Вы находитесь на странице: 1из 23

a

r
X
i
v
:
1
3
1
2
.
6
2
4
7
v
1


[
h
e
p
-
t
h
]


2
1

D
e
c

2
0
1
3
Lagrangian for Frenkel electron and positions non-commutativity due to spin.
Alexei A. Deriglazov

and Andrey M. Pupasov-Maksimov

Depto. de Matem atica, ICE,


Universidade Federal de Juiz de Fora, MG, Brasil
(Dated: December 24, 2013)
We construct relativistic-invariant spinning-particle Lagrangian without auxiliary variables. Spin
is considered as a composed quantity constructed on the base of non-Grassmann vector-like variable.
The variational problem guarantees both xed value of spin and Frenkel condition on spin-tensor.
Taking into account the Frenkel condition, we obtain, inevitably, relativistic corrections to the alge-
bra of position variables: their classical brackets became noncommutative, with the parameter of
non-commutativity proportional to the spin-tensor. This leads to a number of interesting conse-
quences in quantum theory. We construct the relativistic quantum mechanics in canonical formalism
(in physical-time parametrization) and in covariant formalism (in arbitrary parametrization). We
show how state-vectors and operators of covariant formulation can be used to compute mean values
of physical operators of position and spin. This proves relativistic covariance of canonical formalism.
Various candidates for position and spin operators of an electron acquire clear meaning and interpre-
tation in the Lagrangian model of Frenkel electron. We also establish the relation between Frenkel
electron and positive-energy sector of Dirac equation, this allowed us to turn to the long-standing
problem on spin and position operators of Dirac theory. Contrary to widely assumed opinion, our
results argue in favor of Pryces (d)-type operators. This implies that eects of non-commutativity
could be presented at the Compton wave length, in contrast to conventional expectations at the
Planck length. At last, we present the manifestly covariant form of spin and position operators of
Dirac equation.
I. INTRODUCTION AND OUTLOOK
Quantum description of spin is based on Dirac equation, whereas the most popular classical equations of electron
has been formulated by Frenkel [1, 2] and Bargmann, Michel and Telegdi (F-BMT) [3]. They almost exactly reproduce
spin dynamics of polarized beams in uniform elds, and agrees with the calculations based on Dirac theory. Hence
we expect that these models might be proper classical analog for the Dirac theory. The variational formulation
for F-BMT equations represents rather non trivial problem [414] (note that one needs a Hamiltonian to study, for
instance, Zeeman eect). In this work we continue systematic analysis of these equations started in [18]. We develop
their Lagrangian formulation considering spin as a composed quantity (inner angular momentum) constructed from
non-Grassmann vector-like variable and its conjugated momentum [1017].
Non relativistic spinning particle with reasonable properties can be constructed [15, 18] starting from singular
Lagrangian which implies the following Diracs constraints

2
a
3
= 0,
2
a
4
= 0, = 0, where a
3
=
3
2
4a
4
, (1)
while relativistic form of these constraints read
T
3
=
2
a
3
= 0 , T
4
=
2
a
4
= 0 , T
5
= = 0 , (2)
T
6
= p = 0 , T
7
= p = 0 . (3)
Besides, we have the standard mass-shell constraint in position sector, T
1
= p
2
+ (mc)
2
= 0. We denoted the basic
variables of spin by

= (
0
, ), = (
1
,
2
,
3
), then =
0

0
+ and so on.

and p

are conjugate
momenta for

and the position x

.
Since the constraints are written for the phase-space variables, it is easy to construct the corresponding action
functional in Hamiltonian formulation. We simply take L
H
= p x + H, with Hamiltonian in the form of linear
combination the constraints T
i
multiplied by auxiliary variables g
i
, i = 1, 3, 4, 5, 6, 7. The Hamiltonian action with six

Electronic address: alexei.deriglazov@ufjf.edu.br

Electronic address: pupasov@phys.tsu.ru


2
auxiliary variables admits interaction with an arbitrary electromagnetic eld and gives unied variational formulation
of both Frenkel and BMT equations, see [14]. In section II we develop Lagrangian formulation of these equations.
Excluding conjugate momenta from L
H
, we obtain the Lagrangian action. Further, excluding the auxiliary variables,
one after another, we obtain various equivalent formulations of the model. We shortly discuss all them, as they will
be useful when we switch on interaction with external elds [19, 20]. At the end, we get the minimal formulation
without auxiliary variables. This reads
S =
_
d mc

xN x +

a
3

N
1
2
g
4
(
2
a
4
), (4)
where N

2
is projector on the plane transverse to the direction of

. The last term in (4) represents


kinematic (velocity-independent) constraint which is well known from classical mechanics. So, we might follow the
classical-mechanics prescription to exclude this term as well. But this would lead to lose of manifest relativistic
invariance of the formalism. The action is written in a parametrization which obeys
dt
d
> 0, this implies g
1
() > 0, p
0
> 0. (5)
To explain this restriction, we note that in absence of spin we expect an action of a spinless particle. Switching o
the spin variables

from Eq. (4), we obtain L = mc

x
2
. Let us compare this with spinless particle interacting
with electromagnetic eld. In terms of physical variables x(t) this reads L = mc

c
2
x
2
eA
0

e
c
A x. If we
restrict ourselves to the class of increasing parameterizations of the world-line, this reads L = mc

x
2

e
c
A x, in
correspondence with spinless limit of (4).
Assuming
dt
d
< 0 we arrive at another Lagrangian, L = mc

x
2

e
c
A

. So a variational formulation with both


positive and negative parameterizations would describe simultaneously two classical theories. In quantum theory they
correspond to positive and negative energy solutions of Klein-Gordon equation [21].
In [18] we discussed the geometry behind the constraints (1)-(3). The phase-space surface (1) can be identied
with group manifold SO(3). It has natural structure of ber bundle with the base being a two-dimensional sphere.
Components of non relativistic spin-vector are dened by S
i
=
ijk

k
. At the end, they turn out to be functions of
coordinates which parameterize the base. The set (2), (3) is just a Lorentz-covariant form of the constraints (1). In
the covariant formulation, S
i
is included into the antisymmetric spin-tensor J

= 2
[

]
according to the Frenkel
rule, J
ij
= 2
ijk
S
k
.
The constraints have clear meaning in the dynamical theory too.
The role of pure spin-sector constraints (2) is a two-fold. First, they x the value of spin, J

= 6
2
. As in the
rest frame we have S
2
=
1
8
J

=
3
2
4
, this implies the right value of three-dimensional spin, as well as the right
number of spin degrees of freedom.
Second, the equation
2
a
3
= 0 is the rst-class constraint related with local symmetry of variational problem.
The spin-plane symmetry has clear geometric interpretation as transformations of structure group of the ber bundle
acting independently at each instance of time. They rotate the pair

in the plane formed by these vectors. In


contrast, J

turns out to be invariant under the symmetry. Hence the spin-plane symmetry determines physical
sector of the spinning particle: the basic variable

is gauge non-invariant, so does not represent an observable


quantity, while J

does.
Reparametrization symmetry is known to be crucial for Lorentz-covariant description of a spinless particle. The
spin-plane symmetry, as it determines physical sector, turns out to be crucial for description of a spinning particle.
We point out that this appears already in non relativistic model (
2
a
3
= 0 represents the rst-class constraint in
the set (1)). The local-symmetry group of minimal action will be discussed in some details in subsection II C. Curious
property here is that reparametrization symmetry turns out to be combination of two independent local symmetries.
Equations (3) guarantee the Frenkel-type condition J

= 0. They form a pair of second-class constraints which


involve both spin-sector and position-sector variables. This leads to new properties as compared with non relativistic
formulation. The second-class constraints must be taken into account by transition from Poisson to Dirac bracket.
As the constraints involve conjugate momenta p

for x

, this leads to nonvanishing Dirac brackets for the position


variables
{x

, x

}
D
=
J

2p
2
. (6)
We can pass from the parametric x

() to physical variables x
i
(t). They also obey a noncommutative algebra, see
Eq. (39) below. We remind that in a theory with second-class constraints one can nd special coordinates on the
constraints surface with canonical (that is Poisson) bracket, see (46). Functions of special coordinates are candidates
3
for observable quantities. The Dirac bracket (more exactly, its nondegenerated part) is just the canonical bracket
rewritten in terms of initial coordinates [23]. For the present case, namely the initial coordinates (they are x
i
(t)),
are of physical interest
1
, as they represent the position of a particle. So, while there are special coordinates with
canonical symplectic structure, the physically interesting coordinates obey the non-commutative algebra.
In the result, the position space is endowed, in a natural way, with noncommutative structure which originates from
accounting of spin degrees of freedom. It is known that formalism of dynamical systems with second-class constraints
implies a natural possibility to incorporate noncommutative geometry into the framework of classical and quantum
theory [25, 26]. Our model represents an example of situation when physically interesting noncommutative particle
(6) emerges in this way. For the case, the parameter of non-commutativity is proportional to spin-tensor.
We point out that non relativistic model (1) implies canonical algebra of position operators, see [15, 18]. So the
deformation (6) arises as a relativistic correction induced by spin of the particle.
While the emergence of noncommutative structure in a classical theory is nothing more than a mathematical game,
this became crucial in quantum theory. Quantization of a theory with second-class constraints on the base of Poisson
bracket is not consistent, and we are forced to look for quantum realization of Dirac brackets. For instance, instead of
the standard quantization rule of the position, x x = x, we need to set x x = x +

