Sei sulla pagina 1di 49

Foundations of Physics, Vol. 35, No.

8, August 2005 (© 2005)


DOI: 10.1007/s10701-005-6443-7

Cartan–Weyl Dirac and Laplacian Operators,


Brownian Motions: The Quantum Potential and Scalar
Curvature, Maxwell’s and Dirac-Hestenes Equations,
and Supersymmetric Systems

Diego L. Rapoport1
Received April 14, 2005

We present the Dirac and Laplacian operators on Clifford bundles over space–
time, associated to metric compatible linear connections of Cartan–Weyl, with
trace-torsion, Q. In the case of nondegenerate metrics, we obtain a theory of
generalized Brownian motions whose drift is the metric conjugate of Q. We give
the constitutive equations for Q. We find that it contains Maxwell’s equations,
characterized by two potentials, an harmonic one which has a zero field (Bohm-
Aharonov potential) and a coexact term that generalizes the Hertz potential of
Maxwell’s equations in Minkowski space.We develop the theory of the Hertz
potential for a general Riemannian manifold. We study the invariant state for
the theory, and determine the decomposition of Q in this state which has an
invariant Born measure. In addition to the logarithmic potential derivative term,
we have the previous Maxwellian potentials normalized by the invariant den-
sity. We characterize the time-evolution irreversibility of the Brownian motions
generated by the Cartan–Weyl laplacians, in terms of these normalized Max-
well’s potentials. We prove the equivalence of the sourceless Maxwell equation
on Minkowski space, and the Dirac-Hestenes equation for a Dirac-Hestenes
spinor field written on Minkowski space provided with a Cartan–Weyl connec-
tion. If Q is characterized by the invariant state of the diffusion process gen-
erated on Euclidean space, then the Maxwell’s potentials appearing in Q can
be seen alternatively as derived from the internal rotational degrees of free-
dom of the Dirac-Hestenes spinor field, yet the equivalence between Maxwell’s
equation and Dirac-Hestenes equations is valid if we have that these poten-
tials have only two components corresponding to the spin-plane. We present Lo-
rentz-invariant diffusion representations for the Cartan–Weyl connections that
sustain the equivalence of these equations, and furthermore, the diffusion of
differential forms along these Brownian motions. We prove that the construc-

1 Maths Department, FIUBA, University of Buenos Aires and DCyT-UNQ, Argentina.


E-mails: dlrapoport@hotmail.com; drapo@unq.edu.ar; draport@fi.uba.ar.

1383
0015-9018/05/0800-1383/0 © 2005 Springer Science+Business Media, Inc.
1384 Rapoport

tion of the relativistic Brownian motion theory for the flat Minkowski met-
ric, follows from the choices of the degenerate Clifford structure and the Oron
and Horwitz relativistic Gaussian, instead of the Euclidean structure and the
orthogonal invariant Gaussian. We further indicate the random Poincaré–Car-
tan invariants of phase-space provided with the canonical symplectic structure.
We introduce the energy-form of the exact terms of Q and derive the relativ-
istic quantum potential from the groundstate representation. We derive the field
equations corresponding to these exact terms from an average on the invariant
state Cartan scalar curvature, and find that the quantum potential can be iden-
tified with 1/12R(g), where R(g) is the metric scalar curvature. We establish a
link between an anisotropic noise tensor and the genesis of a gravitational field
in terms of the generalized Brownian motions. Thus, when we have a nontrivial
curvature, we can identify the quantum nonlocal correlations with the gravita-
tional field. We discuss the relations of this work with the heat kernel approach
in quantum gravity. We finally present for the case of Q restricted to this exact
term a supersymmetric system, in the classical sense due to E.Witten, and
discuss the possible extensions to include the electromagnetic potential terms
of Q.

KEY WORDS: relativistic Brownian motions; Riemann–Cartan–Weyl con-


nections; Maxwell’s equations; Dirac-Hestenes equations; time-ireversibility;
covariant Schroedinger equations; time-evolution parameter; spinor fields;
quantum gravity; supersymmetric systems; Clifford bundles; generalized Dirac
operators; Quantum potential.

1. INTRODUCTION

In the accompanying article,(1) we have seen that torsion fields are quite
notorious, as they appear as related to the drift vectorfields of generalized
Brownian motions, and we have elaborated as an example, the solutions
of the non-linear dynamics of viscous fluids obeying the Navier–Stokes
equations, or alternatively, the equations of passive transport of magnetic
fields on fluids.(2,3) In this article, we shall give a more embracing theory,
including as examples of Riemann–Cartan–Weyl connections with Car-
tan–Weyl trace-torsion 1-form, electrodynamics, the Dirac-Hestenes equa-
tions, the heat-kernel method of quantum gravity for a spinless mas-
sive particle, and finally, supersymmetric systems. The key to this the-
ory, is that one should consider the Laplacian operators on differen-
tial forms defined by these connections, as the fundamental Hamiltonian
operators of the theory, in contrast with the Laplace–Beltrami opera-
tor which correspond to non-interaction representations. In the case of
Navier–Stokes equations we have seen that the trace-torsion is precisely
the fluid’s velocity and that the Navier–Stokes operator is nothing but an
example of these generalized Laplacians; in the present article, we shall see
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1385

how the electromagnetic potentials of Maxwell’s electrodynamics, as well


as the wave functions for Schroedinger and Dirac-Hestenes equations orig-
inate terms of the Cartan–Weyl form. In contrast with the previous article,
and in view of dealing with the equations of electrodynamics and quan-
tum mechanics, whose most fundamental formulation demand the intro-
duction of Clifford algebras and Clifford bundles over space–time,(4–8) we
shall present these structures and the ensuing generalized Dirac operators
as the basis for the formulation of this theory, which since we shall gener-
ate the RCW laplacians (at least at the level of scalar fields) as the square
of these Dirac operators.

2. TORSION GEOMETRY AND DIRAC OPERATORS

Let M be a smooth oriented n-manifold M, without boundary, g a


smooth metric on T M, the tangent bundle to M, and ∇ a linear connec-
tion on M compatible with g, i.e. ∇g = 0, in short, a triplet (M, g, ∇).
Consider the Clifford bundle on M,(6) defined as the vector bundle defined
by the quotient

Cl (M, g) := T (M)/Jg −1 = Cl (Tx∗ M, gx−1 ), (1)
x∈M

where T (M) denotes the tensor bundle of M, g −1 the induced metric


on T ∗ M by g, and Jg −1 is the ideal on T (M) generated by elements
of the form u ⊗ v + v ⊗ u − 2g −1 (u, v), with u, v ∈ sec(T ∗ M) ⊂
T (M) and Cl (Tx∗ M, gx−1 ) is the Clifford algebra of the metric vector space
(Tx∗ M, gx−1 ), for any x ∈ M, where g −1 is the adjoint metric to g on 1-
forms. Thus, for any two 1-forms u, v, their Clifford product (written as a
justaposition) is defined by

uv + vu = 2g −1 (u, v). (2)

Thus, given a set of orthonormal vector fields {γµ , µ = 0, . . . , n}, with


µ
their dual 1-forms γ µ ∈ secCl (M) so that γ µ (γν ) = δν , their Clifford
product is given by

γµ γν + γν γµ = 2g(γµ , γν ) ≡ 2gµν . (3)

We can further introduce a decomposition of the Clifford product by uv =


u ∧ v + u.v for any 1-forms u, v, where u ∧ v = 1/2(uv − vu) and u.v =
1386 Rapoport

1/2(uv + vu) into the exterior and interior products, respectively. Thus, the
relation in Eq. (3) can be rewritten as

1 µ ν
γ µ .γ ν = (γ γ + γ ν γ µ ) = g µν . (4)
2
For any 1-form u and r-form Ar we define u.Ar := 1/2(uAr −(−1)r Ar a) =
−(−1)r Ar .u and u ∧ Ar := 1/2(uAr + (−1)r Ar u) = (−1)r Ar ∧ u, and con-
sequently, uAr = u.Ar + u ∧ Ar .
We shall present some important examples. Henceforth we shall
denote the Clifford algebras defined by a metric g on n-dimensional space
as Rp,q , where the signature of g is n = p + q, p ones, q minus ones.
Let M = R n be the flat Euclidean space provided with the Euclidean
metric g = diag(1, 1, 1, 1), thus n = 4 = p, q = 0. If n = 3, the
associated Clifford algebra R3,0 is C(2) = End(R 3 ), the Pauli algebra of
2 × 2 complex matrices. The Clifford algebra R3,1 defined by the metric
ν = diag(−1, 1, 1, 1) given by the 4×4 real matrices; if instead we consider
ν = diag(1, −1, −1, −1) so that we have R1,3 corresponding to H(2), the
2 × 2 quaternionic matrices, the so-called Dirac space–time algebra. Both
are of interest when dealing with electrodynamics and quantum mechan-
ics, and in spite of their inequivalence, one can transform objects in one
case to the other. For example, when n = 4 we have the Clifford Euclidean
structure defined by the Euclidean metric diag(1, 1, 1, 1) and we can take
for our original γ ’s the set of orthonormal basis vectors {eµ , µ = 0, 1, 2, 3}
such that eµ .eν = δµν , and denote

γ0 := e0 , γi := ei e0 , i = 1, 2, 3. (5)

Thus, it is easy to check that the set {γµ , µ = 0, . . . 3} form a basis of R1,3 .
If we define instead

γ̃3 := e0 e3 , γ̃i := ei , i = 0, 1, 2, (6)

then, {γ̃µ : µ = 0, . . . , 3} define an orthonormal basis of R3,1 . Thus , it


follows that one can transform from an Euclidean structure to a degener-
ate one, and viceversa, as it can be easily seen. To stress the obvious, the
definition of the Clifford structure is attached to the vectorfields in terms
of which we define a particular structure, yet we can transform them ap-
propiately and still have a unified setting for both the degenerate and non-
degenerate cases, both of fundamental importance.
It is well known(6,9–11) that Cl (M, g) = PSO+ (p,q) ×Ad Rp,q ,
where PSO+ (p,q) is the principal bundle of orthonormal frames and Ad:
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1387

SO+ (p, q) → Aut(Rp,q ) is the adjoint representation of SO+ (p, q) on the


Clifford algebra Rp,q defined by u → Adu , with Adu (x) = uxu−1 , ∀x ∈
Rp,q , u ∈ SO+ (p, q), where p + q = n. The elements of sec(Cl (M.g))
are called multiforms(4,11) , or Clifford fields,(6) or still polyforms.(5) Given
u ∈ sec(T M) and ũ ∈ sec(T ∗ M) ⊂ Cl (M, g) we can consider the tensorial
mapping A → ũ∇u A, for A ∈ Cl (M, g). Since ∇g = 0, then ∇u Jg −1 ⊂ Jg −1 ,
defining thus a covariant derivative operator, ∇, on sec(Cl (M, g)). The Di-
rac operator associated to (M, g, ∇) is (following Ref. 9)

D∇ := Tr(ũ∇u ). (7)

Thus if (γ α (x)) ∈ sec(T ∗ M) ⊂ Cl (M, g) is a coframe on T ∗ M such that


its dual frame is (γα (x)) ∈ sec(T M), i.e. γ α (γβ ) = δβα we have

D∇ = γ α ∇γα , (8)

where we have denoted as before by juxtaposition the Clifford product


between γ α and ∇γα . Thus we have defined

D∇ A = γ α ∧ (∇γα A) + γ α .(∇γα A), ∀A ∈ sec(Cl (M, g)), (9)

which arises from the usual decomposition of the Clifford product defined
by AB = A ∧ B + A.B for any A, B ∈ sec(Cl (M, g)), A a 1-form, into the
exterior and interior products, respectively. Thus,

D∇ ∧ A = γ α ∧ (∇γα A), D∇ .A = γ α .(∇γα A) ⇒ D∇ = D∇ . + D∇ ∧ . (10)

The operator D∇ ∧ satisfies for every A, B ∈ sec(Cl (M, g))

D∇ ∧ (A ∧ B) = (D∇ ∧ A) ∧ B + A∗ ∧ (D∇ ∧ B), (11)


n ∗
where if B = r=0 B r with n B r ∈r sec( (T M)) ⊂ sec(Cl (M, g)),
r

then B =∗ n ∗
r=0 B r = r=0 (−1) B r ., resulting from the applica-
tion of the main automorphism ∗ : Cl (M, g) → Cl (M, g) given by
(AB)∗ = A∗ B ∗ . Another automorphism of Cl (M, g) is defined by B̃ =
n r(r−1)
r=0 (−1)
2 B r , called the reversion operator, and B̃ is the reverse of
B. We shall introduce next the Hodge star operator, i.e. the mapping

∗ : sec(k T ∗ M) → sec(n−k T ∗ M), Ak ; ∗Ak , (12)

where for each Ak , Bk ∈ sec(k (T ∗ M)) ⊂ sec(Cl (M, g)),

(Ak , Bk )volg = Ak ∧ ∗Bk , (13)


1388 Rapoport

So that we note that the Hodge star operator depends on the metric g.
Then, we have the important relations(12) :

∗(∗Ak ) = (−1)k(n−k) Ak , if g is Riemannian, (14)


∗(∗Ak ) = −(−1)k(n−k) , if g is Lorentzian. (15)

In the case when n = 4, introduce a set of orthonormal basis {γµ : µ =


0, . . . , 3}, and define γ 5 = γ 0 γ 1 γ 2 γ 3 ; then we can prove that ∗Ak = γ 5 Ãk ,
where Ãk is the reverse of the k-form Ak . Note that for the Minkowski
space γ 5 γ 5 = −1, so that there is an alternative way of introducing a
square root of −1, which bypasses the complex numbers, and has a geo-
metrical meaning (an oriented volume form) which led to the pioneering
studies by Hestenes on the Dirac equation and its geometrical interpreta-
tion(7,8) .
Let us introduce two operators that will be fundamental to the for-
mulation of this theory. First, the exterior differential d : sec((T ∗ M)) →
sec((T ∗ M)), defined inductively by the relations

d(A + B) = dA + dB, d(A ∧ B) = dA ∧ B + A∗ ∧ dB,


df (u) = u(f ), d2 = 0 (16)

for any u ∈ sec(T M), f : M → R a smooth function. Now we intro-


duce the codifferential operator, δ : sec(k (T ∗ M)) → sec(k−1 (T ∗ M)),
defined as the formal adjoint operator of d with respect to the inter-
nal
 product defined for Ak , Bk ∈ sec(k (T ∗ M)) , then ((Ak , Bk )) :=
−1
⊗k g (Ak , Bk )volg , with volg = |det(g)|dx 1 ∧ · · · ∧ dx n , the metric-vol-
ume n-form and ⊗k denotes the k-th tensor product; it can be defined in
terms of the Hodge star operator: volg = ∗1.(12) Then, since d 2 = 0, then
δ 2 = 0. This internal product is defined for both degenerate (Lorentzian)
metrics, or the non-degenerate (Riemannian) metrics, as long as the inter-
nal product is bounded (see Frankel(12) ). Thus, M will be compact, or
either Ak or Bk will be compact supported. In the Riemannian case, when-
ever ((Ak , Ak )) < ∞, they form a pre-Hilbert space of square integrable
k-forms, which has to be completed in order to obtain a Hilbert space.
Whether M is compact or not, and whether or not M is has a boundary,
if Ak ∈ sec(k (T ∗ M)) ⊂ secCl (M, g), then

δAk = (−1)n(k+1)+1 ∗ d ∗ Ak , if g is Riemannian, (17)


δAk = (−1)n(k+1) ∗ d ∗ Ak , if g is Lorentzian. (18)
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1389

