Vous êtes sur la page 1sur 43

Foundations of Physics, Vol. 35, No.

5, May 2005 ( 2005)


DOI: 10.1007/s10701-005-4564-7

The Mathematical Basis for Physical Laws

R. Eugene Collins1
Received June 2, 2004; revised March 3, 2005

Laws of mechanics, quantum mechanics, electromagnetism, gravitation and


relativity are derived as related mathematical identities based solely on the
existence of a joint probability distribution for the position and velocity of a
particle moving on a Riemannian manifold. This probability formalism is nec-
essary because continuous variables are not precisely observable. These dem-
onstrations explain why these laws must have the forms previously discovered
through experiment and empirical deduction. Indeed, the very existence of elec-
tric, magnetic and gravitational elds is predicted by these purely mathemat-
ical constructions. Furthermore these constructions incorporate gravitation into
special relativity theory and provide corrected denitions for coordinate time
and proper time. These constructions then provide new insight into the rela-
tionship between manifold geometry and gravitation and present an alternative
to Einsteins general relativity theory.

KEY WORDS: mechanics; electromagnetism; gravitation; relativity.

1. INTRODUCTION

This paper is addressed to all who share the desire to not simply know
physics but to understand physics, that is, to know why the laws of phys-
ics have the specic forms we discover through empirical studies. As an
undergraduate in physics more than 57 years ago, my reaction, when rst
exposed to the amazing capability of Newtons laws to predict the motions
of physical objects, was to ask the professor Why F = ma? Of course
he simply assured me that these were empirical relationships which cannot
be explained by more fundamental laws. At the time I accepted this but
as I continued my education I became convinced that not only Newtons
laws but all fundamental laws of physics must have a common basis linked

1 1130 Pheasant Drive, Fort Collins, CO 80525, U.S.A.; e-mail: collins1130@comcast.net

743
0015-9018/05/0500-0743/0 2005 Springer Science+Business Media, Inc.
744 Collins

to the concept of a mathematical continuum. My rst attempt(1) to exploit


this idea was not successful, nor were most in succeeding decades; but now
it seems that this naive conjecture is at last conrmed.
To assign a precise value for any variable having a continuum for
its range would require determining every digit of a non-ending decimal
number. This is obviously impossible(1,2) so we here exhibit formalisms of
mechanics, electromagnetic and gravitational elds, relativity and quantum
mechanics as direct mathematical constructions based on the following
premises:
I. It is impossible to measure any continuous physical variable pre-
cisely so these must be treated as random variables described by
a classical probability formalism.
II. Any physical experiment is properly described by a classical prob-
ability distribution for the state variables of the system under
study with predicted values for observable functions of these vari-
ables dened as expectation values.
III. The probability distribution for any experiment is conditioned by
constant parameters of the system and its environment and val-
ues for variable parameters of the environment labeled by a time
parameter .
We also adopt the basic premise of Albert Einstein that mathematical
formalisms describing physical phenomena must be covariant, meaning
invariant under differentiable coordinate transformations. In the construc-
tions to be presented here this is accomplished with a covariant vector cal-
culus on Riemannian manifolds we have developed, rather than the tensor
calculus used by Einstein. This is described briey in Section 2 following.
These constructions use only laws of mathematics with no calls upon
physical concepts but we do point out the equivalence to familiar physics
when appropriate identications for symbols in the mathematical formalisms
are introduced. Thus it is demonstrated that laws of physics are actually laws
of mathematics imposed by the premises stated above. This exposes previ-
ously unrecognized characteristics of these domains of physical theory and
offers new avenues for further developments in physics. To the best of our
knowledge, there has been no prior work of the character presented here.

2. MATHEMATICAL BACKGROUND; THE COVARIANT VECTOR


CALCULUS

The Whitney embedding theorem(3) states that any differentiable


manifold M of dimension n can be embedded as a closed submanifold of a
The Mathematical Basis for Physical Laws 745

Euclidean manifold of dimension 2n+1. Here closed means that all limit
points for point sets in M are also points of M. However, if the metric of
M is constrained in an appropriate manner M can be embedded closed in
E K with K < 2n + 1 but K > n. For example, the surface of a sphere in
E 3 is a two-dimensional Riemannian manifold embedded closed in E 3 .
Since all points of M are also points of E 2n+1 , the directed line
element dR, from a point P of M to a near-neighbor point P  of M,
exists also in E 2n+1 . Specically, with the set of coordinates, j , j =
1, 2, . . . , 2n + 1, on E 2n+1 and x = x ( ), = 1, 2, . . . , n local rectilinear
coordinates on M, we may write dR, with the Einstein summation con-
vention, as the Euclidean displacement vector
R
dR = dx = 1 dx . (1)
x
This denes the vectors 1 (x) = R/x , = 1, 2, . . . , n at the point P
as local basis vectors on M in the directions of increasing x normal to
the coordinate hypersurfaces in E 2n+1 dened as x ( ) = constant =
1, 2, . . . , n. Thus, forming the Euclidean dot product, we have

dR dR = dR 2 = 1 1 dx dx , (2)

identifying the metric tensor on M as the symmetric form, g = 1 1 ,


which denes the manifold as Riemannian.
Note that if the x have the same physical dimensions and units as
the displacement dR then the basis vectors 1 and the metric elements g
are dimensionless. Of course one may always introduce curvilinear coordi-
nates but here we retain the local rectilinear x in our constructions. Note
also that only on at manifolds, free of curvature, are the 1 globally uni-
form with 1 /x = 0 everywhere for all , and only in this case are
the x actually components of the position vector R on the basis vectors
1 .
Now for any differentiable coordinate substitution the dx transform
as components of a contravariant tensor of rank one so Eq. (1) becomes
R x  R
dx  = 1 dx  ,

dR = 1 dx =  dx = (3)

x x x 
with the 1 = R/x transforming as components of a covariant tensor
of rank one thus assuring dR invariant. The critical point is that in this
vector format with these basis vectors we achieve a formalism that is com-
pletely covariant on Riemannian manifolds of any dimension and any met-
ric. Indeed, any vector V(x) = 1 (x)V (x) is an invariant, local scalar
in tensor language.
746 Collins

This covariant vector formalism is extended with the differential of


any differentiable scalar eld (x) written as
 
 
 
d = dx = 1
1 dx =  dR (4)
x x

by dening contravariant basis vectors 1 such that 1 1 = = 1/0 for


= / = and the gradient operator as  = 1 /x . Then the curl
and divergence of a differentiable vector eld A are introduced as A =
1 1 (A /x A /x ) and  A = (1 /x ) (1 A ) respectively.
With the Christoffel symbols {,} dened as components of the vector
eld 1 /x on the basis vectors 1 by 1 /x = {, }1 , this form for
 A can be shown equivalent to the familiar form from tensor calculus,

[( gA /x )]/ g, with g the absolute value for the determinant of the

metric tensor. This employs {, } = n g/x derived from properties
of determinants. Incidentally, the {,} are obviously symmetric in ,.
Another key relationship follows from (1 1 )/x = 0 and the deni-
tion for the {, } above, namely 1 /x = 1 {, }.
By eliminating tensor indices this vector formalism achieves a simplicity
and economy in notation which produces efcient derivations with very
transparent interpretations. For example we write the gradient of a vec-
tor eld A = 1 A as simply A but in tensor notation this is expressed
in a covariant dyadic as 1 1 (A /x + {, }A ). Here we recognize
A /x + {, }A as the covariant derivative of the contravariant vector
eld A (x) with respect to x , as dened in tensor analysis by the parallel
displacement process, but here this is obtained by differentiating 1 A as a
sum of products using the denition 1 /x = {, }1 and then renam-
ing some indices. This exposes the differential change dA in a vector eld
A(x), due to a differential change in position on a manifold, as the sum
of two parts, 1 dA and A d1 , with the rst due to differences in eld
strength and the second due to differences in manifold geometry at the two
positions. The complete development of this vector calculus is included in a
book just completed, with the same title as this paper, but further clarica-
tions will be provided here as we proceed with our demonstrations.

3. WORLD TIME, KINEMATICS AND THE CONTINUITY


EQUATION

We consider continuous motion of a point particle on a domain D


of a Riemannian manifold of dimension n, with local rectilinear coordi-
nates x , = 1, 2, . . . , n specifying the Euclidean position vector R(x) of
The Mathematical Basis for Physical Laws 747

the particle and a set of velocity coordinates, , = 1, 2, . . . , n, on a


companion manifold dened as components of a vector v represented on
the basis vectors of the x-manifold, 1 (x), = 1, 2, . . . , n, that is,

R
v= v = 1 (x)v . (5)
x
These x and v , being continuous, must be considered as random vari-
ables having a joint probability distribution dP = dP (x, v| ) conditioned
by a continuous, real scalar and constant parameters of the particle
and its environment whose values are xed in the preparation of the sys-
tem for observation, but these are not explicitly displayed here. However,
there are variable environmental parameters whose values are not xed by
the experimenter. Thus we have introduced the real scalar as a continu-
ous one-parameter label for the conguration of the external environment
dened such that if anything in the external environment changes then
the value of must increase. Thus must increase if the sun or moon,
or the hands on a clock, change position. Hence is time measured by
the experimenter with an appropriate clock in his frame of reference. This
frame of reference is the unchanging x and v manifolds. We will refer to
as world time.
We view dP = dP (x, v| ) as the distribution of results in multiple rep-
lications of an experiment predicted by mathematical theory, with pre-
dicted observed values for observables as expectation values under dP. By
the law of large numbers(4) this identication for observed values is sta-
tistically consistent with experimentally determined observed values com-
puted as the arithmetic means of measured values from replications of the
experiment. Here observables of the particle are real, single-valued, covar-
iant functions of the x and v . However, the x and v themselves are
not proper observables because the expectation values, as say v  under
dP , are not covariant. In particular R is the observed value for the posi-
tion vector at time predicted by dP (x, v| ) so dR/d is the predicted
value for the observed velocity v in the experimenters frame of reference.
Thus with R and v under dP the identity

dR
v (6)
d
actually denes the vector v as velocity in this probability formalism.
The probability dP (x, v| ) is now expressed in a more useful repre-
sentation as the factored form

dP (x, v| ) = dPv (v|, x)x (x| ) gd n x (7)
748 Collins


with dPx = x gd n x the marginal probability for the x on the vol-
n
ume element gd x, conditioned only by and the xed parameters, and
dPv (v|, x) the conditional probability for the v which is conditioned by
the x as well as and the xed parameters. Thus dPv has unit integral
over the v-manifold for all and any values of the x D while dPx has
unit integral over the domain D of the x-manifold for all values.
By appropriate choice for the order of x and v integrations, as per-
mitted by Fubinis theorem,(5) this representation for dP (x, v| ) yields the
expectation value for R as