with some operator

. This
leads to interesting consequences concerning the relation between classical and quantum theories, which we start to
discuss in this work.
A natural way to construct quantum observables is based on the correspondence principle between classical and
quantum descriptions. However, this straightforward approach is mostly restricted to simple models like non-
relativistic point particle. Elementary particles with spin were initially studied from the quantum perspective, because
systematically constructed classical models of spinning particle were non known. Construction of quantum observables
for an electron involves the analysis of Dirac equation and the representation theory of Lorentz group. Newton and
Wigner found possible position operator, x
NW
, by the analysis of localized states in relativistic theory [27]. Foldy and
Wouthuysen invented a convenient representation for the Dirac equation [28]. In this representation Newton-Wigner
position operator simply becomes the multiplication operator, x
NW
= x. Pryce noticed that notion of center-of-mass
in relativistic theory is not unique [29]. The Pryce center-of-mass (e) has commuting components and coincides with
the Newton-Wigner position operator, while the Pryce center-of-mass (d) is dened as a covariant object though it
has non-commutative components.
Notion of position observables in the theory of Dirac equation [3133] is in close relation with the notion of
relativistic spin. Current interest to covariant spin operators is related with a broad range of physical problems
concerning consistent denition of relativistic spin operator and Lorentz-covariant spin density matrix in quantum
information theory [3441]. Consideration of Zitterbewegung [42] and spin currents [43] in condensed matter studies
involves Heisenberg equations for position and spin observables. Precession of spin in gravitational elds gives a useful
tools to test general relativity [44]. Surprisingly, coupling of spin to gravitational elds may be important already in
the acceleration experiments due to so-called spin-rotation coupling [45]. In these applications a better understanding
of spinning particle at the classical level may be very useful.
There are a lot of operators proposed for the position and spin of relativistic electron, see [4, 2730, 39]. Which one
is a conventional position (spin) operator? Widely assumed as the best candidate is the pair of Foldy-Wouthuysen (
Newton-Wigner Pryce (e)) mean position and spin operators. Components of the mean-position operator commutes
with each other, spin obeys so(3) algebra. However, they do not represent Lorentz-covariant quantities.
To clarify these long-standing questions, in sections III, IV and V we construct relativistic quantum mechanics of
F-BMT electron. In section III, quantizing our Lagrangian in physical-time parametrization, we obtain the operators
corresponding to classical position and spin of our model. Our results argue in favor of covariant Pryce (d) position
and spin operators
2
. This implies that eects of non-commutativity could be presented at the Compton wave length,
in contrast to conventional expectations [46] of non-commutativity at Planck length.
In section IV, we construct Hamiltonian formulation in the covariant form (in an arbitrary parametrization). The
constraints p
2
+ (mc)
2
= 0 and S
2
=
3
2
4
appeared in classical model can be identied with Casimir operators of
Poincare group. That is the spin one-half representation of Poincare group represents a natural quantum realization of
our model. According to Wigner [4749], this is given by Hilbert space of solutions to two-component Klein-Gordon
(KG) equation. Two-component KG eld has been considered by Feynman and Gell-Mann [50] to describe weak
interaction of spin one-half particle in quantum eld theory, and by Brown [51] as a starting point for QED. In
contrast to KG equation for a scalar eld, the two-component KG equation admits the covariant positively dened
1
In the interacting theory namely the initial coordinates obey the F-BMT equations.
2
Pryce (e)-operators corresponds to the special variables mentioned above, see subsection III B.
4
conserved current
I

=
1
(mc)
2
( p)

( p)

, (7)
which can be used to construct a relativistic quantum mechanics of this equation. This is done in subsection VA,
then in subsection VB we show its equivalence with quantum mechanics of Dirac equation. Taking into account the
condition (5), we conclude that F-BMT electron corresponds to positive-energy sector of the KG quantum mechanics,
see subsection VC. In subsection VD, we establish the correspondence between canonical and covariant formulations
of F-BMT electron, thus proving relativistic invariance of the physical-time formalism of subsection III B. In particular,
we nd the manifestly-covariant operators
x

rp
= x

+
1
2 p
2
( p)

,

j

+
p

( p)

( p)

p
2
, (8)
and show how they can be used to compute mean values of the physical (that is Pryce (d)) operators of position and
spin. In other words, they represent manifestly-covariant form of Pryce (d)-operators.
Using the equivalence between KG and Dirac quantum mechanics, we then found the form of these operators on
space of Dirac spinors. They also can be used to compute position and spin of the Frenkel electron, see subsection
VE.
II. SEARCH FOR LAGRANGIAN
A. Variational problem with auxiliary variables
To start with, we take the Hamiltonian action [14]
S
H
=
_
d p

+
gi
g
i
H,
H =
g
1
2
(p
2
+m
2
c
2
) +
g
3
2
(
2
a
3
) +
g
4
2
(
2
a
4
) +g
5
() +g
6
(p) +g
7
(p) +
gi

gi
.
Here
gi
are conjugate momenta for the auxiliary variables g
i
. We have denoted by
gi
the Lagrangian multipliers
for the primary constraints
gi
= 0. Variation of the action with respect to
gi
gives the equations
gi
= 0, this
implies
gi
= 0. Using this in the equations
SH
gi
= 0 we obtain
3
the desired constraints (2) and (3). Our model
is manifestly Poincare-invariant. The auxiliary variables g
i
, are scalars under the Poincare transformations. The
remaining variables transform according to the rule
x

+a

, p

.
Local symmetries form two-parametric group of transformations. It is composed by the standard reparameterizations
x

= x

, p

= p

, g
i
= (g
i
),
gi
= (g
i
). (9)
as well as by spin-plane transformations with the parameter ():

= s

=
1
s

, (10)
g
3
= s

2sg
5
, g
4
=
1
s

+ 2
1
s
g
5
,
g
6
=
1
s
g
7
, g
7
= sg
6
,
3

obeys the Hamiltonian equation

= g
3

. Together with
2
> 0, this implies
2
> 0.
5
g
5
=
1
s
g
3
sg
4
,
gi
= (g
i
). (11)
We have denoted s
_
a4
a3
. Eq. (10) represents innitesimal form of the structure-group transformations of the
spin-ber bundle [18].
The coordinates x

, Frenkel spin-tensor J

and BMT vector s

BMT
J

() = 2(

), s

BMT
()
1
4
_
p
2

, (12)
are -invariant quantities. For their properties see Appendix 1. Note that the spacial components, s
i
BMT
, coincide
with spin
S
i
=
1
4

ijk
J
jk
, (13)
only in the rest frame. Both transform as a vector under spacial rotations, but have dierent transformation laws
under Lorentz boost. Where this does not lead to misunderstanding, we denote s

BMT
as s

.
Lagrangian of a given Hamiltonian theory with constraints can be restored within the known procedure [23, 24].
For the present case, it is suciently to solve Hamiltonian equations of motion for x

and

with respect to p

and

, and substitute them into the Hamiltonian action (9). Let us do this for more general Hamiltonian action,
obtaining closed formula which will be repeatedly used below.
Consider mechanics with the conguration-space variables Q
a
(), g
i
(), and with the Lagrangian action
S =
1
2
_
dG
ab
DQ
a
DQ
b
K
ab
Q
a
Q
b
M. (14)
We have denoted DQ
a


Q
a
H
a
b
Q
b
, and G(g, Q), K(g, Q), H(g, Q), and M(g) are some functions of the indicated
variables. Let us construct the Hamiltonian action functional of this theory. Denoting the conjugate momenta as P
a
,

gi
, the equations for P
a
can be solved
P
a
=
L


Q
a
= G
ab
DQ
b
,

Q
a
=

G
ab
P
b
+H
a
b
Q
b
, (15)
where

G
ab
is the inverse matrix of G
ab
. Equations for the remaining momenta turn out to be the primary constraints,

gi
= 0. Then the Hamiltonian action reads
S
H
=
_
d P
a

Q
a
+
gi
g
i
H, (16)
H =
1
2

G
ab
P
a
P
b
+P
a
H
a
b
Q
b
+
1
2
K
ab
Q
a
Q
b
+
1
2
M +
gi

gi
. (17)
Thus the Hamiltonian (16) and the Lagrangian (14) variational problems are equivalent. We point out that choosing
an appropriate set of auxiliary variables g
i
, the action (14) can be used to produce any desired quadratic constraints
of the variables Q, P.
Let us return to our problem (9). Comparing the Hamiltonian of our interest (9) with the expression (17), we dene
the doublets Q
a
= (x