2.1. Dirac and Laplacian Operators associated to Linear Connections of


Riemann–Cartan–Weyl

As an example, let ∇ be the Levi–Civita torsionless connection, ∇ g ,


defined by g, thus

D∇ ∧ = d, D∇ . = −δ ⇒ D∇ = (d − δ)
g g g
(19)

is a generalization of the well known Dirac–Kahler operator on Minkowski


space. Now we can introduce the Hodge laplacian operator

 := (D∇ )2 = (d − δ)2 = −(dδ + δd) ⇒ k := (D∇ )2 |sec(k (T ∗ M))) . (20)


g g

This laplacian contains metric curvature terms (Weitzenbock’s for-


mula.(2,12,13) We shall extend finally the previous construction to more gen-
eral laplacians associated to metric compatible linear connections, as the
square of more general Dirac operators. Let ∇ be a g-compatible connec-
tion, with Christoffel coefficients defined by

∇γµ γ ν = −να
µ α
γ , (21)

which has a non-zero torsion tensor Tµν α := [ α −  α ], and let us com-


µν νµ
pute the laplacian on scalar fields f : M → R defined by (see Eqs. (9) and
(10))

(D∇ )2 f = (γ µ ∇γµ )(γν ∇γν )f = γ ν .γ µ ∇γµ ∇γν f


1
+ γ µ ∧ γ ν (∇γµ ∇γν − ∇γν ∇γµ )f
2
1
= g ∇γµ ∇γν f − γ µ ∧ γ ν Tµν
µν α
γα f. (22)
2
The second term of the right hand side of Eq. (22) can be further reduced

1 µ 1 1
γ ∧ γ ν (γα )Tµν
α
= (γ µ (γα )γ ν − γ ν (γα )γ µ )Tµν
α
= (δαµ γ ν − δαν γ µ )Tµν
α
2 2 2
1 µ ν
= (Tµν γ − Tµα γ ) = Qν γ ν = Qν γν ,
α µ
(23)
2
µ
where Tµα := Qα , are the components of the trace-torsion 1-form (which
we shall henceforth call the Cartan–Weyl 1-form), and we have used that
Qν γ ν = Qν g αν γα = Qα γα , which is the vectorfield given by the g-conju-
gate of Q = Qα γ α ; we denoted it Q̂ in our accompanying article.(1) The
1390 Rapoport

first term of the right hand side of Eq. (22) is the same laplacian on real
valued scalar fields defined on M, f , which we computed in(1) (see also
Refs. 2, 3 and 14),

tr∇ 2 f = g αβ ∇α ∇β f = g f + 2Q̂, (24)

which in this Clifford bundle setting arises from the inner product of the
decomposition of the Clifford product. The first term of Eq. (22) is very
important, since depending on the Clifford product we associate to the
{γµ }’s, if degenerate associated to a Lorentzian metric (viz. Minkowski)
then the corresponding laplacian (D∇ )2 turns to be the hyperbolic d’Alem-
bert operator in Minkowski space, or elliptic in the case of a Riemannian
metric (viz. Euclidean). We further note that in view of this one could
instead of performing an analytical continuation in the t-variable of a
space–time M in order to pass from a Riemannian to a Lorentzian met-
ric, or viceversa (which otherwise, from the perspective of non-conver-
gent Feynman integrals transformed to Gaussian integrals, is mandatory
in statistical mechanics, quantum gravity,(15) etc. ), we could transform the
basic orthonormal fields as in Eqs. (5) and (6) or its inverses, with the
same effect, yet keeping in mind that in the Riemannian case we have an
elliptic theory which gives rise to diffusion processes, and for which the
interaction term is represented by Q̂, up to a factor. Nevertheless, in
the degenerate case of Minkowski space, we shall present further below
the construction of a Lorentz-invariant diffusion process.
Altogether we have the laplacian

1 ∇ 2 1 1
(D ) f = g f + Q̂(f ) with Q̂(f ) = g −1 (Q, df ). (25)
2 2 2

We shall keep this 1/2 factor on Q̂ which does not appear unless we use
the Clifford product, and consider similarly to our previous article(1) (in
which this 1/2 factor on Q does not appear)

1
H0 (g, Q)f = (g f + Q̂(f )), (26)
2

which we shall extend as in our previous article to consider the laplacian


on k-forms on M defined as

1
Hk (g, Q) := (k + LQ̂ )|sec(k (T ∗ M)) , 0 ≤ k ≤ n, (27)
2
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1391

where LX is the Lie derivative with respect to the vectorfield X defined on


M,(16) so that we shall denote in the following the laplacian on multiforms

1  n  n
H (g, Q) = ( + LQ̂ ) = Hk (g, Q), ∀A = A k
2
k=0 k=0

n
∈ sec(k (T ∗ M)) ⇒ H (g, Q)A := Hk (g, Q) A k , (28)
k=0

which depends on g, the metric curvature terms through the Weitzenbock


term in the Hodge laplacian , and still an interaction term defined by
Q̂; we note that by definition, H (g, Q) preserves graduation. To resume,
given a triplet (M, g, ∇), with ∇ a linear connection of Cartan such that
∇g = 0 with arbitrary torsion tensor, we have introduced the Dirac opera-
tor on scalar fields and the associated laplacians, first on scalar fields, and
then extended to k-forms.1 Furthermore, we have seen that only the trace-
torsion 1-form appears in the laplacian, so we shall restrict ourselves to
this case, i.e. to Riemann–Cartan–Weyl connections (RCW, for short), with
Christoffel coefficients given by
   
α 2
α
βγ = + δβα Qγ − gβγ Qα , n = 1. (29)
βγ (n − 1)

1 The relation between (D ∇ )2 |sec(k (T ∗ M)) and Hk (g, Q) requires further investigation since
(D ∇ )2 does not preserve graduation; indeed (D∇ )2 = (D∇ .D∇ . + D∇ ∧ D∇ ∧ +D∇ .D∇ ∧
+D∇ ∧ D∇ .), thus the first two terms do not preserve the graduation on acting on mul-
tiforms (they lower it or upgrade it by two, respectively), and thus we cannot regard it
as a laplacian in the usual sense, and in this article we shall disregard them. This is still
related to supersymmetry, as we shall see in this article. As a final remark, the present
approach in terms of the Dirac operator and the d’Alembertian, in the hyperbolic case,
has been analyzed for the Q ≡ 0 and the Levi–Civita covariant derivative case by Pav-
sic,(5) and shown to resolve the long-pending problem of ordering ambiguity of quan-
tum field theory in curved spaces. Since, this is a problem on the meaning of the square
of momentum operator p(f ) := −iγ µ ∂µ (f ) = −iD∇ (f ), so that p2 (f ) = (−i)2 (D∇ )2 (f )
which we computed above, and similarly we can compute p3 (f ) and so on. (By the way,
the square root of minus one should be replaced by γ 5 , which was in the first place the
reason to introduce Clifford algebras in quantum mechanics.(4) ). Thus Pavsic’s results can
be in principle extended to our more general case. We should note furthermore that the
metric-scalar curvature does not appear in our laplacian for scalar fields (metric curvature
terms appear for differential forms through the Weitzenbock’s term), but further below
we shall see how this term does appear when discussing the quantization associated to
this laplacian. The main issue is what is the density one uses for quantization. Related to
this choice, it is interesting to study further whether the expectation value of the momen-
tum operator still follows the geodesic flow (as proved in Ref. 5 for the restricted theory)
or the autoparallel flow associated to the RCW connection.
1392 Rapoport

The last term in Eq. (29) accounts for the metric compatibility, and
the first term designate the Levi-Civita Christoffel coefficients defined by
g (12) . In distinction with the first ever gauge theory due to Weyl,(17) we
no longer have that parallel transport of lengths and angles is not pre-
served, which lead to the “historicity” argument of Einstein rejecting this
approach. We must stress the fact that in distinction with the Poincaré
theory of gravitation, in which torsion is introduced to account for a
non-propagating spin-tensor associated to a completely skew-symmetric
torsion,(18) here it is the trace-torsion that is considered. In our previous
accompanying article,(1) we have seen that in the Riemannian metric case,
it leads to Brownian motion and particularly, to fluid and magnetofluid-
dynamics, in which the Cartan–Weyl 1-form is the velocity 1-form satis-
fying the invariant Navier–Stokes equations for a non-relativistic incom-
pressible viscous fluid.(2,3) Thus, we claimed that torsion fields, contrarily
to common belief, are certainly abundant in Nature. Remarkably, RCW
connections lead to the introduction of spinor fields, which explains a for-
tiori, our approach in the present article through the Clifford structures.
For introducing this, we start by giving the constitutive equations for Q,
and then shall study some of the many theories related to it.

3. DE RHAM–HODGE DECOMPOSITION OF TORSION, THE


HERTZ POTENTIAL, MAXWELL’S EQUATION AND THE
DIRAC-HESTENES EQUATION

To obtain the most general form of the RCW laplacian in the non-
degenerate case, we only need to know the most general decomposition
of 1-forms. The solution of this problem, valid as well for differential
forms of arbitrary degree is given by the de Rham–Kodaira–Hodge the-
orem. So the present approach stands at the crossroads of very impor-
tant subjects, since it has to do with ellipticity theory of p.d.e’s, de Rham’s
cohomology of differential forms, and fundamentally the analytical treat-
ment of discontinuities of fields (say, Dirac’s delta), since this theory as
stated by de Rham, stems from the theory of generalized functions on
Riemannian manifolds, following L. Schwartz’s theory of distributions.(12)
To start with, in this section, we have a smooth orientable n-manifold M
provided with a Riemannian metric g. We consider as above, the Hilbert
space given by the completion of the pre-Hilbert space of square-integra-
ble smooth differential forms of degree k (0 ≤ k ≤ n) on M, with respect to
the Riemannian volume volg , which we denote as L2 (sec(k (T ∗ M)). We
shall focus on the decomposition of 1-forms, so let ω ∈ L2 (sec(T ∗ M)); then
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1393

we have the Hilbert space decomposition

ω = df + Acoex + Aharm , (30)

where f is a smooth real valued function on M, Acoex is a smooth coexact


1-form, i.e. there exists a smooth 2-form, β2 such that δβ2 = Acoex , so that
Acoex is coclosed, i.e.

δAcoex = δ(δβ2 ) = 0 (31)

and Aharm is a closed and coclosed smooth 1-form,

δAharm = 0, dAharm = 0 (32)

or equivalently, Aharm is harmonic, i.e.

1 Aharm ≡ trace(∇ g )2 Aharm − Rβα (g)(Aharm )α γ α = 0 (33)

µβ
with Rβα (g) = Rµα (g) the Ricci metric curvature tensor. Eq. (33) is the
sourceless Maxwell–de Rham equation (cf. page 568, Misner et al. and
page 175 of Eddington).(19) An extremely important fact is that this is a
Hilbert space decomposition, so that it has unique terms, which are fur-
thermore orthogonal in Hilbert space, i.e.

((df, Acoex )) = 0, ((df, Aharm )) = 0, ((Acoex , Aharm )) = 0, (34)

so that the decomposition of 1-forms (as we said before, this is also valid
for k-forms, with the difference that f is a k − 1-form, β2 is really a k + 1-
form and Aharm is a k-form) has unique terms, and a fortiori, this is also
valid for the Cartan–Weyl 1-form. We shall see next that Acoex and Aharm
are further linked with Maxwell’s equations, both for Riemannian and
Lorentzian metrics, still leading to the equivalence of the Maxwell’ equa-
tion and the relativistic quantum mechanics equation of Dirac-Hestenes.
Comments: A number of comments are in order. The appearance of the
Lorentz group and the Minkowski space–time in the formulation of Max-
well’s equations on the vacuum requires more reflection than the customary
attitude of taking them for granted,2 since these equations can be framed as
being purely topological, and furthermore related to de Rham’s cohomol-
ogy of differential forms.(20) If we write on a space–time M with coordinates

2 That Maxwell’s theory can be framed as a Galilean-group invariant theory has been known
for a long time to be valid.(21)
1394 Rapoport

(x, t), the electromagnetic potential 1-form A = Ak dx k −φdt then we get for
F = dA the Maxwell–Faraday p.d.e’s, while from δF = j , with j the electric
current source 1-form, we get the Maxwell–Ampere equations. So whatever
is the signature of the metric, the equations can be written in a completely
invariant form.3 While we can add to β2 a 2-form that is coexact and still
Acoex = δβ2 , while to Aharm we can add still either the differential of an
harmonic function or the codifferential of a 2-form, and still get the same
Aharm , they are unique as terms of the decomposition. This lack of unique-
ness of the expression of each term, is the starting point of gauge theories,
and can affect as well the topological properties of the solutions of the Max-
well equations as stressed by Kiehn in his topological approach through
Pfaffian systems and their topological dimensions.(20) The non-uniqueness
can appear as discontinuities in solution amplitude and its derivatives (i.e.
electromagnetic signals), as a solution multivaluedness (i.e. polarization), as
envelope (Huygens wavelets or Cherenkov) solutions, and many other topo-
logical physically meaningful properties. Fock(24) demonstrated in 1932 that
the point set of M on which the Maxwell system are not uniquely defined,
defines a propagating discontinuity which can be defined in terms of the

3 In the theory of Maxwell’s equations (ME, for short) generally we have to establish con-
stitutive equations; (Note added in proofs: it is important to remark that these equations
introduce the Hodge duality operator in ME; See Hehl and Obukhov, ref. 20.) Thus the
vectors of dielectric displacement D and of magnetic induction B are introduced; the con-
stitutive equations in the simplest case may take the linear form D = E and B = µH . A
formal distinction of the pair of electric and magnetic vectors E, H with respect to D, B is
desirable if one considers the use of coherent units in ME (instead of using together both
electromagnetic and electrostatic units, see Ref. 22), so that they turn to be free of any
parameters or constants which may be linked with the properties of any particular media.
Thus, this distinction is nothing else than the basis to establish ME in the form of general
covariance, or otherwise stated, in an invariant geometrical setting, as we are presenting in
this article. In this regard ME on the vacuum are the embodiment of both field equations
and constitutive equations, albeit trivial: E = D, H = B, and the theory is relativistic
as well known. Yet, if we consider the case of ME on the vacuum or on an isotropic
nonconducting space with no constitutive equations, the geometrically invariant theory for
ME is formulated in Euclidean space R 4 = {(x 0 = −ict, x 1 , x 2 , x 3 )} with the usual analyti-
cal continuation on the Minkowski space with coordinates (ct, x 1 , x 2 , x 3 ) and the Lorentz
invariance gives place to the orthogonal-group invariance; see the excellent monograph by
Post,(22) pages 54, 55. (It is to be remarked that in the former case, the electromagnetic
2-form has purely imaginary components F0j = iEj , j = 1, 2, 3, F12 = H3 , F13 = −H2 and