R = R(x)x (x| ) gd n x. (8)
D

This follows because the rst executed integral of dPv over the v is unity.
Similarly the expectation value for v under dP is the expectation value for
under dPx


v =  = 1 (x) (x, )x (x| ) gd n x, (9)
D

where the components, (x, ), = 1, 2, . . . , n, of the differentiable vec-


tor eld (x, ) are dened by the rst executed integral as

(x, ) = v dPv (v|, x). (10)
v

Clearly these representations for R and v are covariant forms.
The dening relationship for v in Eq. (6) can also be represented in
terms of the vector eld (x, ), dened by Eqs. (9) and (10), in the form

dR
= v =  = (x, )x (x| ) gd n x, (11)
d D

since the expectation value for (x, ) under dPx = x gd n x is equivalent
to the expectation value for v under dP and this is the observed value for
v at time . However,  in Eq. (11) is unchanged if (x, ) is replaced
by u = (x, ) + q(x, ) with q(x, ) a vector eld, of the same charac-
ter as (x, ), but such that q under dPx is identically zero. With u thus
dened Eq. (11) can be written as
 
dR
= v = x gd n x = ux gd n x =u. (12)
d D D
The Mathematical Basis for Physical Laws 749

Then in the ideal limit of absolute precision for observations of the x ,


with x (x xc ( )) a Dirac delta function, this becomes

dRc dRc dxc dxc
= = 1c = c = uc (13)
d xc d d

with the subscript c indicating evaluation with the x replaced by the



xc ( ). Thus the probabilistic denition for v in Eq. (6) is validated
because this yields classical Newtonian kinematics for the idealized case
of precise observations of the x on a manifold with any metric.
Now observe that can be written as R because  =
x and R/x = 1 from Eq. (1). Indeed, R is actually the iden-
tity dyadic, I = 1 1 which can also be written as 1 1 as well as other
forms. Thus, for any vector w, w R is simply the vector w itself. The
form R for is important here because this in Eq. (12) implies that
x must satisfy the familiar continuity equation. This is shown as follows:
Observe that replacing x by x R in Eq. (12) and using the
identity

x R R (x ) +  (x R), (14)

we may apply the divergence theorem to the integral of the last term on
the right obtaining a surface integral of x R that is zero because x is
zero exterior to the domain D. Thus with / taken inside the integral for
R from Eq. (8) on the left and some rearrangement, Eq. (12) becomes
the identity
  
x n
R +  (x ) gd x 0, (15)
D

which states that the quantity in brackets is a scalar function, say Y (x, ),

x
+  (x ) = Y (16)

such that


RY gd n x 0. (17)
D

However, integrating each term of Eq. (16) over the domain D yields
  
x n
gd x +  (x ) gd n x = Y gd n x, (18)
D D D
750 Collins

and both integrals on the left are identically zero, so the integral on the
right must also be identically zero. The rst of these integrals is zero
because any x must have unit integral over D for all values and the
second is zero since the divergence theorem, with x zero exterior to D,
yields a surface integral which is zero, and these are identities.
A form for Y which assures the right member of Eq. (18) identically
zero for any x is

Y =  (x q) (19)

since by the divergence theorem the integral of this over all x D is zero
for any bounded, differentiable, vector eld, q(x, ), because x is zero
exterior to D. However, this form for Y is valid only if q satises the spe-
cic constraint imposed by Eq. (17). This constraint is exhibited in a more
informative representation as follows: Substituting Eq. (19) for Y into Eq.
(17) yields the requirement for q as


R (x q) gd n x 0 (20)
D

and here we again use the identity from Eq. (14) but now with q replacing
. This provides Eq. (20) in the form
  

(x qR) x q R gd n x 0 (21)
D

and again using the divergence theorem, the rst term yields a surface
integral which is zero because x is zero exterior to D. Finally, noting
that R is the identity dyadic, as discussed following Eq. (13), we see that
q R = q (1 1 ) = q so we have the remaining term of Eq. (21) as


qx gd n x 0 (22)
D

which must be an identity. Therefore the vector eld q and the scalar eld
x must be functionally related such that q 0 for any x . Thus q is
the same vector eld used to write  in Eq. (11) as u in Eq. (12) with
u = + q.
Now, with Eq. (19) for Y in Eq. (16), we nd upon rearrangement
that
x
+  (x u) 0 (23)

The Mathematical Basis for Physical Laws 751

is an identity with the vector eld u dened as u = + q with q(x, )


a differentiable vector eld having zero expectation value under any x .
Thus imposing Eq. (6) to provide a probabilistic denition for v in terms
of dR/d requires that x satisfy the continuity equation with the vector
eld u dened as + q. This also implies conrmation for dR/d equiv-
alent to u(x, ) under x as in Eq. (12).

4. PROBABILISTIC DYNAMICS

In conventional mechanics the dynamics of a physical system is deter-


mined by laws specifying dv/d in terms of forces, that is, essentially
Newtons law, F = ma! Thus we now construct equivalent relationships
determining dv/d in this probability formalism as identities.
With v under dP expressed as u under dPx as in Eq. (12) we have
   
dv d u x n
= ux gd n x = x + u gd x. (24)
d d D D

This is then modied with  (x u) for x from the continuity
equation in Eq. (23) and modied further using the identity

u (x u)  (x uu) + x u u (25)

to substitute for u (x u). The integral of the rst term on the right
is then zero by the divergence theorem because x is zero exterior to
D. Thus we obtain
   
dv u n du n
+ u u x gd x x gd x, (26)
d D D d

which actually is an identity because the continuity equation is an iden-


tity. Note that we have here expressed the convective total derivative of u
in the rst integral as du/d in the last integral. With Eq. (24) this shows
that du/d = du/d  for these expectation values under dPx .
Eq. (26) is then modied with yet another identity by substituting for
the u u term with the identity,
 
1
u u  u u + u u, (27)
2
752 Collins

where u is the dyadic, curl u = 1 1 (u /x u /x ), as introduced


in Section 2. Thus we have the identity
   

du n u 1
x gd x + u u + u u x gd n x (28)
D d D 2

with both members being covariant representations for dv/d .


We now call upon the Hodge-DeRham theorem(6,7) which asserts that
any differentiable vector eld on a differentiable manifold has a represen-
tation as the sum of an irrotational eld, a solenoidal eld and a har-
monic eld. With this we have two equivalent representations for u as

u =  + A = P (29)

where  0 and  A 0 while P is a dimensionless function dened,


to within an additive function of alone, by the path integral
 R  R
1
P= u dR = + A dR = + (30)
R0 R0

with and real, scalar constants providing dimensions and units to the
dimensionless scalar elds, and P, and the solenoidal vector eld A. Here
the additive harmonic eld is included in Eqs. (29) and (30) as p + a
with p a harmonic component in P and a a harmonic component in
A. That is, p and  a are identically zero while  p and a
are zero everywhere except on isolated singularities of zero measure with

respect to gd n x. Note also that since A is not everywhere zero the inte-
gral term is path dependent such that the scalar eld P is multivalued.
This characteristic is examined in detail in another paper but need not be
addressed here because P is single valued.
These representations for u are now selectively introduced in Eq. (28)
with P for u in u/ and  + A for u in u to arrive at the
identity
   

du n P 1
x gd x  + u u + u A x gd n x (31)
D d D 2

since  is identically zero in the last term. Also we have interchanged


the order of operations / and  on P in the rst term to group the
elements (P/ + u u/2) under the  operator.
We then make the critically important observation that the integrals
in Eq. (31) remain identical if we replace the real vector quantity
The Mathematical Basis for Physical Laws 753

(P/ + u u/2) by any real, single-valued, irrotational vector eld,


as say 0(x, ), provided these differ only by a real irrotational vector
function of x , Q(x ), such that the expectation value for Q under the

probability dPx = x gd n x is identically zero for any x . Thus if

P 1
+ uu =0+Q (32)
2
with Q(x ) such that Q 0 under any x , then the identity in Eq. (31)
assumes the form
 
du n
x gd x {0 + Q + u A}x gd n x (33)
D d D

with the contribution of the term Q to the integral therefore null.


The integrands in Eq. (33) are not algebraically identical, even though
the integrals are identically equivalent, but these integrands are quantita-
tively equivalent for any x . Therefore

du
= u A + 0 + Q (34)

is a valid equality for all x D and any and constitutes the dynamical
equation of evolution for u(x, ), Furthermore this dynamical equation can
actually be obtained directly as the gradient of Eq. (32) by making use of
Eq. (29) for u together with the identity for (u u/2) from Eq. (27) and
the denition of du/d as u/ + u u.
Now in the limit of absolute precision for observations of the x ,
with the limit for x (x| ) being the Dirac delta function (x xc ( )) and
Q 0 under all x , the identity in Eq. (33) assumes the limiting form

duc
= uc A + 0 (35)
d
in which uc (xc ( ), ) denotes the limiting form for the expectation value
for u(x, ) under x = (x xc ( )), and all functions of the x in this

equation are evaluated on the trajectory x = xc ( ), = 1, 2, . . . , n.
Thus, in the limit of absolute precision for observations of quantities
determined by the x , we obtain dynamics for a particle with Eq. (35) as
the deterministic equation of motion.
Some clarication of notation is called for here since the symbols 0,
A,  and  in Eq. (34) have been retained unchanged in Eq. (35) while in
fact these quantities are changed, but only in one key characteristic. While
0 and the A may have an explicit dependence on , as well as on the x
754 Collins

in Eq. (34), the basis vectors 1 , and therefore the operators  and , are
functions only of the x , but in Eq. (35) all of these quantities have gained

an implicit dependence on through replacement of the x by the xc ( ).
This limit of absolute precision denes conditions necessary to jus-
tify the assumption of classical physics, held since the days of Galileo and
Newton, that the trajectory of a particle can be described by coordinates

xc ( ) as continuous functions of a time parameter.
Before we proceed to further details of this construction of particle
dynamics we show by direct construction that there does exist a form for
Q as a function of x such that the identity Q 0 is valid. Writing x
as Rx2 (x, ) with Rx a real, single-valued function of the x and , differ-
entiable to all orders, we dene Q as

2 Rx 2 x
Q = (36)
Rx x

with a real constant. Since for some x and , Rx may be negative we



have Rx = x but in Q these signs cancel. Then we construct the iden-
tity for x Q