), P
a
= (p

), as well as the matrices

G
ab
=
_
g
1
g
7
g
7
g
3
_
, H
a
b
=
_
0 g
6
0 g
5
_
, K
ab
=
_
0 0
0 g
4
_
,
where g
1
= g
1

and so on. Besides, we take the mass term in the form M = g


1
m
2
c
2
a
3
g
3
a
4
g
4
. With this
choice, the equation (17) turns into our Hamiltonian (9). So the corresponding Lagrangian action reads from (14) as
follows
S =
_
d
1
2 det

G
_
g
3
(Dx)
2
2g
7
(DxD) +g
1
(D)
2

1
2
g
1
m
2
c
2
+
1
2
g
3
a
3

1
2
g
4
(
2
a
4
). (18)
We have denoted
Dx

= x

g
6

, D

g
5

.
6
Using the inverse matrix
G
ab
=
1
det

G
_
g
3
g
7
g
7
g
1
_
,
the action can be written in the form
S =
_
d
1
2
G
ab
DQ
a
DQ
b
+
a
3
g
11
a
1
g
22
2 det G

1
2
g
4
(
2
a
4
),
where DQ
a
= (Dx, D).
B. Variational problem without auxiliary variables
Eliminating the auxiliary variables one by one, we get various equivalent formulations of the model (18). At the
end, we arrive at the Lagrangian action without auxiliary variables g
i
.
First, we write equations for g
5
and g
6
following from (18). They imply (D) = 0 and (Dx) = 0, then
g
5
=
( )

2
, g
6
=
( x)

2
.
We substitute the solution
4
into the action (18), this reads
S =
_
d
1
2 det

G
[g
3
( xN x) 2g
7
( xN ) +g
1
( N )]
1
2
g
1
m
2
c
2
+
1
2
g
3
a
3

1
2
g
4
(
2
a
4
). (19)
It has been denoted
N

2
, then N

= 0. (20)
Together with

N

2
, this forms a pair of projectors N +

N = 1, N
2
= N,

N
2
=

N, N

N = 0. Any vector V

can be decomposed on the transverse and longitudinal parts with respect to

, V

= V

+V

, where V

= N

,
then V

= 0; and V

=

N

=
(V )

. Further, in the action (19) we put g


7
= 0
S =
_
d
1
2g
1
( xN x)
1
2
g
1
m
2
c
2
+
1
2g
3
( N ) +
1
2
a
3
g
3

1
2
g
4
(
2
a
4
). (21)
This does not alter the dynamical equations, whereas the constraint = 0 appears as the third-stage constraint.
The rst two terms in Eq. (21) (as well as the third and the fourth terms) have the structure similar to that of
spinless particle,
1
2e
x
2

em
2
c
2
2
. It is well known, that for the case we can substitute equations of motion for e back
into the Lagrangian, this leads to an equivalent variational problem. So, we solve the equation for g
3
, g
3
=
_
N
a3
,
and substitute this back into (21), this gives
S =
_
d
1
2g
1
( xN x)
1
2
g
1
m
2
c
2
+

a
3

N
1
2
g
4
(
2
a
4
). (22)
Analogously, we solve the equation for g
1
, g
1
=

xN x
mc
and substitute this into (22), this gives the minimal action
S =
_
d mc

xN x +

a
3

N
1
2
g
4
(
2
a
4
). (23)
This depends only on transverse parts of the velocities x

and

. The second term from (23) appeared as a Lagrangian


of the particle [53, 54] inspired by Bag model [55] in hadron physics.
4
There is no guarantee that this gives an equivalent variational problem, the equivalence must be veried by direct computations.
Fortunately, for our case the trick works well.
7
C. Local symmetries of minimal action
Our model is invariant under two local symmetries. For the initial formulation (9) they have been written in Eqs.
(9) and (9). Let us see how they look for the minimal action. This is invariant under reparametrization of the lines
x

() and

() supplemented by proper transformation of the auxiliary variable g


4
(). We use the projectors N
and

N to decompose an innitesimal reparametrization as follows:
x

= x

=

N x

+N x

= N

+

N

,
g
4
= (g
4
) =
_

a
3

2
_
+
_
g
4

a
3

2
_
. (24)
Our observation is that each projection

=

N x

= N

g
4
=
_

a
3

2
_
. (25)

= N x

=

N

g
4
=
_
g
4

a
3

2
_
. (26)
separately turns out to be a symmetry of the minimal action. It can be veried using the intermediate expressions

2
= 0,

2
((N )

+ ( )),

xN x = 0,

N =
_

N
_
(
2
)

N
2
2
.

2
= (
2
),

= 0,

xN x = (

xN x),

N = (
2
)

N
2
2
.
Any pair among the transformations (24)-(26) can be taken as independent symmetries of the minimal action.
Let the functions x(), (), g
4
() represent a solution to equations of motion. Then they obey ( ) = ( x) = 0
and g
4
=

a3

2
. Using this expressions, the transformations (25) and (26) acquire the form

= 0,

g
4
= (g
4
). (27)

= x

= 0,

g
4
= 0. (28)
Hence on true trajectories the symmetries have simple meaning. -transformations (28) represent reparametrizations
of the conguration-space trajectory x

, whereas -transformations (27) represent reparametrizations of the inner-


space trajectory

. Their sum gives the standard reparametrization transformation of the theory, Eq. (24).
III. MINIMAL ACTION IN THE PHYSICAL-TIME PARAMETRIZATION
A. Positions non-commutativity due to spin
Using reparametrization invariance of the Lagrangian (23), we take physical time as the evolution parameter, = t.
Now we work with physical dynamical variables x

= (ct, x(t)) and

= (
0
(t), (t)) in the expression (23). In this
section the dot means derivative with respect to t, x

= (c,
dx
dt
) and so on. Let us construct Hamiltonian formulation
of the model (23).
Computing conjugate momenta, we obtain the primary constraint
g4
= 0, and the expressions
p
i
= mc
N x
i

xN x
, (29)
8

a
3
N

N
. (30)
Comparing expressions for p
2
and p, after tedious computations we obtain the equality which does not involve
time-derivative, p
2
+ (mc)
2
= (
p

0
)
2
. Hence Eq. (29) implies the constraint

_
p
2
+ (mc)
2

0
+p = 0.
This is analog of covariant constraint p = 0. Eq. (30) together with Eq. (20) imply more primary constraints
= 0,
2
a
3
= 0. Computing the Hamiltonian, P

QL +
a

a
, we obtain
H = c
_
p
2
+ (mc)
2
+
3
(
2
a
3
) +
1
2
g
4
(
2
a
4
) +
5
() +
6
(
_
p
2
+ (mc)
2

0
+p) +
4

g4
. (31)
Preservation in time of the primary constraints implies the following chains of algebraic consequences:

g4
= 0,
2
a
4
= 0,
5
= 0.
() = 0,
3
=
a
4
2a
3
g
4
.

_
p
2
+ (mc)
2

0
+p = 0,
_
p
2
+ (mc)
2

0
+p = 0,
6
= 0.
Three Lagrangian multipliers have been determined in the process,
5
=
6
= 0 and
3
=
a4
2a3
g
4
, whereas
1
and
4
remain an arbitrary functions. For the latter use, let us denote
p
0

_
(mc)
2
+p
2
, g

p
2
. (32)
Besides the constraints, the action implies the Hamiltonian equations
dx
i
dt
= c
p
i
p
0
,
dp
i
dt
= 0, (33)
g
4
=
4
,
g4
= 0; (34)

=
a
4
a
3
g
4

= g
4

. (35)
Equations (33) describe free-moving particle with the speed less then speed of light
x
i
= x
i
0
+v
i
t, v
i
= c
p
i
_
(mc)
2
+p
2
, p
i
= const. (36)
The spin-sector variables have ambiguous evolution, because a general solution to (35) depends on an arbitrary function
g
4
. So they do not represent the observable quantities. As candidates for the physical variables of spin-sector, we can
take either the Frenkel spin-tensor,
dJ

dt
= 0, J

= 0, J
2
= 6
2
. (37)
or, equivalently, BMT vector
ds

dt
= 0, s

= 0, s
2
=
3
2
4
. (38)
The constraints
2
a
3
= 0 and
g4
= 0 belong to rst-class, other form the second-class set. To take the latter into
account, we construct the corresponding Dirac bracket. The non vanishing Dirac brackets are
{x
i
, x
j
}
D
=

ijk
s
k
mcp
0
, {x
i
, p
j
}
D
=
ij
, {p
i
, p
j
}
D
= 0, (39)
9
{J

, J

}
D
= 2
_
g
[
J
]
g
[
J
]
_
, (40)
{x

, J

}
D
=
1
(mc)
2
(J
[
p
]

p
0
J
0[
p
]
) , (41)
{s
i
, s
j
}
D
=
p
0
mc

ijk
_
s
k

(s p)p
k
p
2
0
_
, (42)
{x
i
, s
j
}
D
=
_
s
i

(s p)p
i
p
2
0
_
p
j
(mc)
2
, (43)
where g

is the matrix dened in (32). After transition to the Dirac brackets the second-class constraints can be
used as strong equalities. In particular, we can present S
0
in terms of independent variables
s
0
=
(s p)
_
p
2
+ (mc)
2
,
and in the expression for Hamiltonian (31) only rst and second terms survive. Besides, we omit the second term, as it
does not contribute into equations for spin-plane invariant variables. In the result, we obtain the physical Hamiltonian
H
ph
= c
_
p
2
+ (mc)
2
. (44)
As it should be, the equations (33), (37) and (38) follow from physical Hamiltonian with use the Dirac bracket,