F23 = H1 , with i = −1; in the theory of the Hertz potential on Minkowski space–time,(23)
i is replaced by the 4-form γ5 of R1,3 ). Therefore the present formulation of ME on Rie-
mannian manifolds is an extension of the application of the principle of general covariance
to ME on flat Euclidean space with no constitutive equations, while the formulation on a
general Lorentzian manifold is an extension of the principle of general covariance to ME
on Minkowski space with the trivial constitutive equations. The diffeomorphism invariance
of Maxwell´s theory, was first proved – in our understanding – by Post.(22)
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1395

solutions to the eikonal equations, viz. ν −1 (dφ, dφ) = 0, where as above,


ν is the Minkowski metric written, of course, with either signature. So the
Maxwell system perse allows for propagation of discontinuities with velocity
higher than the velocity of light (a problem unconnected with singularities
that has been thoroughly analyzed by numerous researchers, particularly Ro-
drigues and coworkers in connection with the Hertz potential in Minkowski
space; for this and a list of references see Ref. 23). We quote Kiehn’s(20) “it is
the singular ‘characteristic´ solutions of propagating discontinuities (a topo-
logical defect) that ensures the importance of the Lorentz transformations.
Fock also demonstrated that the non-linear fractional Moebius transforma-
tion (i.e. the conformal tranformation) also preserves the eikonal equation.
The projective mapping permits the propagation of the discontinuity to be
less than or greater than c.” Thus, the field dφ which is metric independant
(d is independant of the metric) is an isotropic null 1-form in the Minkowski
metric, and thus, it is a spinor in the sense of Cartan.(25) So spinors play
a role through topological defects both in the quantum realm, and in all
scales.4 A second major topic for discussion, the appearance of the Max-
well equations associated to a term of the trace torsion, which we recall
that it is a U (1) gauge theory in which the electromagnetic potential 1-
forms are real, does seem a far reach from what originally the introduction
of torsion was aimed to do, i.e., the gauging of the group of translations
in R 4 as a semidirect summand to the Lorentz (or orthogonal group) to
yield the Poincaré group, to account for a spin-tensor.(18) Yet, the reduction
of a Cartan connection to gauge this group to a Riemann–Cartan–Weyl
connection, like Weyl’s original gauge theory implemented to unify electro-
magnetism to gravitation, is warranted from the action of the group of the
scalars on both subgroups of the Poincaré group (and in spinors(26) ), i.e. the
Einstein’s λ transformations in the present theory(27) (which contain Weyl’s
scale transformations as a particular case), as we shall see below when study-
ing the problem of conformal invariance and the quantum potential. As a
final remark, the de Rham–Hodge decomposition is valid as well in fluid-
dynamics, where it is the well known Helmholtz decomposition. There, as
we already saw in Ref. 1, the trace-torsion 1-form, uτ is the fluid’s velocity
at time τ and the field intensity is the vorticity, τ = duτ , where d is the
exterior differential with respect to the spatial coordinates of uτ . We can
recover uτ by the Biot–Savart law of electrodynamics, in which the source is
−δτ . This can still be related (as we shall see in a forthcoming article) to

4 In a forthcoming article by this author, a detailed study of the role of discontinuities with
regard to the de Rham decomposition of the Cartan–Weyl 1-form, the eikonal equation
and the description of the photon, the neutrino, etc., in terms of singular sets and both
twistors and spinors will be presented.
1396 Rapoport

the derivation of the Maxwell system from the Euler equations for perfect
fluids, or still, the extension of this system for the viscous case (biquadratic
terms appear there) due to Marmanis,(28) and the subsequent formulation of
turbulence in fluids as a Maxwellian system. Thus, a Galilean-group invari-
ance (local or global) is present both in the Maxwell system and in fluid-
dynamics which perse also leads to the Maxwell system and its extensions,
with a phenomenological velocity of light and discontinuities which may
propagate faster that this velocity.5 The de Rham–Hodge decomposition
although leads to linear constitutive equations, at least for the electromag-
netic terms, is the basis for a linear field theory of non-linear non-equilib-
rium thermodynamics, carried out by this author,(33,34) and by B.Lavenda
in phase-space,(35) in which he invokes Maxwell’s work in electrodynamics
(cf. Sec. 8.3 Ref. 35) as the basis for the de Rham–Hodge decomposition of
the forces and the thermodynamical potentials. Sommerfeld(36) elaborated
as a remarkable curiosity MacCullagh’s work in 1839, further elaborated by
Lord Kelvin,(36) in which the aether is treated as a quasi elastic body, and
derives the Maxwell system from it. Having made these comments, in the
next sections we shall present the formulation of Maxwell’s system which we
shall reduce to a single equation, first on the Clifford bundle Cl (M, g), with
M provided with a Riemannian metric (which as we discussed above, is as
admissible as a Lorentzian metric) for the coexact term (the harmonic term
has a zero field), and second, in Cl (M, ν), with ν the Minkowski metric, in

5 Brownian motion theory, is particularly amenable to the concept of a static world, or


equivalently, of instantaneous correlations. Indeed, Wiener processes, which is the random
element of this theory, have nowhere differentiable paths, so that one can think of infinite
velocity motions. The introduction of the trace-torsion, is precisely the representation of
interactions, with finite average velocity, in the time-evolution parameter of diffusions, τ ,
which is not to be confused with the time variable of M. Yet, as has been stressed by
the author, [2, Rep. Math. Phys. 50, 2002], the RCW laplacian and the Brownian pro-
cesses defined by them, admit a representation in terms of a universal connection built in
Euclidean space, in which the Laplacian and the stochastic differential equations have in
their noise term incorporated the interaction term; thus the present theory has equivalent
static representation, in which, say a viscous fluid which does not flow, but we only see the
appearance of noise. This will also be the case for the relativistic Schroedinger operator we
shall introduce below, and for the Dirac theory. Remarkably, de Broglie as early as 1934,
insisted in the introduction of an evolution parameter and the formulation of a covariant
Schroedinger theory; this was followed by the pioneering work by Stueckelberg(29) followed
by Horwitz and Piron (1973).(29) In the realization of this idea by Kyprianidis,(30) we have
a variable mass ensemble under Brownian motion, with particle and antiparticle creation
and annihilation, the latter being the idea that paved the way to the formulation of quan-
tum field theory by Feynman and Schwinger in terms of τ (31) following Stueckelberg. In
contrast with our formulation, Kyprianidis relates this proposal that he elaborated in terms
of de Broglie’s ideas, to fluctuations of the metric.(30) More contemporay elaborations of
this idea, are due to Land, Shnerb and Horwitz,(29) and finally, Collins and Fanchi.(32)
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1397

which the electromagnetic field will be written in terms of a Dirac-Hestenes


spinor field, and we shall further relate to relativistic quantum mechanics.

3.1. The Cartan–Weyl Trace-torsion, Hertz’s Potential and Maxwell’s


Equation

In this section we shall work on Cl (M, g) with M a space–time man-


ifold with a Riemannian metric g, and still we have a decomposition of
the Cartan–Weyl 1-form given by Eq. (30), which we shall rewrite as

Q = dln ψ 2 + Aharm + Acoex ∈ sec(T ∗ M) ⊂ sec(Cl (M, g)), (35)

where ψ is a real-valued function on M. We shall present now the theory


of the Hertz potential.6 In this section, to keep with notation originally
due to Stratton, we write

Acoex = −δ, for  ∈ sec(2 (T ∗ M)) ⊂ sec(Cl (M, g)). (36)

We take an orthonormal basis {γµ ∈ sec(T ∗ M) : µ = 0, . . . , 3} such that


γ µ .γ ν = g µν and define γ 5 = γ 0 γ 1 γ 2 γ 3 , so that (γ 5 )2 = 1, in distinction
with Minkowski space. We write  = 1/2µν γµ γν . Let

γ 5 S = d ∈ sec(3 (T ∗ M)), for S ∈ sec(T ∗ M). (37)

The 1-form S when defined on Minkowski space is the Stratton potential.(37)


Since γ 5 has maximal degree, and S̃ = S, and then 0 = d(d) = d(γ 5 S) =
−γ 5 δS as follows from Eq. (17), which we shall repeatedly use in the sequel
in the subsection, so that

δS = 0. (38)

We now define the electromagnetic potential 1-form

A = ∂ = (d − δ) = γ 5 S + Acoex ∈ sec(3 (T ∗ M)). (39)

6 We shall extend the work of Rodrigues and associates in Minkowski space (which as we
already discussed should be properly situated as the metric for the propagation of singu-
larities of the electromagnetic potential which according to Fock can have superluminal
propagation);(24) this work has led to find new subluminal and superluminal solutions
(that can be experimentally launched) in Minkowski space,(23,37) and also led to solu-
tions of the Witten–Seiberg monopole equations.(38) In these works, a central role is
played by the Hertz potential, but no connection is established with a RCW geometry
and its coexact term of the Cartan–Weyl torsion.
1398 Rapoport

Then, we define the electromagnetic strength 2-form

F : = ∂A = (d − δ)(Acoex + γ 5 S)
= ∂Acoex + d(γ 5 S) − δ(γ 5 S)
= Fcoex + d 2  + γ 5 dS = Fcoex + γ 5 dS
= Fe + γ 5 Fm , (40)

are the electric and magnetic intensity fields, respectively, with Fe =


Fcoex = dAcoex = −dδ, and Fm = dS, both are 2-forms, as well as F.
We observe that F = 0 if and only if Fcoex = −γ 5 Fm . Also,

∂F = (d − δ)F = (d − δ)Fcoex + (d − δ)(γ 5 dS)


= ∂Fcoex − γ 5 δdS + γ 5 d 2 S
= Je − γ 5 δdS := Je + γ 5 Jm (41)

with Je = Jcoex = ∂Fcoex and Jm = −δdS are the electric and magnetic
current source 1-forms, respectively. Then,

Je = ∂Fcoex = (d − δ)2 Acoex = 1 Acoex (42)

and from Eq. (38) follows that

Jm = −δdS = −(δd + dδ)S = 1 S. (43)

Note the striking similarity between these currents. While the electric
source current is the action by 1 on the coexact potential in Q, the mag-
netic current appears by the same action on the Stratton potential, that
has also appeared from this potential through a duality construction. If we
further add the harmonic term Aharm in Q (see Eq. (35)) to A in Eq. (39),
then it will not produce a contribution to F. Thus, Aharm is a Aharo-
nov–Bohm potential that should not be neglected, which plays an impor-
tant role in breaking the time-reversal of diffusion processes generated by
H0 (g, Q) and further will appear in the Dirac-Hestenes equations associ-
ated to a magnetic monopole when normalized by the invariant density.
Therefore, from Eqs. (42) and (43) follows that the Maxwell’s system for
the Hertz potential can be written as a single equation

∂F = Je + γ 5 Jm ≡ 1 (Acoex + γ 5 S), (44)

where we have used that ∗ = ∗ or still γ 5  = γ 5 in the last term of


Eq. (44). Thus,

∂F = Je if and only if 1 S = 0, (45)


Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1399

so that it follows from Eq. (45) that the vanishing of a source for mag-
netic monopoles is equivalent to the fact that S is an harmonic 1-form.
Then, we can state the following important result proved for Minkowski
space by Rodrigues(37) , from which he deduced the Stratton’s equations(39)
for Maxwell’s theory which play a central role in his theory:

Theorem . 2  = 0 ⇒ ∂Fe = 0.

Proof. We simply note that ∂Fe = (d − δ)(−dδ) = −d 2 δ + δdδ =


−δ(dδ + δd) = −δ2  = 0.

3.2. The Decomposition of the Cartan–Weyl Form and the Stationary


State

We wish to elaborate further on the decomposition of Q in the par-


ticular state in which the diffusion process generated by H0 (g, Q) and its
extensions to differential forms, in the case M has a Riemannian metric
g, has a τ -invariant state corresponding to the asymptotic stationary state.
Thus τ , which is the time-evolution parameter of the diffusion process,
is in the general case not to be confused with the time variable t of M,
whenever M is a space–time. This time parameter is the one introduced by
Stueckelberg (and then introduced in quantum field theory), further elabo-
rated by Piron and Horwitz and in several works by Horwitz and cowork-
ers,(29) Fanchi,(40) Trump and Schieve,(41) Pavsic,(5) and in the context of
a Schroedinger space–time operator, by Kyprianidis,(30) Collins and Fan-
chi,(32) and may still be related to Liouvillian time in Prigogine’s theory
of nonequibilibrium statistical mechanics.(42) Thus, we shall concentrate in
the diffusion processes of differential forms generated by
1
Hk (g, Q) = (k + LQ̂ ), 0 ≤ k ≤ n, with Q = d ln ψ 2 + Acoex + Aharm .
2
(46)

We recall from our previous article, that by taking an isometric immersion


f of M on Euclidean space, the noise tensor becomes X = ∇f , the diffu-
sion of k-forms for 1 ≤ k is completely determined by the diffusion of the
scalar fields, which has for generator H0 (g, Q), on which the diffusion tak-
ing place in sec(k (T ∗ M)) is fibered. So we may restrict ourselves to this
case, in which for a scalar function φ : M → R,
1
H0 (g, Q)φ = (g φ + g(grad ln ψ 2 , grad φ) + g(Â, grad φ)),
2
with A := Acoex + Aharm . (47)
1400 Rapoport

Where  is, as before, the vectorfield given by the g-conjugate of the 1-


form A. This is the invariant form of the (forward) Fokker–Planck oper-
ator of this theory (and furthermore of the Schroedinger operator when
introducing the phase function to the exact term of Q).(34,43) Then, the
random motions generated by H0 (g, Q) are determined by integrating the
invariant non-linear Ito stochastic differential equations
1
dx(τ ) = X(x(τ ))dW (τ ) + Q̂(x(τ ))dτ, (48)
2
where W (τ ) denotes a standard Wiener process on R m (the noise tensor
is a linear mapping X(x) : R m → Tx M such that (XX † )αβ = g αβ ) such
that its mean is zero and covariance dW i (τ )dW j (τ ) = δji dτ , and Q̂ is
as before, the g-conjugate of Q drift vectorfield.7 Note that the knowledge
of X allows us to reconstruct g, and being g positive-definite, the knowl-
edge of Q̂ allows to reconstruct Q and together we reconstruct the RCW
connection. We are interested now in the volg -adjoint operator defined in
L2 (sec(n (T ∗ M))), which we can think as an operator on densities, φ.
Thus,
1
H0 (g, Q)† φ = (g φ − divg (φgrad ln φ) − divg (φ Â)). (49)
2
The operator described by Eq. (49) is the backward Fokker–Planck opera-
tor.(34) The transition density p ∇ (τ, x, y) is determined by the fundamental
solution (i.e. p ∇ (τ, x, −) → δx (−) as τ → 0+ ) of the equation on the first
variable
∂u
= H0 (g, Q)(x)u(τ, x, −). (50)
∂τ
Then, the diffusion process {x(τ ) : τ ≥ 0}, gives rise to the Markovian
semigroup {Pτ = exp(τ H0 (g, Q)) : τ ≥ 0}8 defined as

(Pτ f )(x) = p∇ (τ, x, y)f (y)volg (y), (51)

7 Here we are considering the general situation in which the noise tensor X is not nec-
essarily derived from an isometric immersion of M, as presented above. Generically, X
will have a diffusion constant, say the square root of the kinematical viscosity for Na-
vier–Stokes equations, or (–h/mc)1/2 for a quantum particle with mass m in the case of
quantum mechanics, setting thus a fundamental length for the fluctuations of the system.
8 A Markovian semigroup in a Hilbert space H is a family of bounded positive linear

operators {Pτ , τ ≥ 0} with dense domain contained in H, such that P0 = I d verifying


the following properties: (i)(semigroup property) Pτ ◦ Pτ  = Pτ +τ  , τ, τ  ≥ 0, (ii) (contrac-
tion property)Pτ  ≤ 1, τ ≥ 0, and (iii)τ → Pτ is strongly continuous.
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1401

has a unique τ -independant-invariant state described by a probability den-


sity ρ independant of τ determined as the fundamental weak solution
(in the sense of the theory of generalized functions) of the τ -independant
Fokker–Planck equation:

1
H0 (g, Q)† ρ ≡ (−δdρ + δ(ρQ)) = 0. (52)
2
Let us determine the corresponding form of Q (similarly as in Ref. 44),
say Qstat = d ln ψ 2 + Astat . We choose a smooth real function U defined
on M such that

H0 (g, Qstat )† (e−U ) = 0, (53)

so that from Eq. (49)

−de−U + e−U Q = δ(−δ + Aharm ) (54)

for a 2-form  and harmonic 1-form Aharm ; thus, if we set the invariant
density to be given by ρ = e−U volg , then