2 Rx 
2
Rx   IRx 2 Rx 2 (Rx ) (Rx )+IRx Rx (37)
Rx

with I the identity dyadic 1 1 . The integral of this over the domain D,
with the divergence theorem, yields a surface integral for the bracketed
dyadic which is zero because x and therefore both Rx and Rx are uni-
formly zero exterior to the domain D. Thus the function Q(x ) in Eq. (36)
assures Q identically zero for any x .
Observe that the precision limit for this probabilistic formalism yields
terms on the right in Eq. (35) having forms appropriate for identication
as electromagnetic forces on a moving charged particle. Specically, simply
raising and lowering the index we see the rst term as

uc A = 1 u
c F = 1 uc F ,

(38)

with F denoting A /x A /x , which, in the case of the n = 4


space-time manifold with Lorentz metric (1, 1, 1+1), has the form of
the electromagnetic eld tensor derived from the four potential A. Indeed,
Eq. (38) is then structured precisely as the Lorentz force per unit mass in
the relativistic format if is identied as e/(mc), with e and m being the
charge and mass, respectively, of a particle and c the speed of light. See
for example Barut(8) p. 55. This also requires that the uc be identied as
The Mathematical Basis for Physical Laws 755

components of the proper four velocity of the particle in the Lorentz met-
ric. In this relativistic case there remains the term 0 yet to be identied
in terms of physical quantities. Later this will be shown to be the gravita-
tional force per unit mass acting on the particle for this manifold dimen-
sion and metric.
Note that the equation of motion in Eq. (35) is essentially an iden-
tity. Indeed, these laws of mechanics, as well as the continuity equation,
are simply consequences of the denition for the v by Eqs. (5) and (6),
and in this sense these are identities. Thus we nd that not only Newtons
second law, F = ma, appears here as a mathematical equation relating
two equivalent representations for the same mathematical quantity as F/m
and a, but also this purely mathematical construction yields specic laws
of force. However the elds appearing in the force terms of this equa-
tion of motion must be specied if this equation is to be applied to pre-
dict the motion of a particle. Here it will be shown that eld equations for
these elds are also dictated solely by mathematics, but rst we point out
other aspects of these constructions.
The continuity equation, Eq. (23),

x
+  (x u) = 0 (39)

and the dynamical equation, Eq. (34), with Eq. (36) for Q,
 
du 2 x
= u A + 0  , (40)
d x

are a pair of nonlinear eld equations for x and u. For specied elds
0 and A, with specied boundary and initial conditions for x and
u, these equations determine the scalar eld x (x| ) and the vector eld
u(x, ). In principle, for a given metric, these determine not only the evo-
lution for x and u but also the distribution of possible trajectories for a
particle as streamlines everywhere tangent to the local velocity u. However,
because of singularities noted below, this is not practical.
As already noted just following Eq. (34), this dynamical eld equation
can also be constructed directly from Eq. (32), with Eq. (36) for Q. With
u as P from Eq. (29) and Eq. (36) for Q we may write Eq. (32) as


P 1 2 x
+ |P| = 0
2
, (41)
2 x
756 Collins

or, with P = + from Eq. (30), as


   
1  2
 2 x
+  + A = 0 , (42)
2 x

where is the path integral of A multiplied by / in Eq. (30). In another


paper(9) we show why and how may be single-valued. These have forms
like the familiar HamiltonJacobi equation for a particle.
Note that in view of the representation for Q in the last terms of Eqs.
(40)(42) the vector eld u must acquire singularities on zeros of x which
evolve with . Hence u must contain a harmonic eld component just as
discussed in relation to the HodgeDeRham representation for u in Eq.
(29). This can be shown to be the eld q, introduced in Eq. (12), which
must be -dependent. However, since both Q and q under dPx are
identically zero, the precision limit u uc is free of these singularities.
With assignments for the constants , and as h/m, e/(mc) and
h2 /(2m), respectively, we nd that on the E 3 x-manifold, with  and
A As , Eqs. (39)(42) are identical in form to the so-called quantum

equations of motion with the term (h2 /(2m))( 2 x )/ x called the
quantum potential. Historically Eqs. (39) and (41) were derived from the
Schrodinger equation of quantum mechanics,

h2 2
ih = + 0, (43)
2m

by substituting Rx exp iP for the solution in this equation and separat-


ing into real and imaginary equations with x identied as Rx2 and u iden-
tied as (h/m)P. To be consistent with conventional applications of the
Schrodinger equation to a charged particle, 0 in Eq. (43) must be identi-
ed as eA4 . This formalism was rst constructed by Madelung(10) in 1926
and called the hydrodynamic representation for the Schrodinger equation;
it had limited development until rediscovered by Bohm(11) in 1952 and has
since been highly developed as described in the very comprehensive text of
Holland.(12)
In similar fashion the formalism in Eqs. (39) and (42) for an
n-dimensional at manifold with any diagonal metric is derived by the sub-
stitutions  = Rx exp i and the integral term from Eq. (30) in the
equation

 1
i = (i + A) (i + A)  + (0 + / )  , (44)
2
The Mathematical Basis for Physical Laws 757

with the identications x = Rx2 and u =  + A. For n = 4 with


x 4 = ct and the Lorentz metric tensor (1, 1, 1, 1) Eq. (44) is essen-
tially the Stuckelberg equation,(13) which has been extensively developed as
parameterized relativistic quantum mechanics, or PRQM, as described in
the text of Fanchi.(14)
An alternative exploitation of the probability formalism based on
dP (x, v | ) for at manifolds is found in Collins(9) which demonstrates
a complete equivalence of the Hilbert space formalism of conventional
single-particle quantum mechanics and probability theory, not simply the
Schrodinger and Stuckelberg equations. But now we turn to consideration
of eld equations, as identities, to justify specic physical identications
for the scalar eld 0 and the vector eld A appearing in these equations.

5. THE COVARIANT FIELD EQUATIONS

Since A is the solenoidal element of the velocity eld u required by


the HodgeDeRham theorem, A must satisfy the solenoidal condition, 
A 0 However, A appears in the equation of evolution for u in Eq. (34)
only as the dyadic eld F = A. Therefore
 
A A
F = 1 1 F = 1 1 = A (45)
x x

is the basic denition for the tensor F in this formalism. For manifold
dimension n, the tensor F is antisymmetric with n(n1)/2 distinct scalar
elements, say c1 , c2 , . . . , cN with N being n(n 1)/2. These can be viewed
as the components of one or more differentiable vector elds on the man-
ifold. Since A denes a pseudo-vector on the basis vectors within the
manifold only for n = 3, the smallest manifold dimension to be considered
is n = 3. For n = 3 the matrix [F ] is a 3 3 matrix having only three
distinct elements which are identied as components of a three vector, ci .
These must appear in the matrix [F ] with appropriate signs assuring this
matrix antisymmetric with zeros on the principal diagonal.
For n = 4 there are six distinct elements of F which are the three
components of cI plus the three components of a second three vector, cII .
The latter must appear, with appropriate signs and positions, in an added
row and column appended to the 3 3 matrix containing the components
of cI to form a 44 antisymmetric matrix with zeros on the principal diag-
onal. Proceeding in this manner, components of a four vector, cIII , with
appropriate signs and locations, must be introduced in another added row
and column to form the matrix F corresponding to n = 5. Continuation
758 Collins

of this process for succeeding values of n yields the collection of vectors


ci , cii , ciii , . . . , cn2 lling the matrix [F ]. However, since any vector ca

has two equivalent representations, as 1 ca = 1 ca , the assignment of
components of these vectors in [F ] must stipulate either covariant or
contravariant components.
Each of the ca is a differentiable vector eld on the manifold and
therefore must have a representation of the form,

ca = a + Ka with  Ka 0 (46)

as dictated by the HodgeDeRham theorem. Hence, applying divergence


and curl operations on this yields

(a)  ca , = a , (b)  c a = a , (47)

where a is actually a and a is Ka , since a 0 and Ka


0. These constitute forms for eld equations for each vector ca if the scalar
and dyadic quantities, a and a are determined.
The only characteristic of the F tensor entering the above analysis
was its anti-symmetry. However, the denition of the F as elements of
A, as in Eq. (45), requires that these F satisfy the metric-independent
Bianchi identity

  F F F
F, =
+
+ 0. (48)
x x x

This is an absolute identity, having this same form on any manifold of


dimension n  3. That the F do satisfy this identically
 is 
conrmed sim-
ply by substitution of their dening form, A x A x .
For manifold dimension n  3, this identity provides n!/ [3!(n 3)!]
rst-order, scalar, partial differential equations in the F which must be
consistent with the eld equations for the vectors cI , cII , cIII , . . . , c(n2) in
Eqs. (47). That is, the assignment for covariant, or contravariant, compo-
nents of these vectors in the matrix [F ] must not only carry signs assur-
ing antisymmetry but also these assignments must assure equivalence of
Eqs. (47) and (48). This makes these component assignments in the matrix
[F ] specic and also determines unique forms for some of the a and a
appearing in Eqs. (47).
Executing this program rst for n = 3 determines that I =

cI n g and Fij = ij k cIk for i, j, k in the set 1, 2, 3 with ij k
the LeviCivita tensor. Then for n = 4 this determines that, since the
F are already xed for , in the set 1, 2, 3, the added terms are
The Mathematical Basis for Physical Laws 759

F4 = F4 = cii for = 1, 2, 3 and ii = cI /x 4 , with, of


course, F44 = 0. Continuing in this manner the full matrix [F ] is con-
structed. For example for n = 5


0 ci3 +ci2 cii1 ciii1
+c3 0 ci1 ciii2 ciii2
i

[F ] = ci2 +ci1 0 cii3 ciii3 . (49)

+cii1 +cii2 +cii3 0 ciii4
+ciii1 +ciii2 +ciii3 +ciii4 0

Observe that in this manner equations for ci , cii , ciii , . . . , cn2
as in Eqs. (47), are generated, and assignments for all elements of [F ]
in terms of these vectors are determined. However, there remain the
eld equations for ci , cii , ciii , . . . , cn2 required by Eqs. (47),
yet to be specied. These constitute n additional scalar equations in
the elements F yet to be dened. Note also that the equations for
cii , ciii , . . . , cn2 deduced above are actually metric-independent iden-
tities simply because the Bianchi relationship is an absolute identity.
Before we consider constructions for these additional eld equations
in the F , we should note that the Bianchi identity in Eq. (48) can be
represented in the covariant vector notation as
 

| F = 1 1 1 F, 0 (50)

in which {F, } is the conventional notation in tensor calculus for the


sum of the three derivatives F /x generated by the normal cyclic per-
mutation of indices, , as in Eq. (48). Actually  | F can
be constructed as F + P F + P 2 F with P the cyclic permutation oper-
ator on tensor indices of F, since all terms arising from derivatives of
basis vectors add to zero such that Eq. (50) follows. Thus Eq. (50) is one
covariant eld equation in the F on any manifold of dimension n  3,
having any metric whatever, which is equivalent to n!/ [3!(n 3)!] scalar
partial differential equations in the elements F .
The n additional scalar eld equations required to determine the F
are obtained as follows: Now F can be written as either of the equivalent
forms, 1 1 F or 1 1 F , and the divergence of F must exist. In fact,
 F denes a vector eld J as

1
J =  F = 1 ( gF ) (51)
g x
760 Collins


since 1 /x = {, }1 , 1 1 = and {, } = n g/x from
Section 2. Hence


2 gF
J = 0, (52)
gx x

is an identity simply because F = F . Therefore, if the n compo-


nents J of J are specied functions of the x satisfying  J 0, then
Eq. (51) represents n scalar equations in the F , and these F are sim-
ply related to the F via elements of the metric tensor as g g F . Of
course, these F are already dened in terms of components of the vector
elds, ci , cii , ciii , . . . , cn2 by the constructions based on the Bianchi iden-
tity and the HodgeDeRham theorem as shown in Eq. (49) above.
Thus we now have a system of general covariant eld equations