Q = {Q, H
ph
}
D
.
B. Operators of physical observables: F-BMT electron chooses Pryces (d) -type spin and position
Both operators (except p
i
) and abstract state-vectors of the physical-time formalism we denote by capital let-
ters,

Q, (t, x). In order to quantize the model, classical Dirac-bracket algebra should be realized by operators,
[

Q
1
,

Q
2
] = i {Q
1
, Q
2
}
D
|
Qi

Qi
. To start with, we look for classical variables which have canonical Dirac brackets,
thus simplifying the quantization procedure. Consider the spin variables s
j
dened by the following transformation:
s
j
=
_

jk

p
j
p
k
p
0
(p
0
+mc)
_
s
k
, then s
j
=
_

jk
+
p
j
p
k
mc(p
0
+mc)
_
s
k
.
Vector s is nothing but the spin in the rest frame. Its components have the following Dirac brackets
{ s
i
, s
j
}
D
=
ijk
s
k
, {x
i
, s
j
}
D
=
1
mc(p
0
+mc)
_
s
i
p
j

ij
(ps)
_
.
The last equation together with the following Dirac bracket: {
ikm
s
k
p
m
, s
j
} = s
i
p
j

ij
(ps), suggest to consider the
variables
x
j
= x
j

1
mc(p
0
+mc)

jkm
s
k
p
m
. (45)
The canonical variables x
j
, p
i
and

S
j
have a simple algebra
{ x
j
, x
i
}
D
= 0 , { x
i
, p
j
}
D
=
ij
, { x
j
, s
i
}
D
= 0 , { s
i
, s
j
}
D
=
ijk
s
k
. (46)
Besides, the constraints (38) on s

imply s
2
=
3
4

2
. So the corresponding operators

S
j
should realize an irreducible
representation of SO(3) with spin s = 1/2. Quantization in terms of these variables becomes straightforward. The
Hilbert space consists from two-component functions
a
(t, x), a = 1, 2. A realization of Dirac brackets algebra by
operators has the standard form
p
j
p
j
= i
j
, x
j

X
j
= x
j
, s
j
BMT

S
j
BMT
=

2

j
.
10
TABLE I: Position/spin operators for the relativistic electron [29]
=

1 0
0 1

,
i
=

0
i

i
0

,
i
=


i
0
0
i

.
Dirac representation, itD = c(
i
pi +mc)D F-W representation, it = c p
0
Classical model

X
j
P(d)
x
j
+
i
2mc


k
p
k
p
j
( p
0
)
2

x
j


2mc( p
0
+mc)

jkm
p
k
m position

S
j
P(d)
1
2m
2
c
2

m
2
c
2

j
imc
jkl

k
p
l


2mc

p
0

1
( p
0
+mc)
p
k

k
p
j

spin

X
j
P(e)
= x
j
FW
x
j
+

2 p
0

i
j
+
1
p
0
+mc

jkm
p
k
m
1
p
0
( p
0
+mc)
i
k
p
k
p
j

x
j
x
j

S
j
P(e)
=

S
j
FW

2 p
0

mc
j
im
jkl

k
p
l
+

k
p
k
p
j
p
0
+mc

j
s
j

X
j
P(c)
x
j
+

2( p
0
)
2

jkm
p
k
m +imc
j

x
j
+

2 p
0
( p
0
+mc)

jkm
p
k
m

S
j
P(c)

2( p
0
)
2

m
2
c
2

j
imc
jkl

k
p
l
+
k
p
k
p
j


2 p
0

mc
j
+
1
( p
0
+mc)
p
k

k
p
j

S
j
BMT
The conversion formulas between canonical and initial variables have no ordering ambiguities, so we immediately
obtain the operators corresponding to the physical position and spin of classical theory
x
i


X
i
= x
i


2mc( p
0
+mc)

ijk
p
j

k
, (47)

J
0i
=

mc

ijk
p
j

k
,

J
ij
=

mc

ijk
_
p
0

1
( p
0
+mc)
( p) p
k
_
. (48)

S
i
=
1
4

ijk

J
jk
=

2mc
_
p
0

1
( p
0
+mc)
( p) p
i
_
. (49)
BMT operator reads

S
0
BMT
=

2mc
( p) ,

S
j
BMT
=

2
_

j
+
1
mc( p
0
+mc)
( p) p
j
_
. (50)
The energy operator (44) determines the evolution of a state-vector by the Schrodinger equation
i
d
dt
= c
_
p
2
+ (mc)
2
, (51)
as well the evolution of operators by Heisenberg equations. The scalar product can be dened as follows
, =
_
d
3
x

. (52)
By construction, the abstract vector (t, x) of Hilbert space can be identied with amplitude of probability density
of canonical coordinate x
i
. Since our position operators x
i
are noncommutative, the issue of a wave function requires
special discussion which we postpone for the future.
To compare our operators with known in the literature, we remind that Pryce [29] wrote his operators acting on
space of Dirac spinor
D
, see the rst column in Table I. Foldy and Wouthuysen [28] found unitary transformation
which maps the Dirac equation i
t

D
= c(
i
p
i
+ mc)
D
into the pair of square-root equations i
t
= c p
0
.
After the FW transformation, the Pryce operators acquire block-diagonal form on space , see the second column.
Our operators act on space of solutions of square-root equation (51), so we compare them with positive-energy parts
(upper-left blocks) of Pryce operators of the second column.
Our operators of canonical variables

X
j
= x
j
and

S
j
correspond to the Pryce (e) ( Foldy-Wouthuysen Newton-
Wigner) position and spin operators.
However, operators of position x
j
and spin S
j
of our model are

X
j
and

S
j
. They correspond to the Pryce (d)-
operators.
Operator of BMT-vector

S
j
BMT
is the Pryce (c) spin.
While we have started from relativistic theory (23), working with the physical variables we have loosed, from
the beginning, the manifest relativistic covariance. Whether the quantum mechanics thus obtained is a relativistic
theory? Below we present a manifestly covariant formalism and conrm that scalar products, mean values and
transition probabilities can be computed in a covariant form.
11
IV. MINIMAL ACTION IN COVARIANT FORMALISM. COVARIANT FORM OF
NONCOMMUTATIVE ALGEBRA OF POSITIONS
Obtaining the minimal action (4) we have made various tricks. So, let us conrm that the action indeed leads to
the desired constraints (2) and (3). Computing conjugate momenta we obtain the primary constraint
g4
= 0, and
the expressions
p

= mc
N x

xN x
,

a
3
N

N
.
Due to Eq. (20), they imply more primary constraints, p = 0, p
2
+(mc)
2
= 0, = 0, and
2
a
3
= 0. Computing
the Hamiltonian, P

QL +
a

a
, we obtain
H =
1
2

1
(p
2
+m
2
c
2
) +
3
(
2
a
3
) +
1
2
g
4
(
2
a
4
) +
5
() +
6
(p) +
4

g4
. (53)
Preservation in time of the primary constraints implies the following chains of algebraic consequences:

g4
= 0,
2
a
4
= 0,
5
= 0.
() = 0,
3
=
a
4
2a
3
g
4
.
(p) = 0, (p) = 0,
6
= 0.
As the result, the minimal action generates all the desired constraints (2) and (3). Three Lagrangian multipliers have
been determined in the process,
5
=
6
= 0 and
3
=
a4
2a3
g
4
, whereas
1
and
4
remain an arbitrary functions.
Besides the constraints, the action implies the Hamiltonian equations g
4
=
4
,
g4
= 0, x

=
1
p

, p

= 0,

=
a4
a3
g
4

= g
4

. General solution to these equations in an arbitrary and proper-time parameterizations is


presented in Appendix 2.
To take into account the second-class constraints T
4
, T
5
, T
6
and T
7
, we pass from Poisson to Dirac bracket. We write
them for the spin-plane invariant variables, they are x

