as a function of time, but it predicts a limiting rate of mass
conversion. The limiting mass conversion rate comes from Eqn. (9.1) when v=0 and
g44=1, then
a 0 c.
max
Where does this lead us?

283

You might also like