(a) 
| F0
(b) F=J (53)

(c) J 0

determining the F tensor for any manifold dimension n and any metric.
This system of equations is based on just two denitions and two iden-
tities, F is dened as A and J is dened as  F while  | F 0 and
 J 0 are identities. Thus the equations of evolution for x and u in
Eqs. (39) and (40), and the eld equations in Eqs. (53), are fully covariant
and based on identities.
However it must be noted that Eq. (53b) dening the vector eld J
is not unique; actually any antisymmetric, differentiable dyadic P denes a
vector eld J as J =  P such that  J 0 but only those dyadics
P having elements which are functions of the ca elds dened in F, such
that the equation  P = J yields the eld equations for cI ,  cII , . . .
required by the HodgeDeRham theorem in Eqs. (47), are acceptable. This
alternative can be applied to construct eld equations in material media
but here, in matter-free space, we retain only F itself as this dyadic P.
Not only does Eq. (53a) provide the forms for ci , cii , 
ciii , . . . , cn2 required by Eqs. (47) but also Eq. (53b), provides the forms
for ci , cii , ciii , . . . , cn2 based on the HodgeDeRham theorem
in Eqs. (47). To show this we note that the dyadic F = A can be repre-
sented directly in terms of these eld vectors as

F = Fi + Fii + Fiii + + Fn2 (54)


The Mathematical Basis for Physical Laws 761

with each of the dyadics Fi , Fii , . . . containing only one vector eld. Spe-
cically, with the ca assigned in the matrix F as in Eq. (49), we have

(a) Fi = 1i ij k cik 1j
(b) Fii = 14 cii 1 1 cii 14 (55)

(c) Fiii = 15 ciii 1 1 ciii 15

with forms Fiv , Fv , . . . , Fn2 , just like Fiii for n > 5, each also containing
only one vector eld. This representation contains all F elements, so we
construct  F by applying the linear operator, , onto F in Eq. (54)
which yields a sum of terms,  Fi +  Fii +  Fiii + +  Fn2 , each
containing only one of these vector elds.
With this construction for  F Eq. (53b) does yield explicit equa-
tions for ci , cii , ciii , . . . , cn2 , just as called for by the Hodge
DeRham construction in Eqs. (47).
From the representation for A as the dyadic F, which has the form
Fi + Fii + + Fn2 in Eq. (54) with only one corresponding vector eld,
ci , cii , . . . , cn2 in each of the dyadics Fa , we nd that the general equation
of evolution from Eq. (34) can be written as the covariant form

du 
= u Fi + u Fii + + u Fn2 + G + Q (56)
d

or, with the Fa from Eq. (55), as

du
= [us ci + u4 cii us cii 14 + u5 ciii uciii 15 + un cn2
d
ucn2 1n ] + G + Q, (57)

where G denotes 0. Here, in the rst and third terms of Eq. (57), us
is the three vector of components u , = 1, 2, 3 of u because ci and cii
are three vectors. Also, with the equivalence ij k = ikj and renaming of
indices, j k and k j , Fi in Eq. (55) was modiedto arrive at  the
  j
cross-product form since us Fi then appears as 1 u 1i ij k ci 1k and
j
with the Einstein summation convention this is ijk ui ci 1k = us ci .
In this we see that each of the eld vectors ci , cii , . . . , cn2 con-
tributes separately to an additive force on the particle, but in addition
to these forces arising from F, there also appears the force due to the
n-dimensional eld vector G = 0. Of course all of these forces are
actually forces per unit particle mass.
762 Collins

Field equations for the vector elds cI , cII , . . . , cn2 dened by F =


A were provided in Eqs. (53), but these did not address the addi-
tional eld G. Since G = 0 conforms to the requirements of the
HodgeDeRham theorem, simply applying curl and divergence operators
to G yields

(a) G = 0 0,
20 0
(b)  G =  0 = g
g {, } = JG (58)
x x x
as appropriate eld equations for the vector eld G, with JG a scalar
source term just as in Eqs. (47) for the ca . However, the (a) equation here
is a trivial identity comparable to  | F 0 for the dyadic eld F, while
in the (b) equation g = 1 1 is the
 contravariant metric tensor and
g {, } represents 1 1 x which is clearly zero on any at
manifold. These two equations, together with Eqs. (53), form a complete
set of covariant eld equations for the covariant equation of evolution
in Eq. (57), determined by the scalar function JG and the n-component
vector eld J which must be prescribed.
However these covariant equations are useless without a prescribed
dimension and metric for the coordinate manifold. Since there is a distinct
classication for manifolds as at or curved we begin our comparisons of
this formalism to some elements of familiar physics by restricting the for-
malism to at manifolds. This will provide new insight into important ele-
ments of physics.

6. THE FOUR-DIMENSIONAL FLAT MANIFOLD


AND THE LORENTZ METRIC

Here a particle is called relativistic because it is described as moving


on a four-dimensional space-time manifold with spatial coordinates x ,
= 1, 2, 3 having dimensions of length, and a fourth coordinate, x 4 = ct,
with t the coordinate time of the particle and c a scalar constant providing
length units to x 4 . This coordinate time is dened as a never-decreasing,
one-parameter label for the conguration of the changing external envi-
ronment as perceived in a frame attached to the moving particle, while
the parameter is retained as the one-parameter label for the congura-
tion of the changing external environment as seen in the observers frame
of reference. This manifold is dened to be at with globally uniform
basis vectors, 1 /x = 0 for all , , which provide the metric g =
1 1 diagonal as (gss , gss , gss , g44 ). Shortly we will see g44 = +1 and
gss = 1 as mathematically necessary in the construction of the eld
The Mathematical Basis for Physical Laws 763

equations for this particle. In this diagonal metric we have assigned the
three-dimensional spatial manifold to be isotropic simply because this is
consistent with our local perception of space.
On this four-dimensional x-manifold the eld dyadic F = A contains
only the two three-vector elds cI and cII with elements congured as in
Eq. (49). Therefore we here adopt the notations

B = cI and E = cII (59)

since B will be identied as the magnetic eld and E the electric eld
acting on the particle in constructions following.
With these elements specied, applying Eq. (53a), 
| F 0, to the eld
dyadic yields four scalar equations in the components of B and E which
can be written in the Euclidean three-vector format as

(a) B 0
B . (60)
(b) E + 4 0
x
These are actually trivial identities. This is demonstrated by substituting
the denitions for the E and B as elements of the eld tensor F = A
into Eqs. (60) which yields both equations identically zero. These deni-
tions, based on Eqs. (49) and (59), can be represented in the three-vector
format as

(a) B = As
As , (61)
(b) E = A4 4
x
with As the three-vector of spatial components of the solenoidal eld A
and A4 the coordinate time component. Eqs. (60) have the familiar forms
of the metric-independent Maxwell equations.
The metric-dependent pair of Maxwells equations in free space fol-
low directly from application of Eq. (53b) to this dyadic F, but only with
additional consideration for the form of the metric tensor.
Since Eq. (53b) is equivalent to Eq. (51) and here the metric tensor is
diagonal as (gss , gss , gss , g44 ) with all g either +1 or 1 for the rectilin-
ear coordinates, the determinant g is either +1 or 1 so Eq. (51) becomes
F
 F = 1 = J. (62)
x
Therefore we employ the identity

F = 1 F 1 = 1 g F g 1 = 1 F 1 (63)
764 Collins

to express the F in terms of the known F in Eq. (49) with the iden-
tications cI = B and cI I = E. Since g g 1 this equivalence can be
written as
 1
F = g g F . (64)

For this particular metric all of these coefcients are either +1 or 1. For
2 )1 is positive,
all Fij , i, j = 1, 2, 3 in FI this coefcient is +1 since (gss
while for the F4i and Fi4 , i = 1, 2, 3 in FII the multiplier is (gss g44 )1
which is +1 for the gss and g44 having a common sign or 1 for different
signs.
Thus with FI and FII from Eq. (55) we form F as FI + FII to obtain
Eq. (62) as

1
1 1 i ij k
B j 1 k + (g ss g 44 ) (1 4 E E1 4 ) = 1 J (65)
x
in which we have raised and lowered all corresponding indices in F and
substituted cI = B and cII = E. Also we have used the equivalence ij k =
ikj and renamed indices as j k and k j . Thus executing the diver-
gence operation yields
 
E
B + (gss g44 )1 E1 4 = 1 J

(66)
x 4

since with the Einstein summation convention ij k Bj/x i 1k is B. Then


equating corresponding components on the two sides yields

(a) B + (gss g44 )1 4 = Js
E
x . (67)

(b) (gss g44 )1 E = J4

Here we have collected the three component equations for  = 1, 2, 3 into


Eq. (67a) with Js the three vector of spatial components J  ,  = 1, 2, 3.
Clearly, these are the remaining pair of Maxwell equations if the metric ten-
sor elements gss and g44 differ in sign. Actually these signs must be gss = 1
and g44 = +1 because of mathematical constraints described as follows:
In source-free regions, Js = J 4 = 0, we exploit familiar manipulations
of vector calculus in Eqs. (60) and (67) to show that B must satisfy

1 2B
2B + = 0, (68)
gss g44 (x 4 )2
and by similar constructions E must satisfy this same equation.
The Mathematical Basis for Physical Laws 765

Now if (gss g44 ) > 0 this equation admits solutions for B, and simi-
larly for E, which are exponentially unbounded on the space manifold as
x 4 +, but if (gss g44 ) < 0 then only bounded solutions are admit-
ted. Thus to eliminate these exponentially unbounded solutions, we must
assign (gss g44 ) < 0. Then for |gss | and |g44 | both equal to unity we nd
the metric tensor on this four-dimensional, space-time manifold must be
either (+1, +1, +1, 1) or (1, 1, 1, +1), but there is a requirement
which determines that the latter is the only correct form.
Recall from Section 2 that the embedding theorem yielded the square
of the line element on the Riemannian manifold as the positive product
form in Eq. (2) and here, on this at, pseudo-Euclidean manifold, with
the diagonal Lorentz metric tensor (1, 1, 1, +1), in the precision limit,
x (x| ) (x xc ( )), the expectation value for this is

dRc2 = dxcs dxcs + c2 dtc2 > 0. (69)