, p

and either the Frenkel spin-tensor or BMT four-vector


(12). The non vanishing Dirac brackets are as follows.
Spacial sector:
{x

, x

} =
1
2p
2
J

, {x

, p

} =

, {p

, p

} = 0. (54)
Frenkel sector:
{J

, J

} = 2(g

+g

) , (55)
{x

, J

} =
1
p
2
J
[
p
]
, (56)
BMT-sector:
{s

, s

} =
1
_
p
2

=
1
2
J

, (57)
{x

, s

} =
s

p
2
=
1
4
_
p
2

p
2
. (58)
In the equation (55) it has been denoted g

p
p
2
. Together with g

p
p
2
, this forms a pair of projectors
g + g = 1, g
2
= g, g
2
= g, g g = 0. The transition to spin-plane invariant variables does not spoil manifest covariance.
So, we write equations of motion in terms of these variables
x

=
1
p

, p

> 0, (59)

= 0, J

= 0, J
2
= 6
2
. (60)
12

= 0, S

= 0, S
2
=
3
2
4
. (61)
Besides, we have the rst-class constraint
p
2
+ (mc)
2
= 0 , where p
0
> 0. (62)
Let us compare these results with non manifestly covariant formalism of previous section. Evolution of physical
variables can be obtained from equations (59)-(62) assuming that the functions Q

() represent the physical variables


Q
i
(t) in the parametric form. Using the formula
dF
dt
= c

F()
x
0
()
, this gives Eqs. (33), (37) and (38). The brackets
(39)-(43) of physical variables appeared, if we impose the physical-time gauge x
0
= 0 for the constraint (62), and
pass from (54)-(58) to the Dirac bracket which take into account this second-class pair. Physical Hamiltonian (44)
can be obtained from (53) considering the physical-time gauge as a canonical transformation [23].
Summarizing, in classical mechanics all basic relations for physical variables can be obtained from covariant for-
malism. In the next section we discuss, how far we can proceed towards formulation of quantum mechanics in a
manifestly-covariant form.
V. MANIFESTLY-COVARIANT FORM OF QUANTUM MECHANICS OF THE FRENKEL ELECTRON
According to Wigner [4749], with an elementary particle in QFT we associate the Hilbert space of representation
of Poincare group. The space can be described in a manifestly covariant form as a space of solutions to Klein-
Gordon (KG) equation for properly chosen multicomponent eld
i
(x

). One-component eld corresponds to spin-


zero particle. Two-component eld has been considered by Feynman and Gell-Mann [50] to describe weak interaction
of spin one-half particle, and by Brown as a starting point for QED [51]. It is well-known, that one-component KG eld
has no quantum-mechanical interpretation. In contrast, two-component KG equation does admit the probabilistic
interpretation: the four-vector (65) represents positively dened conserved current of this equation. On this base,
we consider below the relativistic quantum mechanics of two-component KG equation and show its equivalence with
quantum mechanics of Dirac equation. Then we show that the covariantly quantized F-BMT electron corresponds to
positive-energy sector of this quantum mechanics. At last, we establish the correspondence between canonical and
covariant formulations, thus proving relativistic invariance of the physical-time formalism of subsection III B.
A. Relativistic quantum mechanics of two-component Klein-Gordon equation
We denote states and operators of covariant formalism by small letters, to distinguish them from the quantities of
canonical formalism. Consider the space of abstract state-vectors composed by two-component Weyl spinors
a
(x

),
a = 1, 2. Generators of Poincare transformations in this space read
m

= x

+
1
2

, p

= i

, (63)
where the Lorentz generators

=
i
2
(

),
are built from standard Pauli matrices
i
combined into the sets

= (1,
i
),

= (1,
i
).
They are hermitian and obey

= 2

= 2

. Further, on the Poincare-invariant subspace


selected by two-component KG equation
( p
2
+m
2
c
2
) = 0 , (64)
we dene an invariant and positive-dened scalar product as follows. The four-vector
5
I

[, ] =
1
m
2
c
2
( p)

, (65)
5
denotes usual Hermitian conjugation, a

= ( a

)
T
, ( a

b)

=

b

, then ( p

f)

= p

(f)

.
13
represents a conserved current of Eq. (64), that is

= 0, when and satisfy to Eq. (64). Then the integral


(, ) =
_

, d

=
d
4
x
dx

, (66)
does not depend on the choice of a space-like 3-dimensional hyperplane (an inertial coordinate system). As a
consequence, this does not depend on time. So we can restrict ourselves to the hyperplane dened by the equation
x
0
= const, then
(, ) =
_
d
3
xI
0
. (67)
Besides, this scalar product is positive-dened, since
I
0
[, ] =
1
m
2
c
2
( p)

p +

> 0. (68)
So, this can be considered as a probability density of operator x = x. We point out that transformation properties
of the column are in the agreement with this scalar product: if transforms as a (right) Weyl spinor, then I

represents a four-vector.
Now we can conrm relativistic invariance of scalar product (52) of canonical formalism. The operator p
0
is
hermitian on the subspace of positive-energy solutions , so we can write
(, ) =
_
d
3
x
1
m
2
c
2
( p)

p +

=
_
d
3
x
__
1
mc
p +i
_

_
1
mc
p +i
_
,
This suggests the map W : {} {}
= W , W =
1
mc
p +i , W
1
=
1
2 p
0
(i p mc) , (69)
which respects the scalar products (52) and (67), and thus proves relativistic invariance of the scalar product ,
, = (, ) . (70)
We note that map W is determined up to an isometry, we can multiply W from the left by an arbitrary unitary
operator U, W W

= UW, U

U = 1. Here denotes Hermitian conjugation with respect to scalar product , .


The ambiguity in the denition of W can be removed by the polar decomposition of the operator [56]. A bounded
operator between Hilbert spaces admits the following factorization: W = PV , where V = (W

W)
1/2
, P = WV
1
.
Positively dened operator W

W > 0 has a unique square root (W

W)
1/2
. Moreover W

W = W

, therefore V
denes map from {} to {} without ambiguity. We present the explicit form of V in subsection VD.
B. Relation with Dirac equation
Here we demonstrate equivalence of quantum mechanics of KG and Dirac equations. To this aim, let us replace
two equations of second order, (64), by equivalent system of four equations of the rst order. To achieve this, with
the aid of the identity p

, we represent (64) in the form

+m
2
c
2
= 0. (71)
Consider an auxiliary two-component function

(Weyl spinor of opposite chirality), and dene evolution of and

according the equations


6

) +m
2
c
2
= 0, (72)
(

) mc

= 0. (73)
6
Note that

can be considered as conjugated momentum for , than the passage from (71) to (74) is just the passage from a Lagrangian
to Hamiltonian formulation. Similar interpretation can be developed for the Schrodinger equation, see [52].
14
That is dynamics of is determined by (71), while

accompanies :

is determined from the known taking its
derivative,

=
1
mc
( p). Evidently, the systems (64) and (72), (73) are equivalent. Rewriting the system (72), (73)
in a more symmetric form, we recognize the Dirac equation
_
0

0
__

_
+mc
_

_
= 0, or (

W
p

+mc) = 0 , (74)
for the Dirac spinor =
_
,

_
in the Weyl representation of -matrices

0
W
=
_
0 1
1 0
_
,
i
W
=
_
0
i

i
0
_
.
This gives one-to-one correspondence among two spaces. With each solution to KG equation we associate the
solution
[] =
_

1
mc
( p)
_
,
to the Dirac equation. Below we also use the Dirac representation of -matrices

0
=
_
1 0
0 1
_
,
i
=
_
0
i

i
0
_
. (75)
In this representation, the Dirac spinor corresponding to reads

D
[] =
1

2
_
1 1
1 1
__

1
mc
( p)
_
=
1

2mc
_
[( p) +mc]
[( p) mc]
_
. (76)
The conserved current (65) of KG equation (64), being rewritten in terms of Dirac spinor, coincides with the Dirac
current
I

[
1
,
2
] =

[
1
]

[
2
]. (77)
Therefore, the scalar product (66) coincides with that of Dirac.
C. Covariant operators of F-BMT electron
In a covariant scheme, we need to construct operators x

, p

, s

BMT
whose commutators
[ q
1
, q
2
] = i {q
1
, q
2
}
D
|
qi qi
, (78)
are dened by the Dirac brackets (54)-(58). Inspection of the classical equations S
2
=
3
2
4
and p
2
+(mc)
2
= 0 suggests
that we can look for a realization of operators in the Hilbert space constructed in subsection VA.
With the spin-sector variables we associate the operators
s

BMT
s

=
1
4
_
p
2

, (79)
J


2
_
p
2

+
p

( p)

( p)

p
2
. (80)
They obey the desired commutators (78), (57), (55). To nd the position operator, we separate the inner angular
momentum

j

in the expression (63) of Poincare generator


m

=
_
x

+
( p)

2 p
2
_
p

_
x

+
( p)

2p
2
_
p

+
1
2

. (81)
15
This suggests the operator of relativistic position
x

rp
= x

+
1
2 p
2
( p)

, (82)
where x

= x

. The operators p

= i

, (79), (80) and (82) obey the algebra (78), (54)-(58).