A
Qstat = d ln ψ 2 + , with A = −δ2 + Aharm . (55)
ψ2

Now we project A/ψ 2 into the Hilbert-subspaces of coexact and har-


monic 1-forms, to complete thus the decomposition of Qstat obtaining
thus a Hertz and Aharonov–Bohm potential 1-forms for the stationary
state, respectively, yet these potentials have now a built-in dependence on
the invariant distribution, and although they give rise to Maxwell’s theory,
the interpretation is now different.9 Indeed, we have an inhomogeneous
random media, and these potentials depend on the τ -invariant distribution
of the media. We shall next see how these potentials appear in the con-
text of the equivalence of the Maxwell sourceless equation on Minkowski
space written in terms of a Dirac-Hestenes spinor field, and the non-lin-
ear Dirac-Hestenes equation for these fields, albeit in Minkowski space

9 A word of caution. In principle, −δ/ρ and Aharm /ρ may not be the coexact and har-
monic components of A/ρ, respectively. If this would be the case, then we obtain that
d ln ψ is g −1 -orthogonal to both −δ and Aharm ; furthermore d ln ψ ∧ Aharm = 0, so fur-
thermore they are collinear. This can only be for null Aharm or constant ρ, so that the
normalization of the electromagnetic potentials is by a trivial constant. In the first case
the invariant state has the sole function of determining the exact term of Q to be (up
to a constant) d ln ψ.
1402 Rapoport

provided with a RCW connection with trace-torsion given by Qstat . Yet,


we can exploit further the Hodge-decomposition of Qstat to manifest the
quantum potential as built-in. Indeed, if we multiply Eq. (55) by ψ and
apply ∂, then we get that d ln ψ, and the coexact and harmonic terms of
Qstat decouple in the resultant field equation which turns out to be

g ψ = [g −1 (d ln ψ, d ln ψ) − δd ln ψ]ψ (56)

with nonlinear potential 2φ := g −1 (d ln ψ, d ln ψ) − δd ln ψ, which has the


form of (twice) a relativistic quantum potential extending Bohm’s poten-
tial in non-relativistic quantum mechanics;(45) cf. also Eq. (56), from which
follows that φ = 2V , and Eq. (122) below from which follows that φ =
1/6 R(g), where R(g) is the metric scalar curvature. It is interesting to
remark, as we shall see further below, that ψ are the scaling functions
appearing in the Einstein λ transformations,(27) later extended to define the
Dirac-Hestenes spinor fields.(44) In the context of non-relativistic quantum
mechanics, |ψ|2 is the Born density on introducing into ψ the imaginary
phase function iS, which in our formalism is γ5 S, with γ5 = γ0 γ1 γ2 γ3 ,
with {γi : i = 0, . . . , 3} a basis of R1,3 or R3,1 . In statistical mechan-
ics of equilibrium systems, it plays the role of a Gibbs measure.(46) Also,
it is remarkable that the electromagnetic potential in the stationary state
Astat = A/ψ 2 given by Eq. (55) appears in the problem of extending the
gauge transformations due to the presence of torsion on M, where it is
found that this extension is possible if one allows for violation of electric
charge or magnetic flux conservations; here the problem is different from
ours: Does electromagnetism couple to the gravitational field (i.e. torsion
in Minkowski space, or still curvature in curved space)? The answer is
negative unless it comes through the constitutive equations.(47) In our set-
ting, electromagnetism is built into the gravitational field as a term of the
trace-torsion of the RCW connection (see Eq. (26)) whose Hertz potential
term participates in the RCW curvature, and no such violations previously
alluded are assumed.

3.3. Time Ireversibility of Diffusions and the electromagnetic Terms of Q

In this section we shall see that the electromagnetic potential terms in


Q are related to the breaking of the time-evolution reversal of the diffu-
sion processes generated by H0 (g, Q). Consider the τ -dependant probabil-
ity vectorfield defined by the one-point probability density p ∇ (τ, x), for
x ∈ M,
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1403

Q̂(x) 1
J (τ, x) := p∇ (τ, x) − grad p∇ (τ, x), (57)
2 2

with Q̂ as before, the g-conjugate of Q, so that Q̂/2 is the drift vector-


field. Then, the Fokker–Planck equation (50) for the one-point distribution
p∇ (τ, x) can be rewritten as

∂p ∇ (τ, x)
+ divg Jτ (x) = 0, (58)
∂τ

which in the stationary state, yields the probability vectorfield

ρ 1
Jstat = Q̂stat − grad ρ, (59)
2 2

where


Q̂stat = grad ln ρ + , with A = −δ + Aharm . (60)
ρ

Then, it is easy to check that Jstat reduces to


Jstat = , (61)
2

the g-conjugate of the electromagnetic potential terms of Q. Thus, Jstat is


a conserved probability vectorfield, since

δA
divg Jstat = − = 0. (62)
2

Thus, the electromagnetic terms of A or still the unnormalized non-exact


potentials of Qstat must both vanish for a null probability vectorfield of
the diffusion process. A non-null probability vectorfield is equivalent to
the τ -irreversibility of the diffusion process {x(τ ) : τ ≥ 0} in the station-
ary state, i.e. if the probability in the stationary state of a particle at time
τ = 0 in x ∈ M, to be10 at y ∈ M at time τ , i.e.

p∇ (τ, x|y, 0)ρ(y) = p ∇ (τ, y|x, 0)ρ(x), ∀x, y ∈ M, τ ≥ 0, (63)

10 We remark, not to be found; probability has nothing to do, in this context, with the
“observer”; it is τ which has to be related to the observer’s experience of time-flow.
1404 Rapoport

where p ∇ (τ, x|y, 0) denotes the conditional probability which is the funda-
mental solution of the forward Fokker–Planck equation (50) with the ini-
tial condition stated there (see Eq. (5.3.46) in Gardiner(34) ).11 There is a
most simple characterization of irreversibility in terms of the differential
generator of a diffusion process in the stationary state due to Kolmogo-
rov.(40) Thus, in our setting, τ -reversibility is verified whenever for any two
smooth compact supported functions f, h defined on M, we have that

(H0 (g, Q)f )(x)h(x)ρ(x)volg (x) = f (x)(H0 (g, Q)h(x))ρ(x)volg (x) (64)

and thus it can be seen that this is the case if and only if δ and Aharm
vanish completely. The following aim of this article, is to show that the
invariant decomposition of Qstat on Euclidean space (valid in this case
with compact supported terms in its decomposition)(12) leads to the equiv-
alence of the Maxwell equation on Minkowski space and the Dirac-Hest-
enes equation for a Dirac-Hestenes spinor operator field on Minkowski
space provided with a RCW connection. This will place strict conditions
on δ and Aharm .

3.4. The Equivalence of the Free Maxwell Equation and the


Dirac-Hestenes Equation on a RCW Minkowski Space, and
Relativistic Brownian Motions

That there might be a unification of relativistic quantum mechan-


ics and electrodynamics, has been a study of intensive research, proba-
bly starting with Salhoffer’s work (a disciple of Schroedinger), followed by
Campolattaro,(49) and works by Rodrigues and associates.(50) The latter’s
presentation of this equivalence, adquired a simple representation in the
framework of the the Clifford bundle over Minkowski space, where we fur-
ther write the electromagnetic field 2-form in terms of a Dirac-Hestenes
spinor operator field (DHSF, for short). In fact the equivalence between
the Maxwell equation and the Dirac-Hestenes equation, was established
whenever the latter is written as an equation on Minkowski space pro-
vided with the exact potential term d ln ψ, where |ψ|2 defines the prob-
ability density associated to the Dirac equation. An extension of this by

11 For a thorough presentation of time-evolution parameter ireversibility and detailed bal-


ance, which is the implementation of this notion in thermodynamics and in phase-space
following Boltzmann, see Gardiner,(34) from which our definition in configuration space
has followed. In fact we must stress that under the change of sign of τ , the space–time
variables (t, x) do not change sign, i.e., they are even variables.
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1405

including the electromagnetic potential terms in Q was proved by the


author,(51) and this will be the subject of our presentation on this sec-
tion. So the starting point due to Rodrigues et al.(50) is the Rainich–Mis-
ner–Wheeler theorem: A non-null electromagnetic field can be reduced by
a Lorentz transformation and a duality rotation to an extremal field, i.e.
a field such that its electric (magnetic) component is zero (parallel to a
coordinate axis). Recall that we discussed above the fact that the Lorentz
group enters in the Maxwell system, as the invariance group for the propa-
gation of singularities, and in the present approach, on the canonical form
in which the electromagnetic field intensity is written in terms of spinor
fields. Then, on Cl (M, ν), where ν is Minkowski metric,12 a non-null elec-
tromagnetic field F is written as

˜
F = bγ 1 γ 2  (65)

with  a representative of a DHS,13 whose canonical representation is


5 1
 = (ρeβγ ) 2 R, (66)

where β, b ∈ sec(0 M), R ∈ Spin+ (1, 3), such that R R̃ = RR −1 = 1; here


˜ = (ρeβγ 5 )1/2 R̃. Here, β is the Takabayasi phase. We stress the fact that
there exists a global spin structure on M if and only if there exists an even
multiform  defined on M; furthermore, any Dirac spinor is obtained by
multiplying a DHSF by an idempotent of the Clifford algebra.(6)
The sourceless Maxwell equations has as representative on the Clifford
bundle the single equation

∂F = 0. (67)

Inserting Eqs. (65) and (66) in Eq. (67) and further multiplying on the
right by ˜ −1 , we obtain the non-linear Heisenberg-type equation

∂γ 1 γ 2 + F () = −(∂ ln b)γ 1 γ 2 , (68)

12 It has been proved by Pezzaglia(19) that the Maxwell system and the Dirac equation
for even multiforms, say, the Dirac-Hestenes equations, are form invariant under either
choice of the signature of the Minkowski metric. An important difference appears in
taking R3,1 where the spin-plane has an Euclidean metric, and thus one can think on
the possibility of a diffusion process on it.
13 In fact, DHSF depend on the choice of a spinor basis; for a thorough presentation of

DHSF fields on Minkowski space, and more generally, on Lorentzian manifolds, and
the formulation of the DH equation in these settings, we refer to recent works by
Rodrigues.(6)
1406 Rapoport

where

˜
F () = γ 1 γ 2 (∂ )( ˜ −1 ).
 (69)

From the fact that R R̃ = 1, we have that


1
∂µ R = µ R, (70)
2
where

µ = 2(∂µ R)R̃. (71)

Define the spin 2-form as


–h
S= Rγ 1 γ 2 R̃. (72)
2
Since
5
∂µ  = [∂µ ln(ρeβγ )1/2 ] + 1/2µ  (73)

the nonlinear Heisenberg equations (68) and (69) is therefore written as

1 5 1
∂γ 1 γ 2 − – γ µ Sµ  = −(∂ ln b)γ 1 γ 2 − [∂ ln(ρeβγ ) 2 ]γ 1 γ 2 . (74)
h
One can rewrite this equation as

5 1 1 γ5 1
(ρeβγ ) 2 [∂Rγ 1 γ 2 − – γ µ Sµ R] = −(∂ ln b)γ 1 γ 2 −2[∂ ln(ρeβ ) 2 ]γ 1 γ 2.
h
(75)

If  = R satisfies the non-linear DH equation

1
∂γ 1 γ 2 − – γ µ Sµ  = 0 (76)
h
from Eq. (75) we obtain that we must have
5 1
∂ ln b = −2∂ ln(ρeβγ ) 2 (77)

or, equivalently
5
b = K/(ρeβγ ). (78)
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1407

Note that this fixes β to be constant and equal to 0 or π , correspond-


ing to the electron and the positron, respectively. Inserting Eq. (78) into
Eq. (75) we obtain a non-linear DH equation

1 5 1
∂γ 1 γ 2 − – γ µ Sµ  = [∂ ln(ρeβγ ) 2 ]γ 1 γ 2 . (79)
h

Since S and µ , are both 2-forms, their Clifford product Sµ can be
decomposed as the even multi-form

Sµ = Sµ 0 + Sµ 2 + Sµ 4 . (80)

Here we denoted A i , i = 0, 2, 4 as the term of degree i of the multi-form


A. Now we write

Sµ 0 = −pµ ∈ sec(0 M), i.e.pµ : M → R, (81)


Sµ 2 = Eµ,αβ (γ ∧ γ ), Eµ,αβ : M → R,
α β
(82)
5
Sµ 4 = γ rµ , rµ : M → R. (83)

Yet, we need in Eq. (79) the expression γ µ Sµ , so that we compute

γ µ Sµ 2 = Eµ,αβ γ µ (γ α ∧ γ β )
e g e g
= − Aµ γ µ − γ 5 Bµ γ µ = − A − γ 5 B, (84)
c c c c

where A = Aµ γ µ and B = Bµ γ µ are 1-forms such that

e
Aµ = ηνσ Eν,µσ (85)
c

and
g
Bµ = ηµν νσρτ Eσ,ρτ (86)
c

here νσρτ is the canonical completely skew-symmetric tensor. Also,

γ µ Sµ 0 = −γ µ pµ = −p (87)

and

γ µ Sµ 4 = γ µ rµ = r. (88)
1408 Rapoport

If we define the velocity vectorfield

v = Rγ 0 R̃, (89)

so that

˜
ρv = γ 0  (90)

is the Dirac current, the mass of  is defined as a geometrical kinematical


parameter by
5
p = mceβγ v = mcγ 0 . (91)

Similarly, we define a scalar quantity µ, by


5
r = µceβγ v = µcγ 0 . (92)

Therefore, the non-linear DH Eq. (79) takes the form



e g ceβγ
5

∂γ 1 γ 2 + – A + – γ 5 B  + (m + γ 5 µ) – γ 0
hc hc h
βγ 5 1/2 1 2
= ∂ ln[(ρe ) ]γ γ . (93)

We note that although m and µ were introduced similarly, in terms of p


and r where the components of p and r arise from the Clifford product
of γ µ with the scalar and pseudoscalar terms of Sµ , respectively, both
have the units of a mass; yet in Eq. (93) µ appears multiplied by the chiral
factor γ 5 ; then by following Lochak(52) and further Daviau, and Daviau
and Lochak,(52) B can be thought as the potential of a magnetic mono-
pole with mass µ.14

14 If we set A, m, µ and the right hand side of Eq. (93) equal to zero, we obtain the Di-
rac-Hestenes form of the Dirac equation (19), or (144) due to Lochak (up to a sign
in B) for a massless spin-1/2 magnetic monopole; see Lochak.(52) The introduction of
m and µ has the effect of breaking the phase and chiral gauge invariance, respectively,
of the linear Dirac equation. Further below we shall see that A and B can be natu-
rally associated with the normalized Hertz and Aharonov–Bohm potentials, respectively,
in the Hodge decomposition of the trace-torsion Q already discussed, and that indeed
by rescaling  one does obtain a linear Dirac-Hestenes equation. Furthermore, our term
g/ch̄B equals Lochak’s term −1/2, where  is a 1-form equal to ∗φ, with φ a 3-form
which yields the components of a connection with completely skew-symmetric torsion on
Minkowski space. By the way , this explains how one can pass from a RCW connection
to a connection with completely skew-symmetric torsion: For the latter we only have to
consider ∗Q. For this connection in flat space, Lochak obtains a remarkable equation,
namely that its scalar curvature equals −6ν −1 (, ).
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1409