Thus since dxcs dxcs = |dxcs |2 for this diagonal metric, this
implies
  that
 the magnitude of the expectation value for particle velocity
dxcs dtc  on the spatial manifold must be less than the speed c. Here the
notation employs the displacement vector as dxcs = 1i dxcs i on the E 3 spa-

tial manifold while dxc = cdtc is that on the coordinate time axis in the
4

precision limit, with all x xc ( ) in this limit. Were the alternative met-
ric selected the opposite would be implied, that is, the particle speed would
always exceed c in magnitude which is obviously not acceptable.
Thus it is the combination of requiring the space-time manifold at
and spatially isotropic, together with bounded solutions for the source-free
Maxwell equations as x 4 + and the constraint of dRc2 real and posi-
tive imposed by the embedding formalism, which determine that the met-
ric of this at space-time manifold must be the diagonal Lorentz metric
(1, 1, 1, +1).
Hence with gss = 1 and g44 = +1 these constructions, based on
the general covariant eld equations from Eqs. (53), yield eld equations
for the F equivalent in form to the Maxwell equations on this four-
dimensional manifold carrying the Lorentz metric, that is,

E
(a) B = 0 (c) B = Js
x 4
B
(b) E+ =0 (d) E = J 4 (70)
x 4
J 4
(e) Js + =0
x 4
766 Collins

with (a) and (b) from Eq. (53a), (c) and (d) from Eq. (53b) and (e) from
Eq. (53c). Also with J 4 as ce , e being the spatial density of electric
charge, and x 4 = ct, J 4 x 4 is e t so Eq. (70e) is the familiar con-
servation of electric charge on the spatial manifold which identies Js as
the electric current density. Certain factors of 4 and 1/c usually appear-
ing in the Maxwell equations, see Jackson,(15) p. 2, are absent in Eqs. (70)
because our vector eld A is the multiple c/(4 ) of the conventional four-
vector potential of the electromagnetic eld in Gaussian units.
Incidentally we also observe that with the Lorentz metric and x 4 = ct
in Eq. (68), this equation for B, and the corresponding equation for E,
have the form of the wave equation for waves of phase velocity c. Thus
this probability formalism for particle mechanics has identied c as the
constant, invariant speed of electromagnetic waves in free space, that is, c
is the speed of light.
All that remains for a complete set of eld equations is the appro-
priate eld equation for the scalar eld 0 which we tentatively identify as
the gravitational potential. Since this eld 0 must satisfy the inhomoge-
neous equation
 in Eq. (58b) and on any at manifold  0 is simply
g 2 0 (x x ), such that in the diagonal Lorentz metric, with elements
(1, 1, 1, +1), x 4 as ct and JG as 4M , this is the inhomogeneous
DAlembertian equation,

1 20
20 = 4M , (71)
c2 t 2

with the Newtonian gravitational constant and M the spatial density


of source mass. This choice of representation for JG yields the familiar
Poisson equation of Newtonian gravitational theory in the non-relativistic
limit, c in Eq. (71).
One might question the physical validity of this identication for 0;
however, a general solution of Eq. (71) for M and 0 independent of is

   
0 (xs , t) = 4M xs , t  G| xs , t : xs , t  d 3 x  dt  (72)

with xs and xs position vectors on the at spatial manifold and G| the
causal form of the innite-domain Greens function for Eq. (71), as in
Mathews and Walker,(16) pp. 278280,
   
t t  xs xs  c
G| =   . (73)
4 x x 
s s
The Mathematical Basis for Physical Laws 767

execution of the t  integration, with this form for G| yields


    
M xs , t xs xs  c
0 (xs , t) =   d 3x (74)
x x  
s s
 
valid for t  xs xs  c. This is identical to the Newtonian form except
for the retarded time dependence of the source mass density M . Therefore
if M has only negligible
 variation
 with coordinate time, or if the time-
retardation interval, xs xs  c, is negligible compared to t, then this 0
eld will be indistinguishable from a Newtonian gravitational eld. Speci-
cally for a spherically symmetric mass distribution of radius r0 and total
mass M s (t) at the spatial origin, it can be shown that 0 M s (t
r/c)/r for r r0 which becomes 0 M s (t)/r for t r/c. Fur-
thermore, for this latter case or for 0 independent of t, 0 is the Newto-
nian form satisfying the Poisson equation for 0. On the other hand, this
retarded time dependence might account for some anomalies in the New-
tonian formalism.
With cII as B, cI as E and G as the four vector 0 we have the equa-
tion of evolution for u from Eq. (57) on this at space-time manifold as
du  
= us B + u4 E us E14 + 0 + Q. (75)
d
Then the three-vector, spatial component of this is
dus  
= us B + u4 E + 0 + Q, (76)
d
with the s subscript denoting spatial three-vectors, while the fourth, or
time component is

du4 0 Q
= us E + 4 + 4 (77)
d x x
since I4 I4 = g 44 = +1 and u4 = g 44 u4 . Thus in addition to the
quantum mechanical force, Q, Eq. (76) includes the forces (per unit
mass) acting
 on the relativistic particle, as the electromagnetic Lorentz
force us B + u4 E , with = e/(mc), and the force 0 which we iden-
tify as the gravitational force with 0 the gravitational potential.
A note on signs of certain terms is in order here; recall from Sec-
tions 2 and 3 that R, v and u are dened by components on the covariant
basis vectors 1 while the gradient operator  is dened by components
on the contravariant basis vectors 1 . Thus doting 1 onto Eq. (75) yields
du /d having g 0/x in the right member because 1 1 = and
768 Collins

1 1 = g . Thus in the Lorentz metric (1, 1, 1, +1) the compo-


nents of 0 appear as 0/x on the spatial axes but as +0/x 4 on
the time axis.
Forming expectation values for Eqs. (70), (71) and (75) under dPx =
x d 4 x, with all terms involving Q(x ) zero, we nd, in the limit of abso-
lute precision with x (x | ) (x xc ( )), these equations provide the
complete description for the deterministic motion of a charged particle in
electromagnetic and gravitational elds in the formalism of special relativ-
ity theory. This formalism is therefore determined solely by mathematics
based on the underlying probability formalism, not any empirical physical
laws. The relationship of this formalism to special relativity theory will be
examined further in Section 8 following but rst we must examine the sig-
nicance of the two time parameters, and t, in more detail.

7. THE ISOLATED PARTICLE

Recall from Section 3 that world time is dened as a non-decreas-


ing, one-parameter label for the conguration of the uncontrolled environ-
ment external to the physical system in a prepared experiment, as per-
ceived in the experimenters frame of reference. Clearly, an increase in
may or may not correspond to change in some uncontrolled parame-
ters which actually affect the particle motion. However, if any of the vec-
tor elds acting on the particle were explicit functions of these would
be dependent on elements of the changing external environment not con-
trolled by the experimenter. We refer to such a system as non-isolated. On
the other hand, if no eld acting on the particle has an explicit -depen-
dence, then this is an isolated system. Indeed, it is isolated from changes
in the uncontrolled external environment as perceived by the experimenter.
Thus, in an isolated system u evolves as a function of the noncausa-
tive parameter , which is merely a marker provided by an external clock.
Since there are two time parameters, and t, with x 4 = ct an observable
in the relativistic formalism corresponding to n = 4, any of the elds, as
say B, may be transient, or non-stationary with B/t non-zero and con-
currently isolated from uncontrolled changes in the external environment
with B/ zero.

8. THE RELATIONSHIP OF THIS FORMALISM TO SPECIAL


RELATIVITY THEORY

Although we have referred to the particle in motion on the four-


dimensional, space-time manifold as a relativistic particle in Section 6
The Mathematical Basis for Physical Laws 769

we did not call upon the epistemological arguments of Einsteins relativ-


ity theory. Indeed, there the Lorentz metric was required solely to assure
bounded solutions to the source-free Maxwell-type equations. This metric
assignment alone justies the designation relativistic particle as appro-
priate. However, we now provide additional support for this terminology
by establishing the relationship of our world time to the proper time of
the special theory of relativity, and supplying further justication for iden-
tifying 0 as the gravitational potential.
The DAlembertian eld equation for 0 in Eq. (71) is not only Lo-
rentz invariant in form but also the general solution in Eq. (74) clearly
indicates propagation of gravitational interaction at the nite speed of
light, c. Within the solar system this nite speed can be signicant only
as a contributor to small perturbations of the conventional Newtonian
description for planetary orbits, but on the larger galactic scale this may
contribute signicant effects. Pais(17) has pointed out that as early as 1911,
Albert Einstein, as well as others, had considered a theory of gravitation
based on this DAlembertian equation but this was abandoned because
Lorentz invariance was not considered sufcient. Indeed, general covari-
ance for arbitrary coordinate substitutions must be required in a proper
theory. This is assured in our constructions by the covariant vector calcu-
lus with the general eld equations in Eqs. (53) and (58) being covariant
for general coordinate substitutions.
For an isolated system the general equation of motion in the limit
of absolute precision from Eq. (35), with all eld quantities having only
implicit dependence on , we have for any metric and manifold dimension
duc
= uc A + 0, (78)
d
with each vector, or multi-vector element in this equation being covariant
for any substitutions, x = x (x  ) on the manifold.
For linear transformations of coordinates, represented by constant
matrices, independent of the coordinates, the four-vector, space-time rep-
resentation is not only covariant but also retains globally-uniform basis
vectors on any at manifold in the transformed frame of reference. This
holds for the equation of motion in Eq. (78) and for the eld equations
when expressed in the four-vector format. In particular for the Lorentz
metric we nd covariance with globally uniform basis vectors preserved
under proper orthochronous Lorentz transformations.(18) It is important
to note that in all constructions presented in this paper all equations are
written in the frame of reference of the observer because these are derived
from the probability distribution dP (x, v | ) which the observer uses to
describe theoretical predictions for physical observations of the particle.
770 Collins

Now forming the dot product of uc onto Eq. (78) yields

duc
uc = 0 uc , (79)
d
since uc A uc is identically zero simply because A is antisymmetric
in the coordinate indices. Hence, multiplying Eq. (79) by d and executing
the indenite integral provides

1 1
uc u c = 0 + (80)
2 2
with the constant of integration /2, independent of , having units of
energy per unit mass. This integration makes use of the fact that uc d is
equivalent to dRc from Eq. (13), with dR dened as in Eq. (1). This pro-
vides (dR/d )d as dRc in the precision limit x (x| ) (x xc ( ))
such that 0 dRc is simply d0 because in an isolated system 0 is an
explicit function of only the x and not .
Note that for any metric Eq. (80) can also be written as

g xc xc 20 = , (81)

with the uc = xc = dxc /d being components of the proper velocity of
the particle in the reference frame of the observer. For the relativistic par-
ticle on the space-time manifold the value of is determined as c2 by the
following considerations.
For the particle on the space-time manifold having the Lorentz metric
(1, 1, 1, +1), we have
 2
dtc
uc uc = (u4c )2 + ucs ucs = (c2 vcs
2
), (82)
d

because vcs vcs = vcs


2 in this metric and u4 is c dt /d , where
c c

dx 4  dt dtc


=c c = u4c (83)
d d d
in the precision limit, x (x, ) (x xc ( )). Similarly, with xs the
position vector on the E 3 spatial manifold, vcs is dened in this limit as
dxcs /dtc by the relationship

dxs  dt dxs  dtc


= vcs = ucs . (84)
d d dt d
The Mathematical Basis for Physical Laws 771

Thus, with Eq. (82) substituted into Eq. (80), we obtain


 2
dtc 2 20
(1 vcs /c2 ) = + 2 (85)
d c2 c

upon multiplication by 2/c2 .