Equation (61) in this realization states that square of second Casimir of Poincare group has xed value
3
2
4
, and
in the representation chosen is satised identically. The equations (62) just state that we work in the positive-energy
subspace of the Hilbert space of KG equation (64).
We thus completed our covariant quantization procedure by matching classical variables of reparametrization-
invariant formulation to operators acting on the Hilbert space of two component spinors with scalar product (66).
The construction presented is manifestly Poincare-covariant. In the next subsection we discuss the connection between
canonical and manifestly covariant formulations of the F-BMT electron.
D. Relativistic invariance of canonical formalism
Relativistic invariance of the scalar product (52) has been already shown in subsection VA. Here we show how the
covariant formalism can be used to compute mean values and probability rates of canonical formulation, thus proving
its relativistic covariance. Namely, we conrm the following
Proposition. Let
H
+
can
= { (t, x) ; i
d
dt
=
_
p
2
+ (mc)
2
, , =
_
d
3
x

} ,
is Hilbert space of canonical formulation and
H
cov
= { (x

) ; ( p
2
+m
2
c
2
) = 0 , (, ) =
_

, I

=
1
(mc)
2
( p)

} ,
is Hilbert space of two-component KG equation. With state-vector we associate as follows:
= V
1
, V
1
=
1
2
_
p
0
(p
0
+mc)
[mc p] . (83)
Then , = (, ). Besides, mean values of the physical position and spin operators (47)-(49) can be computed as
follows
,

X
i
= Re(, x
i
rp
), ,

J
ij
= (,

j
ij
) , ,

S
i
=
ijk
(,

j
jk
) ,
where x
i
rp
and

j
ij
are spacial components of the manifestly-covariant operators
x

rp
= x

+
1
2 p
2
( p)

,

j

+
p

( p)

( p)

p
2
.
We also show that the map V can be identied with Foldy-Wouthuysen transformation applied to the Dirac spinor
(76).
It will be convenient to work in the momentum representation, (x

) =
_
d
4
p(p

)e
i

px
. Transition to the mo-
mentum representation implies the substitution
p

, x

i

p

,
in the expressions of covariant operators (79), (80), (82) and so on.
An arbitrary solution to the KG equation reads
(t, x) =
_
d
3
p
_
(p)e
i

px
0
+

(p)e

px
0
_
e

(px)
,
p

_
p
2
+ (mc)
2
,
16
where (p) and

(p) are arbitrary functions of three-momentum, they correspond to positive and negative energy
solutions. The scalar product can be written then as follows
(, ) = 2
_
d
3
p
p
m
2
c
2
_

( p)

(p)

_
, where ( p) =
p
+ (p) , (p) =
p
+ (p) .
We see that this scalar product separates positive and negative energy parts of state vectors. Since our classical
theory contains only positive energies, we restrict our further considerations by the positive energy solutions only. In
the result, in the momentum representation the scalar product (67) reads in terms of non-trivial metric as follows:
(, ) =
_
d
3
p

, =
2
p
m
2
c
2
( p) . (84)
Now our basic space is composed by arbitrary functions (p). The operators x
i
, s

and

j

act on this space as


before, with the only modication, that p
0
(p) =
p
(p). The operator x
0
and, as a consequence, the operator x
0
rp
,
do not act in this space. Fortunately, they are not necessary to prove the proposition formulated above.
Given operator

A we denote its hermitian conjugated in space H
+
can
as

A

. Hermitian operators in space H


+
can
have both real eigenvalues and expectation values. Consider an operator a in space H
cov
with real expectation values
(, a) = (, a)

. It should obey a

= a. That is, such an operator in H


cov
should be pseudo-Hermitian. We
denote pseudo-Hermitian conjugation in H
cov
as follows a
c
=
1
a

. Then pseudo-Hermitian part of an operator a


is given by
1
2
( a + a
c
).
Let us check the pseudo-Hermicity properties of basic operators. From the following identities:
(

=
_

+
2i
p
2
(p)(p

)
_
, (

= (

+ 2i[p

(p)

]) ,
( x
j
rp
)

=
_
x
j
rp
+
i
m
2
c
2

p
_
m
2
c
2

p
p
j
p
j
(p)
__
,
we see that operators

and x
j
rp
are non-pseudo-Hermitian, while operators p

, s

,

j

and orbital part of m


ij
are
pseudo-Hermitian.
To construct the map (83) we look for square root of the metric, V =
1/2
. Metric is positively dened, therefore
the square root is unique [56], this reads
V =
1
mc
_

p

p
+mc
[( p) +mc] . (85)
We use this to dene the map H
cov
H
+
can
, = V , which corresponds to the polar decomposition of map W
dened in (69). Then the scalar product (84) can be rewritten as
(, ) =
_
d
3
p(V )

V =
_
d
3
p

= , .
This proves relativistic invariance of the scalar product , of canonical formalism.
Our map dened by operator V turns out to be in the close relation with the Foldy-Wouthuysen transformation.
It can be seen applying the Foldy-Wouthuysen unitary transformation
U
FW
=

p
+mc + ( p)
_
2(
p
+mc)
p
to the Dirac spinor
D
[],

FW
[] = U
FW

D
[] =
_
V
0
_
=
_

0
_
.
The last equation means that operator V is a restriction of operator U
FW
to the space of positive-energy right Weyl
spinors .
The transformation between state-vectors induces the map of operators

Q = V qV
1
, (86)
17
TABLE II: Operators of canonical and manifestly covariant formulations in momentum representation
Canonical formalism (p) Covariant formalism (p)

Pj pj pj pj

S
i
s
i
2mc

p
i

1
(p+mc)
( p)p
i

p
2(mc)
2

p
i
( p)p
i
iimnp
m

X
i
x
i
V
i

p
i


2mc(p+mc)

ijk
pj
k
i

p
i
+
ip
i
( p)
2p
2
p
+
ip
i
2p

i
2p
2
p
i
+

2p
2

ijk
pj
k

J
ij

j
ij
mc

ijk

p
k

1
(p+mc)
( p)p
k

p
m
2
c
2

ijk
(p
k
( p)p
k
i
kmn
p
m

n
)

J
0i

j
0i


mc

ijk
pj
k


m
2
c
2

ijk

p
k
i
kml
p
m

pj

S
0
BMT
s
0
BMT

2mc
( p)

2mc
( p)

S
i
BMT
s
i
BMT

i
+
1
mc(p+mc)
( p)p
i


2mc
(p
i
+i
ijk
pj
k
)
where
V
1
=
1
2
_

p
(
p
+mc)
[mc (p)] .
Then
,

Q = (, q) . (87)
Due to Hermicity of V , V

= V , pseudo-Hermitian operators, q

V
2
= V
2
q, transform into Hermitian operators

=

Q. For an operator q which commutes with momentum operator, transformation (86) acquire the following
form

Q =
1
2
( q + q

)
1
2(
p
+mc)
( q q

)( p) .
Using this formula, we have checked by direct computations that covariant operators p,

j

and s

BMT
transform
into canonical operators p,

J

and

S

BMT
, so the spacial part of

J

,

S
i
=
ijk

J
jk
represents the classical spin
S
i
. This observation together with Eq. (87) implies that mean values of the operators of canonical formalism are
relativistic-covariant quantities.
Concerning the position operator, we rst apply the inverse to Eq. (86) to our canonical coordinate

X
i
= i

p
i
in
the momentum representation

x
i
V
= V
1

X
i
V =

X
i
+ [V
1
,

X
i
]V = i

p
i

ip
i
(p)
2mc
p
(
p
+mc)
+
ip
i
2
p
+
i
2mc

i
+

2mc(
p
+mc)

ijk

j
p
k
.
Our position operator then can be mapped as follows:
x
i
V
= V
1
_
i

p
i
+
1
mc(
p
+mc)

ijk

S
j
P
k
_
V = i

p
i
+
ip
i
( p)
2p
2

p
+
ip
i
2
p

i
2p
2

i
+

2p
2

ijk
p
j

k
.
We note that pseudo-Hermitian part of operator x
i
rp
coincides with the image x
i
V
,
x
i
V
=
1
2
_
x
i
rp
+
_
x
i
rp

c
_
.
Since x

rp
has explicitly covariant form, this also proves covariant character of position operator

X
i
. Indeed, (86)
means that matrix elements of

X
i
are expressed through the real part of manifestly covariant matrix elements
,

X
i
= (, x
i
V
) = Re(, x
i
rp
) .
In summary, we have proved the proposition formulated above. The operators

j

and x

rp
, which act on the space of
two-component KG equation, represent manifestly-covariant form of the Pryce (d)-operators.
Table II summarizes manifest form of operators of canonical formalism and their images in covariant formalism.
18
E. Manifestly-covariant operators of spin and position of Dirac equation
According to Eq. (77), the scalar product (, ) coincides with that of Dirac. This allows us to nd manifestly-
covariant operators in the Dirac theory which have the same expectation values as

j

and x

rp
. Consider the following
analog of

j

on the space of 4-component Dirac spinors

D
=

+
p

p
2
=

+
i
p
2
( p

) ( p) , (88)
where

=
i
2
(

). This denition is independent from a particular representation of -matrices. In the


representation (75) this reads

=
_

0
0 (

_
,
and can be used to prove the equality of matrix elements
_
d
3
x[]