We wish to confront this equation with the DH equation on Min-


kowski space provided with a RCW connection,(11,26) which we shortly
introduce. The essential fact is that  given by Eq. (66) acts on vectors
as γ µ  ˜ = ρRγ µ R̃, and we have the following Einstein λ transforma-
tions: (27) Cρ : γ µ → γ̄ µ := ρ 1/2 γ µ , which for the reciprocal basis
γ̄ .γ̄ν = δν we must have Cρ (γµ ) = γ̄µ = ρ −1/2 γµ . Thus, we have an
µ µ

induced metric ν̄ defined by ν̄(γ µ , γ ν ) = ν(ρ 1/2 γ µ , ρ 1/2 γ ν ) = ρν(γ µ , γ ν ),


which coincides with Weyl’s scale transformations. Thus, on Minkowski
space we can introduce a non-trivial RCW connection, as in the following.
If ∇ denotes the Levi–Civita (trivial) connection defined by ν, we define
the induced connection ∇˜ by ρ 1/2 ∇˜ γµ γ ν := ∇γµ (ρ 1/2 γ ν ).(26) This connec-
tion has non-trivial coefficients ˜ µα ν =  ν − (1/2)(∂ ln ρδ ν ), which pre-
µα γµ α
serves the metric ν̄. If we take orthonormal vectors γ a = θµa γ µ , then,
ν̄(γ a , γ b ) = θµa θνb η̄µν = ν ab , Minkowski metric; note that in our case
θµa = ρ −(1/2) δµa . One can easily check that ˜ µν a =  a . Furthermore, for
µν
˜
the Dirac operator ∂ = γ µ ∂γµ , the corresponding Dirac operator D∇ =
γ̄ µ ∂˜γ̄µ = ρ 1/2 γ µ ∂ρ −(1/2) γµ = γ µ ∂γµ = ∂. Yet, both curvature and torsion
are no longer trivial. In fact
1 µ 1
ν = [(∂ ln ρ 2 )δ − (∂ ln ρ 2 )δ ν ]
T̃µα γα ν γµ α
1
(94)
τ
R̃µνσ = ([∂γσ , ∂γν ] ln ρ 2 )δµτ

the latter vanishes away of the nodes of ρ where it becomes singular. So


we have the already familiar RCW connection, on Minkowski space (in
Euclidean space by replacing ν by the Euclidean metric), with trace-tor-
sion 1-form Q = Tαν α γ ν = 3 ∂ln ρ, so that the scalar field ψ in Eq. (55) is
2
3
such that ψ 2 = ρ 2 . Now for spaces with curvature and torsion the Dirac-
Hestenes equation is (see Eq. (209) in Ref. 11)
5
Q mceβγ
∂ −  γ 1 γ 2 + 0
–h γ = 0, (95)
2

where ∂ = Dη is the spinor-Dirac operator associated to a Lorentzian me-


trix η, {γ̃ µ , a = 0, . . . , 3} are orthonormal vectors, and ∂ = γ µ [∂µ +
2 µ ], where µ = 2 µ γa γb , which for the Minkowski metric, vanishes
1 1 ab

and we are left with the Dirac–Kahler operator in Eq. (95). Yet we shall
consider a more general equation with magnetic monopole mass µ:
  5
1 ceβγ
∂ − Q γ 1 γ 2 + (m + µγ 5 ) – γ 0 = 0. (96)
2 h
1410 Rapoport

If we assume that Q has a Hodge decomposition associated to the


stationary state defined by the invariant density ρ of a Brownian process
on Euclidean space.15 From Eq. (55) 16 we recall that the form of Q is
then
3 δ Aharm
Q= d ln ρ − 3 + 3
, (97)
2 ρ2 ρ2
where −δ and Aharm are, respectively, coexact and harmonic 1-forms,
the Hertz and Aharonov–Bohm potentials, respectively, and that in terms
of the Euclidean metric, these are compact Hilbert orthogonal 1-forms.
Replacing thus Eq. (97) in Eq. (96) we obtain that this equation is written
as the non-linear DH equation with monopole mass µ
   
3 ẽ δ g̃ Aharm
∂ − ∂ ln ρ − – 3 + –  γ 1γ 2
4 2c h ρ 2 2c h ρ 23
5
ceβγ
+(m + µγ ) – γ 0 = 0,
5
(98)
h
where we have written coupling constants ẽ and g̃ for the coexact and har-
monic electromagnetic potentials in the trace-torsion 1-form. This equa-
tion was obtained in Ref. 26 with µ ≡ 0 and further considered in Ref. 50
by Rodrigues and Vaz (1993) always in the restricted case of an exact Q ;
we remark that these authors gave no characterization of the full form of
Q and its relation to Brownian motions), and in the full present case in
Ref. 51. We now put as in de Broglie’s double solution theory (45)
1 3 1
φ = ρ − 4 , ψ = ρ − 4 , so that φ = ρ 2 ψ. (99)

The DH equation on Minkowski space with a RCW connection (96) yields


the following eqts.:
   
ẽ δ g̃ Aharm 5
∂ψ + – 3
−– 3
ψ γ 1 γ 2 + (m + γ 5 µ)ceβγ = 0, (100)
h2c ρ 2 h2c ρ 2

15 Recall that from Eqs. (5) and (6) we can always transform a 1-form on the Euclidean
Clifford structure, to a representation on the non-degenerate case. When dealing with
R3,1 the spin-bundle is left invariant by this transformation, and it has an Euclidean
structure, which in the case of R1,3 is no longer the case. We shall present below the
equations for this diffusion process.
16 It is at this point that the role of the one-half coefficient on Q in (26) will become

apparent; it fixed in Eq. (55) the form of Qstat and in particular the normalization of
the electromagnetic potentials. This will be crucial in the following.
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1411

which is precisely the linear DH equation for ψ submitted to an exterior


3
electromagnetic potential 1-form (−δ + Aharm )/ψ 2 (up to physical con-
stants), i.e. the normalized electromagnetic potentials in Qstat , and we still
have
    5
ẽ δ g̃ Aharm ceβγ
∂φ + – −– φ γ γ + (m + γ µ) – φγ 0
1 2 5
h2c ρ 23 h2c ρ 23 h
5
= [∂ ln (ρeβγ )1/2 ]φγ 1 γ 2 . (101)

If we compare Eq. (101) with Eq. (93) in which we substitute  by φ,


we then note that up to an identification of the electromagnetic exterior
potential 1-forms and the interior ones appearing in Eq. (93) that they are
the same equation. Thus, we are lead to posit the identification of 1-forms
 
ẽδ − g̃Aharm
3
.γ 1 γ 2 = eA + γ 5 B, (102)
2ρ 2

so that
 
ẽδ − g̃Aharm
3
∧ γ 1 γ 2 = 0, (103)
2ρ 2

is valid. Due to the fact that Aharm has a zero field away of the nodes of
ρ (it is harmonic), it is natural to further make the identifications

ẽ δ 1 2
.γ γ = eA (104)
2 ρ 23

and

−1 Aharm 1 2
g̃ 3
.γ γ = gγ 5 B. (105)
2 ρ2

so that it is still natural to take ẽ = e and g̃ = g. Thus, under the iden-


tifications given by Eqs. (104) and (105), we have therefore proved that
the DH equation (101) on Minkowski space with a RCW connection for
−1
φ = ρ 4 , or still Eq. (94) for , coincides with Eq. (67) for φ or still, is
equivalent to the free Maxwell equation for F built in terms of φ instead
1412 Rapoport

of , so that taken in account also Eq. (78) then b = 1


ρβ , and we have
that

R 1 2 S
∂F = 0, with F = bφγ 1 γ 2 φ̃ = 1/2
γ γ R̃ = – 1/2
ρ hρ
= ψγ 1 γ 2 ψ̃, (106)

and additionally we have

eδ − gAharm
3
∧ γ 1γ 2 = 0 (107)
ρ 2

i.e.

Jstat ∧ γ 1 γ 2 = 0, (108)

away of the nodes of ρ. Note that since ψ solves the linear DH Eq. (100),
ρF is precisely the electron’s magnetization or magnetic moment density.
Also, under this equivalence, the electromagnetic potentials appearing as
internal to the structure of the DHSF, more specifically as provided by the
term γ µ Sµ 2 of the non-linear functional of the Heisenberg equation,
can be alternatively interpreted as the interior product of the electromag-
netic potentials appearing in the trace-torsion with the spin-plane two-
form γ 1 γ 2 . Thus, we have a spontaneous reduction of components of δ
(the true electromagnetic potential producing a non-trivial field strength,
as discussed when treating the theory of the Hertz potential) and that of
Aharm , which has a zero-field, to simply two components each of them,
as they are those given in the spin-plane. This condition which enforces
the equivalence between the Maxwell and Dirac-Hestenes equations are
very intriguing, since we have as in the the Gupta–Bleuler quantization
of the electromagnetic field, a reduction of the components to the trans-
versal ones, i.e. to the spin-plane. Furthermore, these conditions actu-
ally give the necessary reduction of degrees of freedom, eight for  to
six for F to be actually an electromagnetic field; this stands in contrast
with the gauge fixing conditions suggested by other authors when study-
ing Maxwell equation with sources and its relations with the Witten–Sei-
berg monopole equations.(37) So we can write the diffusion process given
by Eq. (48) produced by the Euclidean metric (and in fact, as we shall see,
also the diffusion produced by the Minkowski metric ν ) and the Cartan–
Weyl 1-form Q given in Eq. (97) associated to the above equivalence in
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1413

the following form (we set β = 0):


 – 1
h 2 3 –h 1
dx(τ ) = dW (τ ) + grad ln ρ(x(τ ))dτ + A(x(τ ))dτ, (109)
mc 4 mc 2

where if we denote A = 1
3 (−eδ + gAharm ), then
c–
hρ 2

A(x) = A1 (x)e1 (x) + A2 (x)e2 (x) = A1 (x)γ̃1 (x) + A2 (x)γ̃2 (x) (110)

with x = (t, x 1 , x 2 , x 3 ) ∈ R 4 (or alternatively Minkowski space), where


(recall Eq. (6))


2
W (τ ) = W µ (τ )eµ = W i (τ )γ̃i + W3 (τ )γ̃3 γ̃0 (111)
i=0

is a four-dimensional orthogonal-group invariant Wiener processes, with


mean W α = 0, ∀α = 0, . . . , 3 and covariance dW α (τ )dW β (τ ) =
δ αβ dτ is as usual, where − means the averaging with respect to the
Gaussian measure on R 4 . Here {eµ : µ = 0, . . . , 3} represent a basis
for the Euclidean Clifford algebra, and {γ̃µ : µ = 0, . . . , 3} repre-
sent a basis for R3,1 obtained from them by the set of transformations
defined in Eq. (6). In this case, the differential generator is the RCW
Laplacian H0 (–h/mcg, Q), where g denotes the Euclidean metric, and the
gradient term ofthe drift vectorfield is (up to a factor) grad ln ρ =
δ µν (∂µ ln ρ)eν = 2i=0 δ µi ∂µ (ln ρ)γ̃i + δ µ3 ∂µ (ln ρ)γ̃0 γ̃3 . We can now pro-
ceed to give a relativistic (i.e. Lorentz-invariant) meaning to Eq. (109).
For this we shall follow a construction of the (1 + 3)-dimensional rel-
ativistic Brownian motion due to Oron and Horwitz,(53) whose starting
point is the construction of a relativistic Wiener process, yet now writ-
ten in a Clifford algebra setting which with further restrictions  turns out
to be Lorentz-invariant: Indeed, we simply write W (τ ) = 3i=0 W α (τ )γ̃α ,
i.e. we expand the Wiener process in terms of the basis of the Clifford
algebra R3,1 , and now the mean is taken with respect to the relativis-
tic Gaussian due to Oron and Horwitz(53) ,17 so that the covariance is
dW α (τ )dW β (τ ) = ν αβ dτ , with ν = diag(−1, 1, 1, 1); note that this
has the problem that dW 0 (τ )dW 0 (τ ) is negative, which is impossible.

2 √
17 It has the form e−µ /( 2dτ ) for space-like jumps, where µ is the invariant spacelike inter-
val of the standard Brownian motion, and a similar expression for time-like jumps, with
µ replaced by a factor times γ̃5 µ.
1414 Rapoport

To solve this3 problem, we follow Oron and Horwitz in assuming that


W (τ ) = α=0 W α (τ )γ̃ , with dW α (τ ) ∈ R, ∀α = 0, . . . , 3 is such
α
that the jumps dW (τ ) are light-like 3 and αwe further assume that we also
have a Wiener process W (τ ) = α=0 W (τ )γ̃5 γ̃α , where γ̃5 = γ̃0 γ̃1 γ̃2 γ̃3
verifies that γ̃5 γ̃5 = −1, with space-like jumps dW (τ ) with real-valued
components;18 in this setting, multiplication by the density γ̃5 plays the
role of the analytical continuation provided in Ref. 53. If we consider
finally both classes of realizations by space-like and time-like jumps with
an appropiate distribution density for the latter as indicated in footnote
17, it follows from the proof in Ref. 53 that the Fokker–Planck opera-
tor for the relativistic Wiener process is the Lorentz-invariant free d’Al-
embert hyperbolic operator. Consequently, the generator of the noise term
of Eq. (109) is the Lorentz-invariant operator H0 (–h/mcν, 0), so that the
generator of Eq. (109) in which the terms of the drift are also expanded
in terms of the basis {γ̃α : α = 0, . . . , 3} of R3,1 , is the Lorentz-
invariant hyperbolic operator H0 (–h/mcν, Q), with drift vectorfield19 21 Q̂ =
3–
4 h/mcgrad ln ρ(x(τ ))dτ + 2 A(x(τ ))dτ , where
1
the metric gradient oper-
3
ator is defined by ν, i.e. grad ln ρ = µi
i=0 ν ∂µ (ln ρ)γ̃i . Thus, most
remarkably, the development of the relativistic theory follows first from the
choice of the Clifford algebra, in the sense that it amounts to replace the
Euclidean metric g and R4,0 , by the Minkowski metric ν and expand the
drift terms alternatively in the basis of R3,1 , with the further provision of
taking the relativistic Gaussian constructed by Oron and Horwitz instead
of the orthogonal-invariant Gaussian, with the assumptions on the jumps
to be either space-like or time-like. For each fixed sample path W (τ )
as above, we have a random path dependent on it given by integrating
Eq. (109), x(τ, Wτ ), which we may express as x(τ, Wτ ) = x µ (τ, Wτ )eµ =
2 4 –
i=0 x (τ, Wτ )γ̃i + x (τ, W τ )γ̃0 γ̃3 ∈ Cl (R , h/mcg), in the non-degener-
i 3
 3 4 –
ate case, and x(τ, Wτ ) = µ=0 x (τ, Wτ )γ̃µ ∈ Cl (R , h/mcg) in the rel-
µ

ativistic case, both representing a possible trajectory of an electron (or


3
positron), and the invariant distribution of this process is given by ρ 2 ,
20
which is the probability density for particles (antiparticles). Notice that