Recalling the denition of world time, , in the observers frame of
reference in Section 3, together with that for the coordinate time, t, in the
frame of reference of the moving particle, as introduced in Section 6, it
2 /c2 0 and (2/c2 )|0| 0, we must assign
is clear that for both vcs
dtc /d 1. That is, for the particle at rest in the potential-free reference
frame of the observer, perceptions of change in the external environment
must be the same in these two reference frames, so an increment in must
be equal to an increment in tc , With dtc /d 1 Eq. (85) then states that
= c2 . Therefore with /c2 as +1 we have from Eq. (85)


dtc 1 + 20 c2
=  (86)
d 1 vcs
2 c2

with only positive signs on the radicals because by denition both t and
are non-decreasing parameters.
Now for 2 |0| /c2 essentially zero compared to unity, Eq. (86) yields
! "
d = dtc 1 vcs
2 c2 , (87)

which is identical to the relationship of proper time and coordinate time in


special relativity. Thus with no gravitational eld present our world time is
actually the conventional proper time of the special theory of relativity, as
described by Bergmann(18) for example. However, in the conventional the-
ory proper time is interpreted as measured in the frame of the moving
particle while coordinate time is measured in a xed frame in which the
particle is seen to have the velocity vcs , but actually this interpretation
is not required by the mathematics on which Eq. (87) is based in conven-
tional relativity theory.
This interpretation of is arrived at in conventional derivations of
Eq. (87) based on the requirement that the speed of light c is the same in
two frames of reference in uniform linear motion relative to each other.(18)
This yields the familiar linear Lorentz transformation relating spatial and
time coordinates in the two reference frames and the quadratic constraint
772 Collins

of equality for the quantity c2  2 = c2 (tc )2 |xcs |2 in the two ref-


erence frames. Here tc and |xcs | are the elapsed coordinate time and
distance between two events as seen in either one of the reference frames.
Thus the observer (experimenter) sees the same increment  in world time
no matter which of the two frames he uses to describe the two events.
It is then argued that if there exists a frame of reference in which the
two events occur at the same spatial location so that |xcs | = 0 then
 = tc . Based on this equality it is the conventional conclusion that
is time measured in the reference frame rigidly attached to a moving par-
ticle, hence the name proper time for . However, as noted just following
Eq. (85) above, the equality  = tc actually occurs when a particle is
at rest in the reference frame of the observer (experimenter). Indeed, it is
tc that is measured in the reference frame attached to a moving particle,
as dened in Section 6, not .
This is made clear by dening two frames of reference by the
space-time coordinates of two particles, as perceived by the observer
(experimenter). Specically these are xcs , tc , and xcs , t  , respectively, with
c
the latter particle at rest in the observers frame of reference, that is vcs  =

0. Observational events are then detection of both particles at world times


and +  . These yield the quadratic constraints  for 2 constant light
velocity in the two reference frames as c2 ( )2 = c2 tc since xcs =0

because vcs = 0, and c ( ) = c (tc ) |xcs | . Then, in the limit
2 2 2 2 2

as  , xcs , tc and tc approach zero, we obtain these equations as


   
d = dtc and d = dtc 1 vcs 2 c2 since lim x
cs tc = vcs is the
velocity of the moving particle relative to the stationary particle, as seen
by the experimenter. Of course this last result is identical to Eq. (87). Then
eliminating d between these two equations yields the familiar equation
  
2 c2 1/2 relating time t  in the stationary frame and t
dtc = dtc 1 vcs c c
in the moving frame of reference. Indeed, every particle in a multi-particle
system has its own coordinate time but there is only one proper time
common to all.
For the particle at rest, vcs = 0, in the reference frame of the observer,
but exposed to the gravitational eld, Eq. (86) provides
 
dtc = 1 + 20/c2 d 1 + 0/c2 d, (88)

with the approximation valid for 2|0|/c2 1. We view this as additional


support for the identication of 0 as the gravitational potential in this
space-time formalism because Eq. (88) accounts for the modied duration
of temporal events in gravitational elds (0 < 0) which predicts the grav-
itational red shift of spectral lines in light from massive stars. Indeed, the
The Mathematical Basis for Physical Laws 773

rst expression in Eq. (88) is equivalent to that deduced in the metric the-
ory of general relativity, valid for small particle velocity and a weak grav-
itational eld, as described by Adler et al.(19) pp. 131137. However, their
roles for tc and are interchanged here because here it is the coordinate
time tc that is dened in the reference frame of the moving particle, not
, as just emphasized above. These authors(19) also described conrmation
of this relationship to an accuracy of better than one percent in experi-
ments based on the Mossbauer effect reported by Pound and Rebka(20)
and Pound and Snider.(21) It is commented that this illustrates how the
gravitational eld, in the form of a non-Lorentzian metric, inuences mat-
ter, (or light) in its vicinity, but here this is based explicitly on the Lo-
rentz metric. Thus the Einstein concept of a gravitationally determined
metric is not required to account for this particular phenomenon. Actually,
Einstein(22) rst predicted a gravitatational red shift in 1911 based on the
Doppler effect and the equivalence of a uniform, linear gravitational eld
gradient and a uniform, linearly accelerated frame of reference.
Einsteins 1911 paper also predicted the deection of light by a grav-
itational eld but this employed a very obtuse argument about the effect
of a gravitational eld on the operation of clocks, which he claimed to
yield a linear relationship between the speed of light at a point and the
potential 0 existing at this point. He then multiplied this speed by dt,
or d in our notation, to provide displacement distances for points of a
wave front corresponding to different values of 0. These were then used in
an analysis based on Huyghens principle and Euclidian geometry to com-
pute the displaced trajectory of a light ray passing near the Sun. However,
these wave front displacements are given in our constructions by simply
multiplying dt c in Eq.
 (88) by the invariant speed of light c, thus yielding
cdtc cd 1 + 0 c2 for the displacement. One can then proceed with
Einsteins construction for the deection of the light ray using Huyghens
principle.
Einsteins 1911
 calculation
 yielded the angular displacement of the ray
as E = 2Ms / c2 rm , with Ms /rm = |0|, Ms the solar mass, New-
tons gravitational constant and rm the radius from the solar center at
nearest approach. This is incorrect by a missing factor of two; the cor-
rected deection, = 2E , predicted by general relativity theory,(19)
has been conrmed by numerous astronomical observations. In 1995 this
was conrmed to very great accuracy using very-long-baseline-interferom-
etry observations of radio waves from the quasar 3C279, at three differ-
ent frequencies, within a few days of it being occulted by the Sun. This
employed observations in Massachusetts and California, Lebach et al.(23)
Thus the deection of light passing near the Sun seems to conrm gravi-
tation as the determining factor for manifold geometry.
774 Collins

9. MANIFOLD CURVATURE AND EINSTEINS GENERAL


THEORY OF RELATIVITY

Actually the general formalism for particle motion developed here


could admit a relationship between gravitation and manifold geometry as
assumed in Einsteins general relativity theory, but this is not necessary and
raises interesting questions about his assumptions. This is shown by writ-
ing our equation of evolution for uc from Eq. (78), for an uncharged par-
ticle, = e/(mc) = 0, in terms of components on the space-time basis
vectors as

d 2 xc dxc dxc 0
2
+ {, } = g a , = 1, 2, 3, 4. (89)
d d d x

This employed duc /d = d(1 uc )/d with d1 /d = (1 /xc ) (dxc /d )
and uc = dxc /d together with 1 /xc = {, } 1 followed by equating
components in left and right members by dotting 1 onto both sides with
1 1 = and 1 1 = g .
Note that the entire left members of Eq. (89) represent the accelera-
tion components, 1 (duc /d ), = 1, 2, 3, 4, so these equations conform
to Einsteins principle of equivalence for gravitational force and accelera-
tion which is independent of the particle mass. The negatives of the terms

in Christoffel symbols {, } uc uc in Eq. (89) are viewed as constraint
forces which conne the particle motion to the curved manifold, while the
terms g 0/x are real gravitational forces which accelerate the particle.
On a at Lorentzian manifold, with globally uniform basis vectors, all of
the {, } are everywhere zero so Eq. (89) can be written with the uc =
dxc /d as

d 2 xc 0
2
= g , = 1, 2, 3, 4. (90)
d x
Thus on this at manifold no constraint forces are required.
In Einsteins general theory of relativity, the gravitational eld merely
determines the manifold metric while the particle moves with zero acceler-
ation along a geodesic. Specically, Einstein(24) assumed the equations of
motion to be

d 2 xc dxc dxc
+ {, } = 0, = 1, 2, 3, 4, (91)
d 2 d d
with the {, } determined by the gravitational eld of material bodies
acting on the particle whose motion is under study. Thus this conicts
with the principle of equivalence.
The Mathematical Basis for Physical Laws 775

These Einstein equations of motion actually dene geodesics on a space-


time manifold which, in the absence of gravity, are simply straight lines on
a curvature-free, at manifold with the Lorentz metric since all {, } are
then zero. It was these facts which led Einstein(24) to generalize this eld-
free case to the general case with Eq. (91) as the most plausible assump-
tion for the motion of a material particle in a gravitational eld with the
= {, } he described as components of the gravitational eld.

However, these equations do not describe the motion of a material particle
but are instead only dening equations for geodesics which have been iden-
tied with trajectories of rays of light. Indeed, this interpretation is used
to derive the general relativity alternative to Einsteins 1911 calculation for
light deection yielding the factor of two correction, see Adler et al.(19) ,
p. 214 Actual calculations for particle motion in a gravitational eld based
on the Einstein theory employ EulerLagrange equations derived from the
alternative denition of a geodesic as the extremum for a path integral of
the line element on the manifold. See Adler et al.(19) , pp. 199213 In 1916
Einstein(24) showed that in the non-relativistic limit, corresponding to a weak
gravitational eld and low particle velocity, Eqs. (91) could be approximated
as equations of motion for a particle with 0 a solution of the Poisson equa-
tion. The validity of Einsteins interpretation of this construction is examined
in Section 11.
Our equations of motion in Eq. (89) provide an alternative to Ein-
steins geodesic theory which does not explicitly require the determination
of the manifold metric by the gravitational eld of other material bodies.
This is compatible with the principle of equivalence as cited above and
yields one result directly identifying this as a possible alternative to the
Einstein theory. Specically we will show that our formalism for particle
motion supplies the factor of two required to correct Einsteins 1911 cal-
culation for the deection of light by the Sun, as described in Section 8
above. To show this we proceed as follows:
First we note that the Christoffel symbols, dened as {, } = 1
1 /x , are readily shown to be related to the metric tensor elements,
dened as g = 1 1 by the equations

 
1 g g g
{, } = g + , (92)
2 x x x

and, since g = g for any Riemannian metric, there always exists a lin-
ear transformation of coordinate differentials which diagonalizes the met-
ric tensor. Thus for a diagonal metric, g = 0, = , with Eq. (92)
for the {, } it is readily veried that the constraint forces in Eq. (89)
776 Collins

are represented by
 
1 g
{, }u
c uc = g

u u (uc g )uc (93)
2 x c c

for = 1, 2, 3, 4. We also note that, from the derivations of Eqs. (79) and
(80) from our equation of motion in Eq. (78), the relationships,
 
1 0
uc uc = , = 1, 2, 3, 4, (94)
x 2 x

are true equalities. Thus, since g uc uc uc uc , Eqs. (93) appear as
 
0 uc
{, }u
c uc = g

g u
c

(u c g )u
c

x x (uc g )uc
(95)

for = 1, 2, 3, 4. We then evaluate these for a particle following the light


ray path at the point of nearest approach to the Sun. We assume the metric
to be that of Schwartzschild in Eq. (108) of Section 11.
Since the light ray is normal to the solar midplane at nearest approach
this plane is selected as that of = /2 in the spherical coordinates r, , .