D
[] = (,

) ,
for arbitrary solutions , of two-component KG equation. The covariant position operator can be dened as follows:
x

D
= x

2 p
2
+
i(
5
1) p

2 p
2
= x

+
i

2 p
2
( p) +
i
5
p

2 p
2
, (89)
where
5
= i
0

3
. Again, one can check that matrix elements in two theories coincide
_
d
3
x
[
]

[
] = (, x

rp
) .
As a result, the manifestly-covariant operators

j

D
and x

D
of the Dirac equation represent position x and spin S (13)
of the Frenlel electron (23). Their mean values can be computed as follows
,

X
i
=
1
2
Re([], [ x
i
D
+ x
i
D
][])
D
, ,

S
i
=
ijk
([],

j
jk
D
[])
D
. (90)
VI. CONCLUSIONS
The content and the main results of this work have been described in Introduction. So, here we nish with some
complementary comments.
There are a lot of candidates for spin and position operators of the relativistic electron. Dierent position observables
coincide when we consider standard quasi-classical limit. So, in absence of a systematically constructed classical model
of an electron it is dicult to understand the dierence between these operators. Our approach allows us to do this,
after realizing them at the classical level. As we have seen, various non-covariant, covariant and manifestly-covariant
operators acquire clear meaning in the Lagrangian model of Frenkel electron developed in this work.
Starting with variational formulation we described the relativistic Frenkel electron with the aid of singular La-
grangian. Equations of motion for the classical model are consistent [19, 20] with experimentally tested BMT
equations. We showed that the classical position variables are non-commutative quantities. Selecting physical-time
parametrization in our model in the case of free electron, we have done canonical quantization procedure. As it
should be, we arrived at quantum mechanics which can be identied with positive-energy part of Dirac theory in the
Foldy-Wouthuysen representation. The Foldy-Wouthuysen mean-position and spin operators correspond to canonical
variables x
j
and s
j
of the model, whereas the classical position x and spin S are represented by Pryce (d)-operators.
Since all variables obey the same equations in the free theory, the question of which of them are the true position and
spin is a matter of convention. The situation changes in interacting theory, where namely x and S obey the expected
F-BMT equations and thus represent the position and spin.
Concerning the position, in his pioneer work [29], Pryce noticed that except the particles of spin 0, it does not
seem to be possible to nd a denition which is relativistically covariant and at the same time yields commuting
coordinates. Now we know, why this happens. At the classical level, an accurate account of spin (that is of Frenkel
19
condition) in a Lagrangian theory yields, inevitably, to relativistic corrections to the classical brackets of position
variables.
It seems to be very interesting to study

X
j
P(d)
as the true relativistic position operator in more details. The rst
reason is an interesting modication of quantum interaction between the electron and background electromagnetic
elds coming from non-local interactions p

+
e
c
A

(

X
j
), F

(

X
j
)

J

. The second reason is due to its natural


non-commutativity. Instead of introducing non-commutativity to theoretical models by hands, one can study usual
relativistic electron as a physical model of non-commutative particle. We will return to these issues in the next paper
[20].
We also quantized our model in an arbitrary parametrization, keeping the manifest Lorentz-invariance. The co-
variant quantization gives positive-energy sector of two-component Klein-Gordon equation (quantum eld theory of
two-component KG proposed by Feynman and Gell-Mann [50]). We have found a covariant conserved current for the
two-component KG equation, which allows us to dene an invariant, positive-denite scalar product with metric in
the space of two-component spinors. The resulting relativistic quantum mechanics represents one-particle sector of
the Feynman-Gell-Mann quantum eld theory. Classical spin-plane invariant variables p

, S

and J

produce well
dened quantum operators of momentum and spin, p

, s

and

j

.
The square root of metric, V =
1/2
, denes the map from canonical to covariant formulations. This allows us
to establish relativistic covariance of canonical formalism: scalar product and mean values of operators of canonical
formalism can be computed using the corresponding quantities of covariant formalism, see the proposition of subsection
VD. And back, the transformation V allows us to interpret the results of covariant quantization in terms of one-
particle observables of an electron in the FW representation (see Table II). The relativistic-position operator x

rp
is
non-Hermitian and does not correspond to a physical observable. However, Hermitian part of x
j
rp
coincides with
image of physical-position operator x
j
V
= V
1

X
i
V .
Our classical model may provide a unication in modern issues of quantum observables in various theoretical and
experimental setups [31-46]. Since the model constructed admits an interaction with electro-magnetic and gravitational
elds, one can try to extend the obtained results beyond the free relativistic electron.
VII. ACKNOWLEDGMENTS
This work has been supported by the Brazilian foundation CNPq. AMPM thanks CAPES for the nancial support
( Programm PNPD/2011).
Appendix 1. Some identities

= (, +, +, +),
0123
= 1,
0123
= 1,

abcd

ab
= 2(
c

).

abc

ijk
= [
a
i
(
b
j

c
k

b
k

c
j
)
a
j
(
b
i

c
k

b
k

c
i
) +
a
k
(
b
i

c
j

b
j

c
i
)].
Given quantities J

= J

and p

, we dene the vectors


s

=
1
4
_
p
2

, then s

= 0, (91)

= J

, then

= 0.
Then both J

and its dual,



J

=
1
2

ab
J
ab
, can be decomposed on these vectors
J

p
2

2
_
p
2

ab
p
a
s
b
,

ab
J
ab
= 4
p

_
p
2

2
p
2

ab
p
a

b
.
20
we have the identity
s

=
1
4p
2
(J

)
2
+
1
8
J

.
If J

obeys
J

= 0,
then J

and S

turn out to be equivalent


s

=
1
4
_
p
2

, J

=
2
_
p
2

,
and obey the identities

ab
J
ab
= 4
p

_
p
2
,
s

=
1
8
J

.
In the rest system of p

, p

= (p
0
,

0),
_
p
2
= |p
0
| = mc we have
s
0
= 0, s
i
=
p
0
4|p
0
|

ijk
J
jk
.
The last equality explains our normalization for the BMT vector s

, Eq. (91).
Appendix 2. General solution to equations of motion
Lagrangian equations. Variation of the minimal action (23) implies the equations
S
g
4
= 0 :
2
= a
4
, ( ) = 0, (92)
S
x
= 0 : mc
_
N x

xN x
_
= 0, mc
N x

xN x
= p

= const,
(p) = 0, p
2
= (mc)
2
, (p ) = 0, (93)
S

= 0 :

a
3
_
N

N
_
+

a
3
( )

N
N

+
( x)

2
mcN x

xN x
+g
4

= 0. (94)
Using the consequences pointed in Eqs. (92) and (93), we simplify the equation (94)

a
3
_


2
_

( x)
a
4
p

+g
4

= 0.
Contraction of this equation with

gives the expression for g


4
g
4
=

a
3


2
a
4
,
whereas contraction with p

implies ( x) = 0. Collecting all this, the initial Lagrangian equations can be presented
in the equivalent form
_
x

x
2
_
= 0, (95)
21
_


2
_
+


2
a
4

= 0, (96)

2
= a
4
, ( x) = 0. (97)
We have second-order equations (95) and (96). Besides, there are presented two Lagrangian constraints (97).
General solution to equations (95)-(97) reads
x

= x

0
+p

1
(), (98)

=
_
a
4
a
3
A

sin f() +
_
a
4
a
3
B

cos f(), (99)


where
1
and f are arbitrary functions of the evolution parameter. Constants of integration obey the restrictions
p
2
= (mc)
2
, (pA) = (pB) = 0, A
2
= a
3
, B
2
= a
3
, (AB) = 0. (100)
Eq. (98) determines straight line (as geometric place of points) in Minkowski space, whereas (99) is an ellipse which
lies on the plane of inner space formed by the vectors A

and B

. Due to the arbitrary functions () and f(),


evolution along the trajectories is not specied, as it should be in a reparametrization invariant theory.
General solution to Hamiltonian equations. Hamiltonian formulation leads to the same result. Hamiltonian
constraints and equations written in section IV do not determine the multipliers
1
and
4
. As a consequence, the
variable g
4
() can not be determined neither with the constraints nor with the dynamical equations. This implies
the functional ambiguity in solutions to the equations of motion for the basic variables x

and

: besides the
integration constants, solution depends on these arbitrary functions.
Denoting
f() =
_
a
4
a
3
_
dg
4
,
general solution to the Hamiltonian equations is given by Eqs. (98)-(100) and

= A

sin f() B

cos f().
Covariant dynamics in proper-time parametrization. The physical variables obey to non degenerate equa-
tions, but they are not manifestly covariant. The standard way to work with nondegenerated equations keeping
covariance is to x parametrization to be proper time of the particle, x

= x

(s). Here s is the time measured in


Lorentz frame moving with the particle (the rest frame). As the proper time coincides with interval between the
particle positions, in the proper-time parametrization we have the relation
7
( x

(s))
2
= c
2
.
This equation together with Eq. (59) xes
1
=
1
m
, so we arrive at the deterministic equations
x

(s) =
p

m
, p

(s) = 0; p
2
= (mc)
2
,
with the solution being
x

= x

0
+
p

m
s, p
2
= (mc)
2
, p

= const.
As before, physical dynamical variables x
i
(t) obtained from x

(s) excluding the parameter s.