18 The explicit construction of the dW α is given in Ref. 53.


19 Recall that from Section 2.1, the interaction drift vectorfield in the operators H0 (g, Q)
is (one-half) the g-conjugate of the trace-torsion, Q̂, independently of the signature of g.
In the present case we only need to take care of the exact term, since due to the reduc-
tion of the electromagnetic terms of Q in Eqs. (97) and (110), then in the spin-plane
the metric is always Euclidean and equal to diag(0, 1, 1, 0), so that the electromagnetic
drift terms are identical in both the non-degenerate and degenerate cases.
20 The relativistic or Euclidean Brownian motions generated by the RCW connections with

flat metric given by Minkowski and Euclidean metrics, respectively, can be extended to
the Brownian motions of differential forms, moving along these paths. Indeed, by taking
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1415

we have a diffusion of the time-variable x 0 = t, so that the particle paths


can move both backwards in t (viz. antiparticles, into the future in τ )
as well as forwards in t (viz. particles), just alike in quantum field the-
ory (that was precisely the reason for Feynman and Schwinger separately
introducing τ in quantum field theory, and has remained so in develop-
ing the theory on curved spaces(31) ). These trajectories in fact represent
an infinite number of virtual particles and antiparticles (with pair creation
and annihilation, as the random motion is also on t), which may leave
or either reentry the light-cone. So in Minkowski space, we have a dis-
continuous process describing this entry-exit of the virtual particles (anti-
particles). The x 1 (τ ) and x 2 (τ ) components of x(τ ) = x µ (τ )eµ which
coincide with those of the relativistic invariant diffusion x(τ ) = x µ (τ )γ̃µ ,
µ = 0, 1, 2, 3 represent the irreversible diffusion process on the spin-plane
(which as already remarked has an Euclidean structure when working
with R3,1 ) which has appeared from the equivalence between Maxwell and
Dirac-Hestenes equations. While, the x 0 (τ ) and x 3 (τ ) components are pro-
cesses like one-dimensional Schroedinger equation, with drift terms given
by 3–h/4mc∂ρ/∂t and 3–h/4mc∂ρ/∂x 3 , respectively; for the modern presen-
tation of non-relativistic quantum mechanics, Euclidean field theory and
Bernstein random process; see Zambrini(54) and references therein.
Remarks. From the basic kinematical considerations of the geomet-
rical study of the Dirac equation due to Hestenes in his zitterbewe-
gung interpretation (which in fact was justly claimed by Hestenes as the
main reason for introducing Clifford structures which otherwise had not
appeared before(23) ), we have been able to prove the equivalence of these
equations, in terms of the reduction of the degrees of freedom of the elec-
tromagnetic potentials to two transversal components (see Eqs. (110)), and
more precisely, those of the spin-plane, and those potentials are undis-

the invariant Jacobian process(40) and its k-th exterior product 1 ≤ k ≤ 4 we obtain the

h ν̃, Q) where ν̃ stands for the Euclid-
stochastic differential equation generated by Hk ( mc
ean and Minkowski metrics, alternatively; now, since ν̃ is flat, the Weitzenbock curva-

h ν̃, Q) coincide with H ( – h
ture terms vanish and Hk ( mc 0 mc ν̃, Q) acting componentwise on
k-differential forms (0 ≤ k ≤ 4), so essentially we have determined how multiforms
move randomly under the RCW geometry with Q given by Eq. (97); thus the random
motions of both fermions and bosons have been determined as follows from our discus-
sion of supersymmetric systems below. Therefore, with the choice of the metric and the
associated basic vectorfields that define the Clifford algebra (the reader is suggested to
return to our discussion of this point in Sec. 2), and with the provision of the choice
of the Gaussian density, the present theory for multiforms is Lorentz-group invariant,
or orthogonal-group invariant. Finally, we can construct random Hamiltonian processes
and their Poincaré–Cartan invariants associated to these motions, following our previous
article and the construction of the random invariants of the Navier–Stokes equations
and also of the kinematic dynamo(3) .
1416 Rapoport

tinguishably corresponding to the internal kinematics of the rotational


degrees of freedom, or the exterior potentials of a background geometry
on flat space, which has trace-torsion describing all the degrees of freedom
of the spinor-field. The above equations of diffusion, which have appeared
from this analysis that has allowed us to determine Q, suggest an original
characterization of the quantum paths for both the Maxwell (with a kine-
matical mass term Eq. (91)) and DH equations. The quantization of the
Dirac equations in terms of Brownian motions has been a long standing
problem;(55) the naturality of the present approach is apparent, in spite of
its rawness that calls for further investigation. Indeed, one could attempt
to build, alternatively, the relativistic diffusion process introducing Poisson
jump processes, following Marra and Serva;(56) furthermore, the possibility
of deriving these diffusion processes from the DH equations, is an open
problem (in fact, we have only constructed the random representations
determined by either the Laplacian or d’Alembertian operators given by
the square of the generalized Dirac operator determined by a RCW con-
nection, whose associated DH equation through its equivalence with the
Maxwell equation, lead to this particular form of the diffusions.) We have
identified the node set as supporting magnetic monopoles through the har-
monic component of the trace-torsion, or still by Eq. (105) the set on
which the electromagnetic field associated to the DHSF φ becomes singu-
lar. Remarkably enough, the node set of the amplitude becomes an excep-
tional set for the diffusion, in the sense that only with probability zero the
sample paths of the diffusion can go through the node set because of the
fact that the exact component d ln ρ of the trace-torsion becomes singular
there.(57) Thus, the node set becomes associated to a generalized Meissner
effect in which not only the electromagnetic fields are excluded from the
node set, but actually the diffusing particles are excluded as well; thus,
appears to be natural to study the extension of the present theory to the
consideration of the source-full Maxwell equation. Away from the node
set, the diffusion takes place and the equivalence between the free Maxwell
field and the non-linear DH equation for a DHSF produces on each γ̃ 1 γ̃ 2
spin-plane the breaking of detailed balance of the diffusion process gen-
erated by the RCW geometry, due to the electromagnetic components of
the trace-torsion, whose natural interpretation is that of a radiative pro-
cess on the spin-plane. While this condition was hitherto unknown in the
theory of the electron as approached from the DH equation, the problem
of detailed balance (albeit, in phase space) is central in Stochastic Elec-
trodynamics , i.e. the theory of the electron with the zero-point radiation
field(83) . It appears to be an interesting problem to develop an understand-
ing of the relation between this breaking of detailed balance and the sta-
bility of the electron. Indeed, d ln ρ which yields the quantum potential,
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1417

must play a central role in balancing this radiative process to yield a sta-
ble picture of the electron, where we understand the electron as an ensem-
ble system radiating in the spin-plane (this is the natural interpretation of
having an irreversible diffusion in the spin-plane; a somewhat related pic-
ture appears from the zitterbewegung(23) ). Remarkably, this coincides fully
with de Broglie and Bohm, which succeeded in explaining the stability of
atoms, in the guise of the quantum potential cancelling all classical poten-
tials (viz. Coulomb potential) leading to a non-radiating electron in the
stationary state, which is precisely defined in terms of the real objective
field, here the DHSF  (45) , instead of Bohm’s restriction to work with the
Schroedinger equation, which as shown by Hestenes cannot be taken as
fundamental, since it is an equation for an electron in a fixed homoge-
neous spin state(7,8) . By the way, one should recall that de Broglie insisted
on having a non-linear equation to be able to describe the particle, here
the DH-equation on Minkowski space provided with a RCW connection,
while keeping a linear equation for the field.(45) So, we have obtained a
somewhat related perspective to de Broglie and further Bohm’s recovery of
causal trajectories (incorporating randomness as an additional feature) in
quantum mechanics, here described by the random flow given by integrat-
ing the stochastic differential equations defined by the RCW connection,
and in the case of relativistic quantum mechanics, to a diffusion which
has irreversible terms on the spin-plane. These random flows are consistent
with the basic tenents of Einstein’s principle of general covariance, since
they define active semigroups of diffeomorphisms of space–time,(58) , by
the way, extending the role of classical diffeomorphisms such as the solu-
tions of the Euler equations for perfect fluids,(59) which can be obtained
as the zero viscosity case(3) of the random flows that yield the analytical
integration of the Navier–Stokes equations, on manifolds with or without
boundaries, and in Euclidean and semieuclidean space.(2) In forthcoming
articles, we shall present the diffusions for non-relativistic and relativis-
tic Schroedinger equations generated by RCW connections, placing in the
current perspective Schroedinger’s work in Brownian motion and his equa-
tion.(60) Finally, we want to remark that at the level of the field equation
(52) for the spin-0 field ρ 1/2 -obtained from applying the Dirac operator
to the Weyl form in the non-degenerate case—there is no coupling to the
electromagnetic potentials, if we look for a second order equation for 
by applying the Dirac operator to Eqs. (98) and (101), then the coupling
of the scalar field ρ 1/2 to both A and ∂A will appear in this equation.
This is very natural, since the electromagnetic potentials are derived from
the rotational degrees of freedom of the DHSF, as we have already seen.
This is to be contrasted with the fact that the invariant density for the
present Lorentz-invariant theory yields a Klein–Gordon like equation with
1418 Rapoport

additional term given by the drift in which the electromagnetic terms A


are present. As a final comment to this section, we would like to remark
that the present discussion cannot be final and that a number of addi-
tional questions are left unexplored: What is the evolution parameter of
the irreversible diffusion process on the spin-plane? Is it the invariant cen-
ter-of-mass parameter of the classical electron as a point particle having
zitterbewegung oscillations on the spin-plane?(8,61) We believe these prob-
lems to be of great importance to elucidate the equivalence between elec-
trodynamics and quantum relativistic mechanics.

4. ENERGY FORMS, THE QUANTUM POTENTIAL AND RCW


DIFFUSIONS

In this section we shall show that the RCW geometries yield a natural
formulation of quantum mechanics on manifolds, as an operator the-
ory on two Hilbert spaces. So, in this section and the next, we will dis-
cuss basic issues which on the usual setting have been somehow obvi-
ated and are far from being obvious.(31,62) The basic formalism which
leads to this is the well known remarkable correspondence explored in
flat Euclidean space between the Dirichlet forms of potential theory, Mar-
kovian semigroups and their diffusion processes(63–65) and RCW laplacian
operators,(14) and originates in the canonical commutation relations.(66) In
fact, in quantum field theory on curved space–time, the starting point is
an energy functional for the field associated to a self-adjoint operator on
the Hilbert space determined by the Riemannian volume element.(31,62)
In our theory, this self-adjoint operator will appear to be the conformal
transform of the self-adjoint extension of the RCW laplacian as defined on
an adequate subspace of the ground-state Hilbert space with a weighted
inner product defined by the invariant density. Thus, two Hilbert spaces
are needed: the ground-state Hilbert space on which we have a diffusion
generated by the RCW laplacian which acts as the Fokker–Planck oper-
ator, and the Hilbert space defined by the Riemannian volume in which
this operator transforms into the Schroedinger operator. We shall present
below the above mentioned correspondences.
We assume that M has a Riemannian metric; we assume further that
is four-dimensional space–time (and thus, we are in the situation discussed
by Kyprianidis(32) and Collins and Fanchi(52) ) and a diffusion process with
stationary state ψ 2 volg with null electromagnetic terms in Eq. (55), gen-
erated by H0 (g, d ln ψ 2 ), a Hamiltonian operator on the Hilbert space
L2 (ψ 2 volg ).(14) With abuse of notation, let us denote still as H0 (g, dln ψ 2 )
the Friedrichs self-adjoint extension(63,67) of the infinitesimal generator
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1419

(27) with domain given by C0∞ (M). We can now define the inner product

(f1 , f2 )ρ = 1/2 g −1 (df1 , df2 )ψ 2 volg . (112)

By integration by parts, we obtain

(f1 , f2 )ρ = −(f1 , H (g, dln ψ 2 )f2 )ρ , (113)

where (., .)ρ denotes the weighted inner product in L2 (ψ 2 volg ). Let us
consider now the closed quadratic form, (the Dirichlet form) q associated
to (·, ·)ρ , i.e. q(f ) = (f, f )ρ(63,67,75) We see from Eq. (112) that there is a
unique Hamiltonian operator which generates q, it is the self-adjoint oper-
ator −H0 (g, d ln ψ 2 ). Since the quadratic form is positive, q(f ) ≥ 0 , for
any f ∈ L2 (ψ 2 volg ), then H0 (g, d ln ψ 2 ) is a negative self-adjoint oper-
ator on L2 (ψ 2 volg ) and the Markovian semigroup exp(τ H (g, d ln ψ 2 )) is
defined. Let us see how this construction is related to the usual formu-
lation of Quantum Mechanics in terms of quadratic forms in L2 (volg ),
which in the non- relativistic flat case has been elaborated by several
authors.(64) Consider the mapping Cψ : L2 (ψ 2 volg ) → L2 (volg ) defined
by multiplication by ψ; this is the groundstate transformation and defines
a conformal isometry between the two Hilbert spaces. This map takes
C0∞ (M) into itself. For any f in C0∞ (M) we have

q(ψ −1 f ) = (ψ −1 f, ψ −1 f )ρ

= 1/2 {g −1 (df, df ) − 2g −1 (df, d ln ψ)f

+g −1 (d ln ψ, d ln ψ)f 2 }volg

= 1/2 {g −1 (df, df ) + (divg (b)f 2 + g(b, b)f 2 }volg

1
= f {− g + V }f volg = (f, Hf )L2 (volg ) , (114)
2
where we denoted b = grad lnψ which is the drift vector field of the
process generated by H0 (g, d ln ψ 2 ) since by Eqs. (47) and (55) this is
1 2
2 grad ln ψ and

H = Cψ ◦ H (g, dln ψ 2 ) ◦ Cψ−1 = −1/2g + V , (115)

where in the weak sense,


1 g ψ
V = (divg b + g(b, b)) = (116)
2 2ψ
1420 Rapoport

is the relativistic quantum potential ; here, in distinction with Bohm’s


quantum potential in non-relativistic Quantum Mechanics(45,68) (which is
retrieved in the case of n = 3 and g the Euclidean metric), it depends
on both the space and time-t coordinates. Then, we have proved that
−H (g, d ln ψ 2 ) is unitarily equivalent to the Hamiltonian operator H :=
− 21 g + V defined on L2 (volg ) and ψ is a generalized groundstate eigen-
function of H with 0 eigenvalue. The non-linear dependence of V on the
invariant density introduced by ψ introduces non-local correlations on
the quantum system We shall see below that this dependence of V on
ψ is removed due to conformal invariance. This will establish that the
Schroedinger operator H has for quantum potential one-twelfth of the
Riemannian scalar metric and thus H coincides with the Riemannian con-
formal invariant wave operator considered in quantum gravity in curved
spaces (see de Witt, Birrell and Davies(31) ). We shall now elaborate on
these aspects.

5. THE MEAN CURVATURE EXTREMAL PRINCIPLE

Since at the level of constitutive equations for Q, the electromag-


netic potentials decouple from the ψ-field (see the discussion that lead
to Eq. (56)) we can study independently the field equations from which
the RCW connection with exact Q can be derived. We shall assume that
n = 4. We start with a general Riemann–Cartan connection (αab ), (where
Greek letters denote space–time indices as until now, and Latin letters
denote anholonomic indices), and we introduce its scalar curvature
β
R() = eaα eb Rαβ
..ab
, (117)

where the eaα is a field of invertible tetrads with gαβ = δab eαa eβb , with δab
the Euclidean metric,21 and Rαβ
..ab is the curvature tensor of ( ab ).(18) Now
α

21 All the following definitions of the λ transformations and the ensuing field equations
are valid as well if we take here the Minkowski metric; since we do not know whether
our construction of a relativistic Brownian motion carries from the Minkowski space
to general Lorentzian metrics , in this section we shall keep the metric to be positive-
definite for which we take the initial metric to be Euclidean. Note added in proofs:
Brownian motions for general Lorentzian metrics have been recently constructed on
the unit tangent manifold (see J. Franchi and Y. Le Jan, Relativistic Diffusions, ar-
Xiv:math.PR/0403499 and arxiv:math.PR/0414085); this restriction is placed by normali-
zation of velocities by the velocity of light, as kindly explained to the author by Prof.
R. Rebolledo, to whom we express our gratitude. The relation of this construction, with
the Lorentz-invariant Brownian motions on Minkowski space presented in Ref. 53 and
the present article is unknown.
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1421

we recall the Einstein’s λ transformations of above (here ρ will be substi-


tuted by a scalar field φ): Let φ be a real function on M. Then λ(αb a ) :=
a α −1
αb , and λ(ea ) := φ ea so that λ(gαβ ) = φ gαβ and then the scalar curva-
α 2

ture transforms as λ(R()) = φ −2 R(), and finally volλ(g) = φ 4 volg . Since


the scalar fields ψ transform as λ(ψ) = φ −1 ψ, we get that the functional

A(, ψ, g) = R()ψ 2 volg (118)

is invariant by the set of λ transformations, i.e.: A(λ(), λ(ψ), λ(g)) =


A(, ψ, g). Notice that if from the field equations we obtain that ψ 2 volg
can be identified with the unique invariant density of the diffusion pro-
cess generated by H0 (g, d ln ψ 2 ), then (118) is the mean Riemann–Cartan
scalar curvature.(69) Taking variations with respect to g we obtain that