Thus at this point urc = uc = 0 while uc and u4c are non-zero. Further-
more for the Schwartzschild metric in Eq. (108), g / = 0, g44 / = 0,
g /x 4 = 0 and g44 /x 4 = 0. Thus the last term in Eq. (95) is zero
at this point. Since this is a turning point of the trajectory rst derivatives

of the non-zero uc are zero, so the middle terms of Eq. (95) are also zero.
Indeed, in the neighborhood of this point Eq. (95) is a good approximation
with both middle and last terms deleted. Thus with these terms omitted we
use Eq. (95) to substitute for the constraint forces in Eq. (89) with the metric
assumed diagonal. This yields the rearranged forms

d 2 xc 0
2
= 2g , = 1, 2, 3, 4, (96)
d x
which reveal that at this point on the manifold with curvature the particle
appears to have a at manifold acceleration caused by a gravitational force
twice that seen on a at manifold with only the radial component nonzero.
Furthermore in the Schwartzschild metric g rr = (1 2/r) which for a
weak eld, |/2r| 1, is essentially minus one. Therefore Eq. (96) yields
d 2 rc /d 2 20/r. The above statement assumes the 0/x to be at
least approximately equivalent to those on a at manifold. In Section 11 it
is shown that on a manifold with the Schwartzschild metric the eld 0 is
The Mathematical Basis for Physical Laws 777

described by our Eq. (58) in this metric, which is well approximated by the
DAlembertian equation in Eq. (71) in the weak eld limit. Therefore the
0/x are well approximated by the at manifold values. This factor of
two on the gravity potential 0 is consistent with the factor of two required
to correct Einsteins 1911 calculation of light deection by the Sun. Thus,
if deection of light by gravitation arises because a ray of light is equiva-
lent to a beam of particles of negligible mass, then this is a plausible expla-
nation for the required factor of two for this correction and is equivalent
to that provided by general relativity theory.
With this background for Einsteins general relativity equations of
motion, and their relationship to our mathematical constructions, we now
show how the covariant vector calculus formalism yields eld equations
constraining the manifold metric.

10. SINGLE-VALUED VECTOR FIELDS AND MANIFOLD


CURVATURE

The change in a vector eld corresponding to displacement along a


path on a differentiable manifold is intimately related to the geometry of
the manifold. Specically the change, dA, in any vector eld, A(x), corre-
sponding to a displacement dR = 1 dx along a path dened by paramet-
ric equations x = x () , = 1, 2, . . . , n is given by dA = dRA. Then
the total change in A = 1 A for a nite displacement along this path is
simply the integral of dA along the path, but this integral may be path
dependent. For a closed path the net change, A, in A upon arriving
again at the initial point, is evaluated using Stokes theorem for a vector
eld but with the dyadic A replacing the vector eld; that is,
# # 
(2)
A = dA = dR A = dRs(1) [A] dRS . (97)
S

Clearly A must be zero if A(x) = 1 (x)A (x) is a single-valued vector


eld.
(1) (2)
Here the surface S is bounded by , with dRs and dRs non-colin-
ear displacement vectors, tangent to the surface, jointly dening elements
of area on the two-dimensional surface S. Then executing the operation
, followed by , acting on A yields
 
1 1
A = A 1
1 , (98)
x x x x
778 Collins

because all terms involving derivatives of the A add to zero. Thus this,
(1) (2)
with dRs = 1 dx and dRs = 1 dx , yields Eq. (97) as
# 
A = dA = A R

1 dx dx (99)
S
1 for the vector elds, dened by
with the notation R

1 1

R 1 =
, (100)
x x x x
which characterize curvature of the x-manifold. Indeed, considering any
area element corresponding to dx dx on the coordinate surface x =con-
stant, having local normal vector 1 (x) this reveals that if the vector elds
in Eq. (100) are zero then directions of 1 at diagonal corners of the sur-
face element are parallel so this surface element is at, but if these elds
are not zero then the directions of 1 at these corners are not parallel,
indicating curvature of this x surface. Executing the derivatives of 1 in
Eq. (100) with d1 /dx = {, } 1 we nd that in conventional tensor
notation


R = { , } {, }
x x
+ { , } {, } {, } {, } , (101)

which is recognized as the RiemannChristoffel curvature tensor.



Thus, if R is zero on a surface S bounded by the closed curve , then
for any vector eld A(x) the integral of dA around the path is zero; that is
A is zero. The vanishing of A for all such closed paths on a domain of
the manifold identies A(x) as a single-valued vector eld on the domain.
Thus, in order that such elds be denable on a domain of the manifold,

all R must be zero on that domain.
This RiemannChristoffel tensor for a four-dimensional space-time
manifold was a cornerstone of Einsteins general theory of relativity.(24)
However, the equation for the gravitational eld in free space introduced
by Einstein was not expressed in terms of R but instead in terms of
the tensor of lowerorder called the Ricci tensor, which we construct as

R = 1 1 R = R . This inner-product operation corresponds
to contraction in tensor analysis. Clearly the R will be zero or non-
zero on the same manifold elements as the vector elds 1 R . Indeed

this denition is equivalent to


 
1 1
R = 1 = R (102)
x x x x
The Mathematical Basis for Physical Laws 779

with summation on  the repeated index . Executing the derivatives using


the denition 1 x = {, } 1 for the {, } yields Einsteins general
relativity eld equation in matter-free space as

R = { , } {, } + { , } {, }
x x
{, } { , } = 0. (103)

However, these Einstein eld equations merely assure the existence of sin-
gle-valued vector elds, including the gravitational eld, 0, in matter-free

space. Since all { , } = n g x and n g are zero for g = 1,
Einstein (24) eliminated the rst and last terms of this equation by requir-
ing local coordinates such that g = 1, but here this requirement is not
imposed.
However, when used in conjunction with the Einstein equations of
motion in Eq. (91), these Einstein eld equations do determine the gravita-
tional eld in matter-free space, as dened by Einstein with his extensions
to include eld sources described below. Alternatively, when used with Eq.
(89) these Einstein free-space eld equations merely determine the metric
of the space-time manifold, with this constrained only by symmetry con-
ditions. An example is the Schwartzschild metric described in Section 11.
The gravitational eld 0 on a curved space-time manifold is determined
by our Eq. (58) with JG expressed as 4M , just as in Eq. (71) for the
at Lorentzian manifold. The metric in these equations for 0 is therefore
determined by the Einstein eld equations in Eq. (103) and this same met-
ric must also be applied on this space-time manifold in the electro-mag-
netic eld equations which are exhibited in Eq. (53) for arbitrary manifold
dimension and metric. Thus an alternative explanation for the deection
of light by the Sun may reside in this gravitationally altered metric in
the eld equations of electromagnetism.
Einsteins general relativity eld equations, in domains free of gravita-
tional sources in Eq. (103), are expressed in our covariant vector notation
as the dyadic form

R = 1 1 R = 0. (104)

This can be extended to domains containing gravitational source materials


in the manner used by Einstein.(24,25) Thus a new dyadic eld G is dened
as R plus another covariant dyadic which is also zero on all space-time
domains on which R is zero. Then a dyadic source eld T can be intro-
duced by the relationship

G = kT, (105)
780 Collins

with a constant k which Einstein showed to be proportional to Newtons


gravitational constant.(19) Clearly this reduces to G = 0 and therefore R = 0
on all domains where T = 0. Furthermore if G is dened as

1
G = R 1 1 g R, with R = R g , (106)
2

then not only does T = 0 imply all R = 0 but also this provides the
identity  G 0 which, with Eq. (105) implies the identity  T
0. Indeed, the numerical coefcient, 1/2 in G, was selected to assure
 G 0 by Einstein(25) because with his chosen form for T the iden-
tity  T 0 provides conservation laws for the momentum, energy and
mass of the source material of the gravitational eld.
It was in the design of the tensor T , having divergence zero, that
Einstein called upon empirical conservation laws of physics. In the covar-
iant vector notation this form is

T = M u u , (107)

with M the spatial mass density of the continuous source distribution


and u = 1 u the proper velocity eld of this mass distribution on the
four-dimensional space-time manifold on which the motion of a particle
is described by dP (x, v| ). Einstein also included an additive term, not
shown here, for the contribution of the energy density of the electromag-
netic eld to the gravitational source. This was expressed in terms of the
electromagnetic eld tensor F in a covariant expression. This term also
had divergence equal zero. See Einstein,(24) Section 20. However, this term
is orders of magnitude less than that for massive materials.