Spin-sector is described either by Eq. (37) or by Eq. (38).
[1] J. Frenkel, Die elektrodynamik des rotierenden elektrons, Zeitschrift f ur Physik, 37(4-5) (1926) 243.
7
In an arbitrary parametrization we have ( x

())
2
= c
2
s
2
(), that is nothing interesting.
22
[2] J. Frenkel, Spinning Electrons, Nature 117 (1926) 653.
[3] V. Bargmann, L. Michel, and V. L. Telegdi, Phys. Rev. Lett. 2, 435 (1959).
[4] H. C. Corben. Classical and quantum theories of spinning particles, Holden-Day, San Francisco, 1968.
[5] A. J. Hanson and T. Regge, The relativistic spherical top, Annals of Physics, 87(2) (1974) 498.
[6] F. A. Berezin and M. S. Marinov, Particle spin dynamics as the Grassmann variant of classical mechanics, Ann. Phys.
104 (1977) 336.
[7] S. P. Gavrilov and D. M. Gitman, Int. J. Mod. Phys. A15 (2000) 4499.
[8] A. O. Barut and W. Thacker, Phys. Rev. D 31 (1985) 1386.
[9] A. O. Barut and A. J. Bracken, Zitterbewegung and the internal geometry of the electron, Phys. Rev. D 23 (1981) 2454.
[10] P. Grassberger, Classical charged particles with spin, Journal of Physics A: Mathematical and General, 11(7) (1978) 1221.
[11] G. Cognola, L. Vanzo, S. Zerbini, and R. Soldati, On the lagrangian formulation of a charged spinning particle in an
external electromagnetic eld, Physics Letters bf B 104(1) (1981) 67.
[12] A. A. Deriglazov, Spinning-particle model for the Dirac equation and the relativistic Zitterbewegung, Phys. Lett. A 376
(2012) 309.
[13] A. A. Deriglazov, Classical-mechanical models without observable trajectories and the Dirac electron, Phys. Lett. A 377
(2012) 13.
[14] A. A. Deriglazov, Variational problem for the Frenkel and the Bargmann-Michel-Telegdi (BMT) equations, Mod. Phys.
Lett. A 28 (2013) 1250234; arXiv:1204.2494.
[15] A. A. Deriglazov, Nonrelativistic spin: `a la Berezin-Marinov quantization on a sphere, Modern Physics Letters A 25(32)
( 2010) 2769.
[16] A. A. Deriglazov, Semiclassical description of relativistic spin without use of Grassmann variables and the Dirac equation,
Ann. Phys. 327 (2012) 398.
[17] A. A. Deriglazov, B. F. Rizzuti, G. P. Z. Chauca, P. S. Castro, Non-Grassmann mechanical model of the Dirac equation,
J. Math. Phys. 53 (2012) 122303; arXiv:1202.5757.
[18] A. A. Deriglazov and A. M. Pupasov-Maksimov, Geometric constructions underlying relativistic description of spin on the
base of non-Grassmann vector-like variable, arXiv:1311.7005
[19] W. G. Ramirez, A. A. Deriglazov and A. M. Pupasov-Maksimov, Frenkel electron and a spinning body in a curved back-
ground, arXiv:1311.5743.
[20] A. A. Deriglazov and A. M. Pupasov-Maksimov, Lagrangian of F-BMT spinning electron on electromagnetic background,
in preparation.
[21] G.Fulop, D. M. Gitman and I. V. Tyutin, Reparametrization Invariance as Gauge Symmetry, Int. J. Theor. Phys. 38
(1999) 1941.
[22] P. A. M. Dirac, Lectures on quantum mechanics, Yeshiva University, New York 1964.
[23] D. M. Gitman and I. V. Tyutin, Quantization of elds with constraints, Springer-Verlag, Berlin, 1990.
[24] A. Deriglazov, Classical Mechanics: Hamiltonian and Lagrangian Formalism, Springer-Verlag, 2010.
[25] A. A. Deriglazov, Poincare covariant mechanics on noncommutative space, JHEP 0303 (2003) 021; arXiv: hep-th/0211105.
[26] A. A. Deriglazov, Noncommutative relativistic particle on the electromagnetic background, Phys. Lett. B 555 (2003) 83;
arXiv: hep-th/0211105.
[27] T. D. Newton and E. P. Wigner, Localized states for elementary systems, Rev. Mod. Phys. 21 (1949) 400.
[28] L. L. Foldy and S. A. Wouthuysen, On the Dirac theory of spin 1/2 particles and its non-relativistic limit, Phys. Rev. 78
(1950) 29.
[29] M. H. L. Pryce, The mass-centre in the restricted theory of relativity and its connexion with the quantum theory of
elementary particles, Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 195
(1948) 62.
[30] R. P. Feynman, Quantum Electrodynamics, W A Benjamin, 1961.
[31] G. N. Fleming, Covariant position operators, spin, and locality, Physical Review 137(1B) (1965) B188.
[32] M. Bunge and A. J. Kalnay, A covariant position operator for the relativistic electron, Progress of Theoretical Physics
42(6) ( 1969) 1445.
[33] A. J. Kalnay and E. Mac Cotrina, On proper time and localization for the quantum relativistic electron, Progress of
Theoretical Physics 42(6) (1969) 1422.
[34] C. Chicone, B. Mashhoon, and B. Punsly, Relativistic motion of spinning particles in a gravitational eld, Physics Letters
A 343(1) (2005) 1.
[35] D. Singh and N. Mobed, The implications of noninertial motion on covariant quantum spin, Classical and Quantum
Gravity 24(10) (2006) 2453.
[36] L. M. Slad, Spin rotation as an element of polarization experiments on elastic electron-proton scattering, Physics Letters
A 374(10) (2010) 1209.
[37] W. T. Kim and E. J. Son, Phys. Rev. A 71 (2005) 014102.
[38] T. F. Jordan, A. Shaji and E. C. G. Sudarshan, Phys. Rev. A 73 (2006) 032104.
[39] M. Czachor, Phys. Rev. A 55 (1997) 72.
[40] A. G. S. Landulfo and G. E. A. Matsas, Phys. Rev. A 80 (2009) 044302.
[41] P. Caban and J. Rembieli nski, Lorentz-covariant reduced spin density matrix and einstein-podolsky-rosen-bohm correlations,
Phys. Rev. A 72 (2005) 012103.
[42] J. Cserti and G. David, Unied description of zitterbewegung for spintronic, graphene, and superconducting systems,
Physical Review B 74(17) (2006) 172305.
23
[43] Semiconductor Spintronics and Quantum Computation, edited by D. Awschalom, D. Loss, and N. Samarth (Springer,
Berlin, 2002).
[44] Yu. N. Obukhov, A. J. Silenko and O. V. Teryaev, Phys. Rev. D 80 (2009) 064044.
[45] G. Lambiase and G. Papini. Spin-rotation coupling in compound spin objects, Physics Letters A 377(14) ( 2013) 1021.
[46] A. Kempf, G. Mangano, and R. B. Mann, Phys. Rev. D 52 (1995) 1108; hep-th/9412167.
[47] E. Wigner, On unitary representations of the inhomogeneous lorentz group, The Annals of Mathematics 40(1) (1939) 149.
[48] V. Bargmann and E. P. Wigner, Group theoretical discussion of relativistic wave equations, Proceedings of the National
Academy of Sciences 34(5) (1948) 211.
[49] S. Weinberg, The quantum theory of elds, Cambridge University Press, 2009.
[50] R. P. Feynman and M. Gell-Mann, Theory of the Fermi interaction, Phys. Rev. 109 (1958) 193.
[51] L. M. Brown, Two-component fermion theory, Physical Review 111(3) (1958) 957.
[52] A. A. Deriglazov, On singular Lagrangian underlying the Schrodinger equation, Phys. Lett. A 373 (2009) 3920;
arXiv:0903.1428.
[53] C. Dullemond and E. van Beveren, Canonical formalism for the relativistic harmonic oscillator, Phys. Rev. D 28 (1983)
1028.
[54] A.N. Tarakanov, Homogeneous space-times as models for isolated extended objects, arXiv:hep-th/0611149.
[55] A. Chodos, R. L. Jae, K. Jonson, C. B. Thorn and V. F. Weisskopf, New extended modal of hadrons, Phys. Rev. D 9
(1974) 3471.
[56] J. B. Conway, A course in functional analysis, Springer, 1990.

Вам также может понравиться