Rαβ () − 1/2gαβ R() = 0, (119)

i.e. the Einstein–Cartan equations for  in the vacuum, while by taking


γ
variations with respect to αβ , we obtain that similarly to Eq. (94) we have
γ γ
Tαβ = δαγ ∂β ln ψ − δβ ∂γ ln ψ, (120)

so that, up to factor of 3 which we shall absorb so we shall take Q =


dln ψ and thus the field equations have yielded a RCW structure with
exact Q. Taking variations with respect to ψ we get the teleparallelism:
R() = 0; replacing Eq. (120) in Eq. (119) we get the field for the Ein-
stein metric tensor Gαβ (g) = Rαβ (g) − 21 R(g):

6 1
Gαβ (g) = − 2
∂α ψ ∂β ψ − 1/2gαβ ∂γ ψ∂ γ ψ − (∇α ∇β ψ 2 − gαβ g ψ 2 ),
ψ 6
(121)

where in the r.h.s. we identify (up to a factor) minus the improved energy-
momentum density of the scalar field in renormalizable gauge theories.(70) .
Now, by taking the trace in this equation we finally get

1
Hψ ≡ (g − R(g))ψ = 0, (122)
6
so that ψ is a generalized groundstate of the conformal invariant wave oper-
ator defined on L2 (volg ). Note that from Eqs. (115), (116), (121) we con-
clude that the quantum potential is 1/12R(g) which does not depend on
the scalar field ψ at all. Therefore, the correlations on the quantum system
1422 Rapoport

under Brownian motion with drift given by b = grad lnψ are mediated by
the metric scalar curvature (which, of course, does not depend on ψ any
more; this is the form invariance of the quantum potential(45) )! Otherwise
stated and in view of the relation between the noise tensor and the Rie-
mannian metric (see the discussion after Eq. (48)), when we have an aniso-
tropic noise tensor we have constructed a non-trivial metric and quantum
non-local correlations which are due to the metric scalar curvature.
Solving the conformal invariant wave equation with Dirichlet regu-
larity conditions on the closure of an open neighborhood of M,(63) we
obtain a conformally conjugate Dirichlet form whose associated Hamil-
tonian operator is −H0 (g, dln ψ 2 ), with ψ a solution of Eq. (122) and
thus the Markovian semigroup determined by it can be reconstructed by
reversing the steps in the previous Section. We shall finally establish the
relation between the heat kernel pconf (τ, x, y) of the Markovian semigroup
exp( τ2 H ) and the heat kernel pψ (τ, x, y) of the RCW semigroup. We have

τ
exp(τ H0 (g, dln ψ 2 ))f (x) = ψ −1 (x) exp H (ψf )(x)
2
= ψ −1 (x)pconf (τ, x, y)φ(y)f (y)volg (y)
(123)

so that we conclude that

pψ (τ, x, y) = ψ −1 (x)ψ(y)pconf (τ, x, y). (124)

Thus, we have linked the kernels of the quantization in the two Hil-
bert spaces, the groundstate Hilbert space L2 (ψ 2 volg ), and L2 (volg ). The
former corresponds to the RCW geometry, while the latter is the usual
Hilbert space for the quantization of the kinetic energy of a spinless mas-
sive free-falling test-particle, in terms of the Riemannian invariants of the
manifold M described in terms of g. We remark that the introduction of
both spaces and the unitary transformation between them, has allowed us
to identify the quantum potential, while working only in the usual Hil-
bert space would not have allowed for this identification; finally, the scalar
curvature term so much discussed has been found to be a resultant of
the λ invariance of the theory, and not the resultant of technicalities in
computing the propagators(20,45) ; as discusssed already, this theory has no
ordering problem . Thus, in the L2 (volg ) space we have found the Ham-
iltonian operator considered by B.de Witt, and reencountered by several
researchers in quantum field theory in Riemannian geometries through the
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1423

short-τ expansion of pconf (τ, x, x)(20) in geometrical and topological invar-


iants , and for the path integral representations for Fokker–Planck opera-
tors(27) , which as we already saw, when g is Riemannian, are precisely of
the form H0 (g, Q). Yet, our result is in disagreement with the path inte-
gral representation of the classical kinetic energy of a massive particle in a
Riemann–Cartan geometry due to Kleinert, in which he obtains twice the
quantum potential (see, chap. X,(45) ).22 We would like to finally comment
that the relation between a Klein-Gordon field and the drift vector field,
is implicit in a covariant formulation (in the usual Riemannian setting) of
space–time fluctuations due to Graham(59) .

6. SUPERSYMMETRY AND EXACT RCW CONNECTIONS

In this last section we shall proceed to construct a supersymmetric


system for the case of RCW with Q reduced to the exact term d ln ψ pre-
viously encountered; thus, g is a Riemannian metric, and M is further pro-
vided with a RCW connection with Q = d ln ψ. The approach we shall
follow next is the one which has been at the foundations of supersymmet-
ric systems, due to Witten(71) . This is the approach which has proved to be
extremely fruitful, both in theoretical physics and in mathematical-physics
where it has lead to the proof of the Morse inequalities, the Gauss-Bon-
net-Chern theorem of differential geometry, to the proof of the index the-
orem for Dirac operators and fundamentally to the study of many-body
quantum mechanical systems(32) . We recall the basic definition due to Wit-
ten. Let H be a Hamiltonian operator on a Hilbert space H, together
with a self-adjoint operator Q and a bounded self-adjoint operator P both
defined on H , such that

H = Q2  0, P 2 = 1, and {Q, P} = QP + PQ = 0. (125)

Then, the triple {H, P, Q} is said to be a supersymmetric system, or, to


have supersymmetry. Since P is self-adjoint and P 2 = 1, then P has ei-

22 For a discussion on the work of Kleinert and the role autoparallels we suggest the
reader to return to Remarks 1 and footnote no. 8 before Section 5 of our accompa-
nying article(1) . Another role for autoparallels appears on the formulation of a geo-
metrically invariant theory of non-linear non-equilibrium thermodynamics systems in the
framework developed in this article and subsequently in Refs. 33, 46 and 72, in which
the approach to equilibrium of these systems can be related to the autoparallels of a
RCW connection, where as we have proved already, the spin structure is built-in; we
shall elaborate this elsewhere.
1424 Rapoport

genvalues 1 and −1. Denote

Hferm = { ∈ H : P = −}, Hbos = { ∈ H : P = }, (126)

which are called the fermionic and bosonic states, respectively. Then, Q :
Hbos → Hferm and Q : Hferm → Hbos , and thus Q maps boson-
ic to fermionic states, and conversely. In the present theory, we take for
Hilbert space H the space given by the ncompletion of the pre-Hilbert
space defined by the multiforms A = k=0 A k , which are square-inte-
grable with respect to the following pairing (that extends the pairing dealt
with when treating the energy-forms and the ground-state representation
for quantum mechanics)

(( A k , Bk )) := ⊗k g −1 ( A k , B k )(x)ψ 2 (x)volg (x); (127)

for A k , B k ∈ sec(k (T ∗ M)) ⊂ sec(Cl (M, g)) The Hamiltonian operator


is the laplacian on multiforms of the exact RCW connection (see Eqs. (26)
and (27)):

1  n
H := H (g, d ln ψ 2 ) = ( + Lgrad ln ψ 2 ) = Hk (g, d ln ψ 2 ). (128)
2
k=0

We note that H (g, d ln ψ 2 ) preserves graduation on sec(Cl (M, g)), i.e.


k-forms are mapped into k-forms. Furthermore,
 
1
Q := √ (d − [δ − igrad ln ψ 2 ]), (129)
2

where iX denotes the right interior product with respect to the vectorfield
X on M, and P is defined by its restriction on H to k-forms:


n
P A k = (−1)k , for any A= A k ∈ H (130)
k=0

Then, it is easily verified that we have a supersymmetric system. Thus, in this


setting, fermionic (bosonic) states are given by odd (even) multiforms, but
we remark that here the inner product is defined in terms of the invariant
density ψ 2 volg . We remark that when g is Euclidean metric, H essentially
reduces to H0 (g, d ln ψ 2 ) acting componentwise. If we would introduce the
electromagnetic potential terms in the de Rham–Hodge decomposition of
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1425

Q, we shall see that Q2 will not longer be equal to H (g, Q). In a more gen-
eral setting we would take for H the laplacian produced by the square of the
Dirac operator defined by a metric-compatible Riemann–Cartan connection
(D∇ )2 , where of course Q = D∇ is the Dirac operator of this connection
(see Sec. 2.1) and P as before (recall, we computed the square of the Dirac
operator for scalar fields, and took for Hamiltonian a natural extension, not
the square of the Dirac operator acting on multiforms; as discussed then,
this Hamiltonian operator is not graduation preserving, which is precisely
what one obtains in adding the electromagnetic potential terms in the defi-
nition of Q and taking its square).

7. FINAL COMMENTS

This article has treated several theories in a single framework, in which


geometrical structures originate random mechanics, and viceversa, they can
alternatively be seen as originated by them. We have seen that in the case of
the equations of Dirac-Hestenes, the Cartan–Weyl connection is related to
the appearance of the spinor fields, in fact that all fields can be introduced
as Einstein’s lambda transformations in the vacuum. On the electromagnetic
potential 1-forms, they act as a gauge transformation, so that in this the-
ory, everything can be defined in terms of local tetrad and local rotational
fields,23 the flat Minkowski and Euclidean metrics, scalar fields acting on
them to produce non-trivial structures, and last but not least, the Wiener
process, whose amplification by the noise tensor defined from a square-root
of the metric is the single random input of the theory. Many researchers
have pointed out to the vacuum oscillations these unique random feature
(the ‘apeiron’), which as we have seen, is all pervasive in the domain of think-
ing of theoretical physics, appearing in fluid-dynamics, quantum mechanics
(in which the kinematical viscosity is substituted by –h/mc) and as the con-
stituents of the gravitational field, whenever the metric is non-flat. So, in this
article we have furthered a minimalist approach that stems mostly from the
non-triviality of the vacuum upon being acted by scale transformations. We
can go one step forward to treat chaos, turbulence and irreversibility, fol-
lowing Kiehn(20) in applying a topological approach based on the Pfaffian
dimension and Frobenius integrability to the Pfaffian 1-form Q, which holds

23 In fact our treatment of the equivalence between Maxwell and Dirac-Hestenes equation,
has substantiated the idea that spin and torsion are associated, yet in distinction with
the usual approach, here the trace-torsion is what matters, since the  representative for
the DHSF yields all the terms of the Cartan–Weyl form Q, and equivalently, solving the
non-linear DH equation or its electromagnetic equivalent permits the construction of Q.
1426 Rapoport

the interaction terms of the theory; the analysis presented above of the form
of Q is by itself topological, since it originated from the de Rham–Kodaira–
Hodge decomposition, so we can make full use of the geometrical approach
with a topological one which we would like to suggest that may lead us to a
more complete picture, in which the role of discontinuities will be explicited.
Furthermore, this will lead to the construction of a theory of the formation
and evolution of coherent structures in diffusion processes, which we shall
present elsewhere. It is an objective of this program to elaborate further the
relation between these geometries and their random counterparts in physics,
notably in fluid dynamics, which as we mentioned in our previous article, can
be taken as the foundations to actually derive electrodynamics (as a setting
for the formulation of a theory of turbulence) which in this article, as we
have further indicated, has some puzzling relations with relativistic quan-
tum mechanics. The present approach has intended to follow a tradition
of non-dualistic thinking in physics, of a Universe in which chance cannot
be conceived as an ad hoc entity (see Zeldovich’s discussion on chance(73)
and the classics, Stratonovich and Prigogine and collaborators,(73) ) or an
expression of our finite capabilities to describe many degrees of freedom,
and which we know today from the study of simple unidimensional recur-
sively defined dynamical systems to be unavoidable. To this vision, Einstein
himself gave the initial description of Brownian motions, at the same time
as he presented his special relativity. The epistemological implications of the
present descriptions are far from the author’s capacity to review them (in
particular, if it amounts to resume them in a few lines),24 yet we have some-
what placed it in the line of thought by Einstein (as a founding father of
the statistical approach to theoretical physics as well as the founding of the
theory of general relativity), Clifford’s vision of the status of geometry with
regards to physical fields, de Broglie and Bohm’s vision of an implicate order,
and still Prigogine’s vision of chance as a driving constitutive force which in
Biology took the form of the theory of evolution due to Charles Darwin.

ACKNOWLEDGEMENTS

The author would like to express his deep gratitude to the Organizing
Committee of IARD 2004, for their kind invitation to submit a contribution

24 The present work is further related to the approach to physics and science that stemmed
from the geometrization of the work by Fisher on statistical inference.(74) For a most
interesting discussion of the relation between the so-called Fisher information, recursivity
and the ideas of Baruch Spinoza and John A.Wheeler on the Universe as a cognitive
system, we recommend the article by Frieden in Ref. 74.
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1427

to the Proceedings. Particularly, our gratitude to Prof.Larry Horwitz for dis-


cussions and instructing the author on his work, his kind interest on ours and
for sending preprints of his articles; to the referees for pointing out the need
for some clarifications and last, but not least, to the editors for their kind indi-
cations. Our gratitude as well to Prof. Leopold Halpern for his kind support
along the years, and to Prof. W. Rodrigues, for sending reprints of his work.

REFERENCES

1. D. Rapoport, Found. phys. 35(7), (2005).


2. D. Rapoport, Rep. Math. Phys. 49(1), 1–27, (2002); ibid. Rand. Operts. Stoch.Eqs.
11(2), (2003); ibid. Discrete and Cont. Dyn. Syst. A special issue 2000, Procs. III’d
Inter. Conf. Dyn.Syst. Diff. Eqts., Atlanta, 2000, S.Hu ed. ibid. Rep. Math. Phys. 50(2),
211–250 (1992).
3. D. Rapoport, Random Operts. Stoch. Eqs. 11(2), 359–380 (2003); ibid. in Trends in
Partial Differential Equations on Mathematical Physics, in Honor of Prof. V.A. Solonni-
kov, Obidos (Portugal), May 2004, Progress in Nonlinear Differential Equations and
Their Applications, 61, 225–241, J. F. Rodrigues et al. eds., (Birkhauser, Boston, 2004).
4. D. Hestenes and G. Sobczyck, Clifford Calculus to Geometric Calculus (D.Reidel,
Dordrecht, 1984).
5. M. Pavsic, The Landscape of Theoretical Physics: A global view (Kluwer, Dordrecht
2001); ibid., arXiv:gr-qc/0111092 v4.
6. W. A. Rodrigues Jr., J. Math. Phys. 45, 2908–2944 (2004); R. Mosna and W. A. Ro-
drigues Jr., J.Math.Phys. 45, 2945–2966 (2004).
7. D. Hestenes, J. Math. Phys. 8, 798–808 (1975); 16, 556–571 (1975); 14 (1973), 893–
905; 15 (1974), 1768–1777 (1974); 15 (1974), 1778–1786; 16 (1975), 556–572; Found.
Phys.15, 63–87 (1985); Found. Phys. 12, 153–168 (1982); Int. J. Theor. Phys. 25, 1013–
1028 (1986).
8. D. Hestenes, Found.Phys.20, 1213–1332 (1990); ibid. in D. Hestenes and A.Weingart-
shofer eds,The Electron pp. 21–26, (Kluwer, Dordrecht, 1991).
9. W. Rodrigues and Q. de Souza, in, Gravitation, The Space Time Structure, Proceedings,
Silarg VIII, 1994, W. Rodrigues et al. eds., pp. 170–210 (World Scientific, Singapore,
1995).
10. A. M. Moya, V. V. Fernandez and W. A. Rodrigues Jr., Metric Clifford algebras, Adv.
Appl. Clifford Alg. 11(S3), 53–73 (2001).
11. W. A. Rodrigues J., Q. A. G. de Souza and P. Lounesto, Int. J. Theor. Phys. 35, 1854–
1900 (1995).
12. Th. Frankel, The Geometry of Physics, An Introduction (Cambridge University Press,
Cambridge, 1997); G. de Rham, Differentiable Manifolds (Springer, Berlin, 1984);
F.Warner,Introduction to Differentiable Manifolds and Lie Groups (Holden Day, San
Francisco,1971).
13. H. L. Cycon, R. G. Froese, W. Kirsch and B. Simon, Schroedinger Operators with Appli-
cations to Quantum Mechanics and Global Geometry (Springer, Berlin, 1987); G. Junker,
Supersymmetric Methods in Quantum and Statistical Physics (Springer, Berlin, 1996).
14. D. Rapoport, Int. J.Theor. Phys. 30(11), 1497 (1991).
15. S. Hawking, in General Relativity, an Einstein Centenary Survey, S. Hawking and
W. Israel (Cambridge University Press, Cambridge 1979); C. Itzykson and J. M.
1428 Rapoport