11. THE SPECIAL RELATIVITY LIMIT FOR THE GENERAL


THEORY OF RELATIVITY

Apparently the fact that the Christoffel symbols can be expressed in


terms of the metric tensor elements, as in Eq. (92), motivated Schwartzs-
child(26) in 1916 to construct an exact solution to the free-space equa-
tions R = 0 in Eq. (103), for a time-independent spherical source
at the spatial origin, by determining the line element with the constraint
of spherical symmetry for the three-dimensional, spatial submanifold of
the space-time manifold. This solution process made explicit use of the
Einstein geodesic equations in Eq. (91). This is described in detail by
Adler et al.(19) , pp. 185194 where it is also pointed out that Birkhoff(27)
The Mathematical Basis for Physical Laws 781

demonstrated that the constraint of time-independence introduced by


Schwartzschild was not necessary, so this solution is also valid for a time-
dependent spherical source. In spherical coordinates the square of the
Schwartzschild line element is found to be
 
2 (dr)2
2
(ds) = 1 (dx 4 )2 r 2 [(d )2 + sin2 (d)2 ], (108)
r 1 2/r
which contains only the one free parameter characterizing a spherical
source centered at the origin of spatial coordinates.
Note that this metric determination, made by imposing symmetry
constraints in the solution of Eq. (103), R = 0, does not explicitly
dene this as a gravitational eld. Indeed, by this type of solution we may
eliminate Einsteins identication of this eld as a gravitational eld, thus
retaining only our Eqs. (58) dening the gravitational eld 0.
This Schwartzschild metric is diagonal with all g , = 1, 2, 3, 4 dis-
played explicitly in Eq. (108). The determinant of this matrix is then g =
r 4 sin2 . These g can then be used in Eq. (92) to construct all non-
zero Christoffel symbols {, }. These g and {, } are then used to
express the eld equation for the gravitational potential 0 in Eq. (58b) in
the Schwartzschild metric. With JG as 4M this appears as
   
2 2 0 1 20 1 20 2 1 2 0
1 + + 1
r r 2 r 2 2 r 2 sin2 2 r (x 4 )2
 
2 2 0 cos 0
+ 1 + 2 = 4M , (109)
r r r r sin
which retains the one free parameter yet to be specied.
It is quite evident that as 0 the Schwartzschild line element (ds)2
assumes the Lorentzian form (dx 4 )2 + dxs dxs since on the spatial mani-
fold dxs dxs appears as [(dr)2 + r 2 (d )2 + r 2 sin2 (d)2 ] in spherical
coordinates. Furthermore, in the limit as 0 the eld equation for
0, in the Schwartzschild metric, constructed as in Eq. (109), assumes the
inhomogeneous DAlembertian form in Eq. (71) in spherical coordinates.
Thus 0 must satisfy this equation in this special relativity limit. Further-
more the equations of motion in Eq. (89) will reduce to Eq. (90) with 0
a solution of the DAlembertian equation in Eq. (71) since all {, } are
then zero.
We also examine this limit for the Einstein theory as follows: Since
the Einstein equations in Eq. (91) dene geodesics on the curved manifold,
the special relativity limit for the constraint forces in these is simply

{, }xc xc 0, = 1, 2, 3, 4. (110)
782 Collins

Hence the result is Eq. (90) with all 0/x zero. This merely reverses the
process used by Einstein to arrive at Eq. (91), as described in our Section
9. However, this result has other consequences which are exposed with Eq.
(92) for the {, }. With this Eq. (110) are
 
1 g g g
g + xc xc 0, = 1, 2, 3, 4. (111)
2 x x x

Then since xc = (dxc /dtc )(dtc /d ) for = 1, 2, 3, 4 and from Eq. (86)
dtc /d [(1 + 20/c2 )/(1 vcs 2 /c2 )]1/2 , it follows that dt /d 1 for
c

|20/c | 1 and vcs /c 1. Thus since dxc4 /dtc = c and |dxc /dtc | c
2 2 2

for = 1, 2, 3 we nd that for any diagonal metric, g = 0, = , the


limit in Eq. (110) is well approximated by

1 g44 2
{, }xc xc g c 0, = 1, 2, 3, 4, (112)
2 x
for a weak gravitational eld and low to moderate particle velocities.
In particular for spherical symmetry, as described by the Schwartzschild
metric in Eq. (108), g44 = (1 2/r), so this yields

g44 
2
c2 c , = 1, 2, 3, 4, (113)
x x r

and since /r = 0 as r the integral is c2 /r which is equivalent


to the solution 0 = Ms /r of the DAlembertian equation in Eq. (71) if
is identied as Ms /c2 . However, the special relativity limit requires this
to be zero, as in Eq. (110), so must be zero in this limit, not Ms /c2 .
Actually Einstein(24,25) used an analysis similar to that in our Eqs.
(111)(113) to approximate his geodesic equation by an equation of
motion for a particle in the non-relativistic limit. See Section 21 in
Einstein.(24) Indeed, he identied Ms /r as a solution of the Poisson
equation on the spatial manifold rather than as a solution of the inhomo-
geneous DAlembertian equation on the space-time manifold as in our con-
struction. Clearly this is inappropriate because the Einstein equations in
Eq. (91) describe geodesics, not particle motion.
The most signicant factor showing that general relativity theory
must become a special relativity formalism in the weak-eld limit is that
the square of the Schwartzschild line element in Eq. (108) assumes the
four-dimensional Lorentz form as 0, not the Euclidean three-dimen-
sional form. Furthremore, since the Einstein equations in Eq. (91) actu-
ally describe geodesics on the curved manifold, the correct special relativity
limit is simply that obtained with the {, } all set to zero in Eq. (91).
The Mathematical Basis for Physical Laws 783

Thus the identication of in the Schwartzschild metric as a gravitational


parameter as above is not warranted. Indeed, a physical identication for
this parameter has not been established in our alternative formalism.

12. CONCLUSIONS

Based on the fact that continuous variables are not precisely observ-
able we have shown that the motion of a particle on a Riemannian man-
ifold is properly described by a joint probability distribution dP (x, v | )
for the position and velocity of the particle conditioned by a continuous
time parameter as well as other assigned parameters. Then predicted
observed values for all observables associated with the particle, dened as
covariant functions of the x and v , are provided as expectation values
under dP . Based solely on the existence of this dP (x, v ) and the deni-
tion of particle velocity v by the identity dR(x)/d v with R(x) the
position vector of the particle, basic laws of mechanics, quantum mechan-
ics, electromagnetism, gravitation and relativity are derived as mathemat-
ical identities. Specically, by rst limiting the manifold to be at and
four-dimensional, we have obtained mathematical equations having the
same forms as these laws with physical constants occurring in these equa-
tions which must be properly assigned to achieve complete equivalence.
This physical identication process must also be applied to eld quanti-
ties which appear in the formalism based solely on mathematics, as say the
four-vector eld A as the electromagnetic potential. Thus we have shown
that these familiar laws, established by experiment and empirical deduc-
tion, must have these familiar forms based solely on laws of mathematics.
Indeed, even the manifold metric was determined as the Lorentz metric
based on mathematical requirements for this at manifold case.
These constructions have added to our understanding of these laws by
exposing a new role for the gravitational eld in special relativity theory
with a new relationship between proper time and coordinate time which
includes the gravitational eld as well as particle velocity. This exposes an
alternative to conventional interpretations of coordinate time and proper
time and provides an improved understanding of gravitational effects on
temporal events in the formalism of special relativity.
Einsteins identication of manifold geometry as a manifestation of
gravitation is also given a new understanding by these constructions in
the form of two representations for general relativity theory. One of these
is Einsteins original form while the other is based on our equation of
motion for a particle which includes the gravitational force as the gra-
dient of a scalar potential 0. In Section 9 the differences in these two
784 Collins

formalisms are pointed out and this new formalism is shown to provide
an alternative explanation for the deviation of a light ray passing near the
Sun. In this second form for general relativity Einsteins free-space eld
equations do not determine the gravitational eld 0 acting on a material
particle, instead these determine the metric of the space-time manifold on
which particle motion occurs. The scalar gravitational eld 0 then satises
a generalized inhomogeneous DAlembertian-type equation, expressed in
the metric determined by Einsteins free-space eld equations. This is illus-
trated by the example of the Schwartzschild metric for a spherically sym-
metric source at the spatial origin. Before this new theory can be accepted
as a real alternative to the Einstein theory it must be demonstrated that
it correctly predicts precession of planetary perihelions. This will require
a proper physical identication for the parameter in the Schwartzschild
metric. Indeed, much work remains to really evaluate this new theory.
It is also pointed out that this manifold metric must be applied in the
generalized equations for the electromagnetic eld which were written for
an arbitrary manifold dimension and metric in Eq. (53) of the text. Thus
these Einstein free-space eld equations determine the space-time metric
for all elds which may act on a particle. Thus we have a unied eld the-
ory but we must defer development of this to a future study.
Last but not least these constructions expose the intimate relationship
of classical probability theory, classical relativistic mechanics and quan-
tum mechanics by exposing a proper basis for the hydrodynamic repre-
sentation for the Schrodinger equation including the so-called quantum
potential as named by David Bohm.

REFERENCES

1. R. E. Collins, The Continuum and Wave Mechanics, Ph.D. dissertation, Texas A&M
University (1954).
2. R. R. Dedekind, Essays on the Theory of Numbers trans. by W. W. Beman (1901)
(Open Court, La Salle, Illinois, 1924).
3. L. Auslander and R. E. Mackenzie, Introduction to Differentiable Manifolds
(McGraw-Hill, New York, 1963).
4. W. Feller, An Introduction to Probability Theory and Its Applications, Vols. I and II
(Wiley, New York, 1950, 1966).
5. P. R. Halmos, Measure Theory (Van Nostrand, Princeton, NJ, 1950).
6. G. DeRham, Varietes Differentiables, Formes, Courants, Formes Harmonique (Her-
mann, Paris, 1955).
7. W. V. D. Hodge, The Theory and Application of Harmonic Integrals (Cambridge Uni-
versity Press, London, 1963).
8. A. O. Barut, Electrodynamics and Classical Theory of Fields and Particles (McMillan,
New York, 1964).
The Mathematical Basis for Physical Laws 785

9. R. E. Collins, The probabilistic basis for quantum mechanics, submitted to Found.


Phys. April 2005.
10. E. Madelung, Z. Phys. 40, 322 (1926).
11. D. Bohm, Phys. Rev. 85, 166178 (1952); Phys. Rev. 85, 180193 (1952).
12. P. R. Holland, The Quantum Theory of Motion (Cambridge University Press, New
York, 1993).
13. E. C. G. Stuckelberg, Helv. Phys. Acta 14, 23 (1942).
14. J. R. Fanchi, Parameterized Relativistic Quantum Theory (Kluwer Academic, Dordr-
echt, 1993).
15. J. D. Jackson, Classical Electrodynamics, 2nd edn. (Wiley, New York, 1975).
16. J. Mathews and R. L. Walker, Mathematical Methods of Physics, 2nd edn. (Benjamin,
Menlo Park, CA, 1970).
17. A. Pais, Subtle is the Lord, The Science and Life of Albert Einstein (Oxford Univer-
sity Press, New York, 1982).
18. P. G. Bergmann, Introduction to the Theory of Relativity (Dover, New York, 1976).
19. R. Adler, M. Bazin and M. Shiffer, Introduction to General Relativity (McGraw-Hill,
New York, 1975).
20. R. V. Pound and G. A. Rebka, Phys. Rev. Lett. 3, 439441 (1959).
21. R. V. Pound and J. L. Snider, Phys. Rev. Lett. 13, 539540 (1964).
22. A. Einstein, Ann. Phys. (Leipzig) 35, (1911); also in The Principle of Relativity
(Dover, New York, 1952).
23. D. E. Lebach et al., Phys. Rev. Lett. 75, 1439 (1995).
24. A. Einstein, Ann. Phys. (Leipzig) 49, (1916); also in The Principle of Relativity
(Dover, New York, 1952).
25. A. Einstein, The Meaning of Relativity (Princeton University Press, Princeton, NJ,
1950).
26. K. Schwartzschild, Sitzber. Preuss. Akad. Wiss. (Berlin, 1916), pp. 189196.
27. G. Birkhoff, Relativity and Modern Physics (Cambridge University Press, Mass, 1923).

Vous aimerez peut-être aussi