Drouffe, eds. Statistical Field Theory, vol. I (From Brownian Motion to Renormaliza-
tion, and Lattice Gauge Theories), (Cambridge University Press, Cambridge 1989)
16. D. Hurley and M. Vandyck, Geometry, Spinors and Applications (Springer, Berlin, 1999).
17. H. Weyl, Space, Time and Matter (Dover, New York, 1952).
18. V. de Sabbata and C. Sivaram, Spin and Torsion in Gravitation (World Scientific, 1994);
F. Hehl, P. von der Heyde, G. D. Kerlick and J. M. Nester, Rev. Modern Phys., 48,
3 (1976); F. Hehl, J. Dermott McCrea, E. Mielke and Y. Ne’eman, Phys. Reports vol.
258, 1–157 (1995).
19. C. Misner, K.Thorpe and J. A. Wheeler Gravitation (Freeman, New York, 1973); A. S.
Eddington, The Mathematical Theory of Relativity (Chelsea, London, 1995) (reedited).
20. R. M. Kiehn, A topological perspective of electromagnetism, in http://www.car-
tan.pair.com.
21. G. F. Rubilar, Y. N. obukhov and F. W. Hehl, Int. J. Mod. Phys. D. 11, 1227; F. W.
Hehl and Yu. N. Obukhov, Foundations on Classical Electrodynamics: Charge, Flux,
and Metric (Birkhauser, Boston, MA, 2003); Le Bellac and J. M. Levy-Leblond, Nu-
ovo Cimento 143(2), 217–233 (1972).
22. E. J. Post, Formal Structure of Electromagnetics, reprinted (Dover, New York, 1997).
23. W. Rodrigues and J. Lu, Found.Phys. 27, 435–508 (1997); W. Rodrigues and E. C.
Oliveira, Ann. der Physik 7, 654–651 (1998); W.Rodrigues and J.E. Maiorino, Sci. Tech.
Mag. 2(4), 1–167 (1999).
24. V. A. Fock, Theory of Space, Time, and Gravitation (Pergamon Press, London, 1964).
25. E. Cartan, The Theory of Spinors, reprinted (Dover, New York, 1996); A. Lasenby,
C. Doran and S. Gull, in Spinors, Twistors, Clifford Algebras and Quantum Deforma-
tions, pp. 233–245, Z. Oziewicz et al. eds. (Kluwer, Dordrecht, 1993).
26. D. Rapoport, W. Rodrigues, Q. de Souza and J. Vaz, Algebras, Groups and Geometries
11, 25–35 (1995).
27. A. Einstein and Kauffman, Annals Maths. 56 (1955); Yu Obukhov, Phys. Letts. 90 A,
13 (1982).
28. H. Marmanis, Phys. Fluids 10, (6), 1428 (1998).
29. E. C. Stueckelberg, Helv. Physica Acta 14, 322, 588 (1941); L. P. Horwitz and C.
Piron, Helv. Physics Acta 46, 316 (1973); L. P. Horwitz and C. Piron, Helv. Physi-
ca Acta 66, 694 (1993); M. C. Land, N. Shnerb and L. P. Horwitz, J. Math. Phys.
36, 3263 (1995); L. P. Horwitz and N. Shnerb, Found. of Phys. 28, 1509 (1998).
30. A. Kyprianidis, Phys. Rep. 155(1), 1–27 (1987) and references therein.
31. B. de Witt, “Quantum field theory in curved space–time,” Phys. Rep. C 19(6) (1975),
295–357. G. W. Gibbons, in General Relativity, An Einstein Centennary Survey, S. W.
Hawking and W. Israel (Cambridge University Press, Cambridge, 1979); N. D. Birrell
and P. C. W. Davies, Quantum Field Theories in Curved Space, (Cambridge University
Press, Cambridge 1982); J. Schwinger, Phys. Rev. 82, 664 (1951).
32. R. E. Collins and J. R. Fanchi, Nuovo Cimento A 48, 314 (1978); J. Fanchi, Found.
Phys. 30(8), 1161–1189 (2000) & 31(9), 1267–1285 (2001).
33. D. Rapoport, in Instabilities and Nonequilibrium Structures vol. VI, Proceedings of the
Sixth International Workshop, E. Tirapegui et al. eds. (Kluwer, Dordrecht, 2000); ibid.
in Proceedings of the International Workshops on the Frontiers of Mathematics, Phys-
ics and Biology, Monteroduni, Italy, August 1995, 2, G.Tsagas ed. (Hadronic Press and
Ukraine Academy of Sciences, Palm Harbor Florida-Kiev, 1996).
34. C. W. Gardiner, Handbook of Stochastic Processes, 2nd. ed. (Springer, Berlin, 1993);
Z. Schuss, Stochastic Differential Equations and its Applications (Academic Press, New
York, 1987); Risken, The Fokker–Planck Equation (Springer, Berlin, 1993).
35. B. Lavenda, Thermodynamics of Irreversible Processes (Dover, New York, 1973).
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1429

36. A. Sommerfeld, Mechanics of Deformable Bodies, Lectures on Theoretical Physics vol.


II (Academic Press, New York, 1964); Sir William Thomson (Lord Kelvin), Mathe-
matical and Physical Papers, vol. 3, art. 49, 50,52.
37. W. A. Rodrigues Jr., Int. J. Math. Math. Sci. 2003 2007–2734 (2003).
38. E. Witten, Monopoles and four-manifolds, Math. Res. Lett. 1, 769 (1994).
39. J. Stratton, Electromagnetic Theory (Mac Graw-Hill, New York 1941).
40. J. Fanchi, Parametrized Relativistic Quantum Theory (Kluwer, Dordrecht, 1993).
41. M. Trump and W. Schieve Classical Relativistic Many-Body Dynamics (Kluwer, Dordr-
echt, 1994).
42. I. Prigogine, From Being to Becoming (Freeman, New York, 1995).
43. F. Langouche, D. Roenkarts and E. Tirapegui, Functional Integration and Semiclassical
Expansions (Reidel, Dordrecht 1981).
44. D. Rapoport, Int. J. Theor. Phys. 36(10), 2115–2152 (1997).
45. D. Bohm, Phys. Rev. 85, (1952), 166 and 180; D. Bohm and J. P. Vigier, Phys. Rev.
96, 208 (1953); L. de Broglie, La réinterpretation de la Mécanique ondulatoire, (Gau-
thier-Villars, Paris, 1971); ibid. Etude critique des bases de l’interpretation actuelle de
la Mécanique ondulatoire (Gauthier-Villars, Paris, 1963); ibid. Jalons pour une nouvelle
microphysique (Gauthier-Villars, Paris 1978).
46. D. Rapoport, in Instabilities and Nonequilibrium Structures vol. IX, Proceedings of the
Ninth International Workshop, O. Descalzi et al. eds. (Kluwer, Boston, 2003).
47. S. Hojman, M. P. Rosenbaum, L. C. Shepley, Phys. Rev. D17, 1341 (1978); F. Hehl
and Yu. Obukhov, arXiv:gr-qc/0001010 v2 (3 may 2000.)
48. A. N. Kolmogorov, Zur Umkehrbarkeit der statistichen Naturgesetze, Math. Annalen
113, 776–772 (1937).
49. H. H. Sallhoffer, Z. Naturforsch. 33a, 1378 (1978); 45a (1990), 1361; ibid. in Essays
on the Formal Aspects of Electromagnetic Theory, 268–286, A. Lakhtakia ed. (World
Scientific, Singapore 1993); A. A. Campolattaro, Int. J. Theor. Phys. 19 (1980), 99, 19;
127 (1980); 29, 141 (1990).
50. J. Vaz Jr. and W. A. Rodrigues, Int. J. Theor. Phys. 32, 945–949 (1995); W. Rodrigues
and J. Vaz, in R.Delanghe ed. Clifford Algebras and their Applications in Mathematical
Physics, Proceedings of the III Workshop (Kluwer, Dordrecht, 1993); W. Rodrigues, J.
Vaz and E. Recami (1993b), in Courants, Amers, Écueils en Microphysique, Centennial
Celebration of L. de Broglie, Annales Fondation L. de Broglie, Paris (special issue)
(1993); G. Lochak, Int. J. Theor. Phys. 24, 1019 (1985).
51. D.Rapoport, Adv.Appl.Cliff.Alg. 8(1), 129–146 (1998); ibid. in Group XXI, Physical
Applications and Mathematical Aspects of Algebras, Groups and Geometries, Proceedings,
Clausthal, 1996, H. Doebner et al. eds. (World Scientific, Singapore, 1997); D.Rapoport
& M. Tilli, Hadronic J. Suppl., 2 (2) 682 (1986).
52. G. Lochak,Int. J. Theor. Phys. 24, 1019 (1985); C. Daviau, Ann. Fond. L. de Broglie
14, 373 (1989); C. Daviau and G. Lochak, Ann. Fond. L. de Broglie 16, 43 (1991).
53. O.Oron and L. Horwitz, Relativistic Brownian Motion as an Eikonal Approximation
to a Quantum Evolution Equation, Found. Phys. IARD 2004, special issue; ibid. in
Progress in General Relativity and Quantum Cosmology Research, V. Dvoeglazov ed.
(Nova Science, Hauppage, 2004); ibid., Phys. Letts. A280, (2001), 265.
54. J. P. Zambrini, Phys. Rev. A33, 1532–1548 (1986) and A 35, 3631–3649 (1987); K.
L. Chung and J. P. Zambrini, Introduction to Random Time and Quantum Random-
ness, 2nd. ed., World Scientific, Singpore (2003); A. B. Cruzeiro, Wu Liming and J. P.
Zambrini, in Stochastic Analysis and Mathematical Physics, ANESTOC ’98 (Santiago,
Chile), R. Rebolledo ed. (Birkhauser, Boston, 2000).
55. M. Nagasawa, Quantum Theory and Brownian motions (Birkhauser, Boston, 1999).
1430 Rapoport

56. M. Serva, Annals Inst. H. Poincaré, Phys. Theor. 49, 312 (1998); R. Marra and M.
Serva, Annals Inst. H. Poincaré Phys. Theor. 53, (1), 97–108 (1990).
57. E. Nelson, The theory of Brownian Motion (Princeton University Press, Princeton (New
Jersey), 1967); ibid. Quantum Fluctuations (Princeton University Press, NJ 1985).
58. P. Baxendale, K.D. Elworthy, and Z.Wahrschein. verw. 65, 245. (1983) K. Kunita, Sto-
chastic Flows and Stochastic Differential Equations, (Cambridge University Press, Cam-
bridge 1994); N. Ikeda and S.Watanabe, Stochastic Differential Equations and diffusion
Processes (North-Holland-Kodansha, Amsterdam-Tokyo, 1989).
59. R. Graham, “Lagrangian for diffusions in curved space–time”, Phys. Rev. Letts. 38(2),
51 (1977).
60. E. Schroedinger Sitzunsberger Press Akad. Wiss. Math. Phys. Math., 144 (1931). Ann.
I. H. Poincaré 11, 300 (1932).
61. M. Pavsic, E. Recami, W. A. Rodrigues Jr., G. D. Macarrone and G. Salesi, Phys.
Lett. B318, 481–488 (1993).
62. H. Kleinert, Path integrals in Quantum Mechanics, Statistics and Polymer Physics
(World Scientific, Singapore, 1991).
63. E.B. Davies, Heat kernels and Spectral Theory (Cambridge University Press, Cam-
bridge, 1989).
64. S. Albeverio et al., J. Math. Phys. 18, 907 (1977) and Stochastic Methods in Physics,
Math. Phys. Rep. 77, in K. D. Elworthy and de C. Witt-Morette, eds. no. 3, (1977).
See also F. Guerra contribution in the last reference.
65. Fukushima, Markov Processes and Dirichlet forms (North-Holland, Amsterdam, 1981).
66. H. Araki, J. Math. Phys. 11, 492 (1960). A.I. Kirillov, Theor. Math. Phys., 345–353,
447–453 (1991).
67. M. Reed and B. Simon, Modern of Modern Mathematical Physics II, Fourier Analysis,
Self-adjointness (Academic Press, New York, 1975).
68. P. R. Holland, The quantum theory of motion (Cambridge Univestity Press, Cambridge
U.K., 1994).
69. D. Rapoport, in Gravitation, The space–time Structure, W. Rodrigues et al. eds. Sin-
gapore, 1995; ibid. in Chaos and Dyn. Systems II, Proc. Conf. Dynamical Syst. and
Chaos, Tokyo 1994, Y. Aizawa ed. (World Scientific, Singapore, 1995).
70. C. G. Callan, S. Coleman and R. Jackiw, Annals Phys. 59, 42 (1970).
71. E. Witten, J. Diff. Geom. 17, 661 (1982).
72. D. Rapoport, in Proceedings, International Conference on Dynamical Systems and
Chaos, Tokyo, May 1994, Y. Aizawa et al. eds. vol. 2, (World Scientific, Singapore,
1995).
73. Ya. B. Zeldovich, A. A. Rumauzkin and D.D.Sokoloff, The Almighty Chance (World
Scientific, Singpore 1990). R. Stratonovich, Non-linear non-equilibrium Thermodynam-
ics, I, II, (Springer, Berlin, 1992, 1994). I. Prigogine, Introduction to Thermodynamics
of Irreversible Processes (Thomas, Springfield, 1955). G. Nicolis and I. Prigogine, Self-
organization in Non-equilibrium Systems (Wiley, New York, 1977).
74. B. R. Frieden, Physics from the Fisher Information, (Cambridge University Press, Cam-
bridge, 1999); www.optics.arizona.edu/Fisher/; R.A. Fisher, Phil. Trans. R. Soc. London
222, 309, (1922).
75. E. B. Davies, Heat Kernels and Spectral Theory (Cambridge University Press, Cam-
bridge, 1990).
76. D. Rapoport, in Proc. IX th. Marcel Grossman Meeting, Rome, June 2000, R. Ruffini
et al. eds. (World Scientific, Singapore, 2003); ibid. ibid. Adv. Appl. Clifford Alg. 8(1),
127–169, 1998.
77. D. Ebin and J.Marsden, Annals Math. 92, 102–163 (1971).
Cartan–Weyl Dirac and Laplacian Operators, Brownian Motions 1431

78. D. Rapoport and S. Sternberg, Annals Phys. 158, 447 (1984); ibid, Lett. N. Cimento
80A, 371 (1984).
79. A. Lasota and M. Mackey, Probabilistic Properties of Dynamical Systems (Cambridge
University Press, Cambridge, 1985).
80. S. Sternberg, Annals Phys. 162, 85 (1985); J.M. Souriau, Annales I.H. Poincaré 20
A(1974).
81. S. Gupta, Proc. Royal Soc. 63A (1950), 681; K. Bleuler, Phys. Helv. Acta 23, 567
(1950).
82. W. Pezzaglia Jr., gr-qc/9704048
83. T. W. Marshall, Physica 103 A, 172 (1980).

Potrebbero piacerti anche