Sie sind auf Seite 1von 19

NIKHEF-H/92-08

Symmetries and Motions in Manifolds1


arXiv:hep-th/9205074v1 20 May 1992

J.W. van Holten2 and R.H. Rietdijk3


NIKHEF-H, P.O. Box 41882, 1009 DB Amsterdam (NL)

Abstract
In these lectures the relations between symmetries, Lie algebras,
Killing vectors and Noethers theorem are reviewed. A generalisation
of the basic ideas to include velocity-dependend co-ordinate trans-
formations naturally leads to the concept of Killing tensors. Via
their Poisson brackets these tensors generate an a priori infinite-
dimensional Lie algebra. The nature of such infinite algebras is clar-
ified using the example of flat space-time. Next the formalism is
extended to spinning space, which in addition to the standard real
co-ordinates is parametrized also by Grassmann-valued vector vari-
ables. The equations for extremal trajectories (geodesics) of these
spaces describe the pseudo-classical mechanics of a Dirac fermion.
We apply the formalism to solve for the motion of a pseudo-classical
electron in Schwarzschild space-time.

1 Lectures at the 28th Karpacz Winterschool of Theoretical Physics (Poland, 1992), by

J.W. van Holten


2 Research supported by the Stichting FOM
3 Address after oct. 1, 1992: Dept. of Mathematics, Science Labs., Univ. of Durham U.K.
1 Motions of Scalar Points in Curved Space-
Time
1.1 Introduction
In the following we connect a number of old and venerable topics related to
symmetries and conservation laws, such as Lie algebras and Noethers theorem,
with differential geometric structures like Lie derivatives and Killing vectors.
Although most of the basic ideas are well-known4 , we present extensions and
generalisations of interest in the description of certain physical systems; in par-
ticular we apply our methods to study the motion of spinning particles in a
curved space-time.
According to Einsteins equivalence principle the world line of a massive
scalar point particle in curved space-time is a time-like geodesic, described by
the equation

D2 x
= x + x x = 0, (1)
D 2
with overdots denoting proper time derivatives. Its time-like nature is expressed
by the condition
2
dx dx

ds
= g = c2 < 0, (2)
d d d
where the universal constant c (the light velocity) is a real number. In the
following we usually take c = 1 and consider particles of unit mass, but oc-
casionally we re-instate the explicit mass dependence when this is physically
relevant.
Note, that eq.(1) implies that the acceleration due to gravity is quadratic in
the four-velocity, whereas according to the Lorentz force law the acceleration of
a particle subject to electro-magnetic forces is linear in the four-velocity. For
particles coupled to force fields of higher spin one expects the acceleration to
depend on higher powers of the four-velocity [4].
The geodesic law of motion (1) can be derived from an action principle. The
simplest form of the action is
2
1
Z
S= d g (x)x x , (3)
1 2
the stationary points of which are precisely given by eq.(1). Indeed, the general
variation of S reads
Z 2
D2 x
 
d
S = d x g + (x
p ) , (4)
1 D 2 d
4 See for example refs.[1]-[3].

1
where p = g x is the canonical momentum. Thus S vanishes for any arbi-
trary variation of x with fixed end points if and only if the equations of motion
(1) are satisfied.

1.2 Symmetries and Noethers Theorem


In regard to eq.(4) we can now ask, whether there exist variations x for which
S = 0 modulo boundary terms even when the equations of motion are not
satisfied:
Z 2
d
S = d (x p J (x, x)) . (5)
1 d
Here we have already split off the total derivative coming from partial integra-
tions in the derivation of expression (4). The quantity J (x, x) is obtained from
variations x of the type
1
x = R (x, x) = R (x) + x K (x) + x x L (x) + ... (6)
2
We restrict ourselves to variations which depend on the first derivative x only,
because the second and higher derivatives can always be rewritten in terms
of these modulo the equations of motion (1). Comparing eqs.(4) and (5) one
immediately finds

dJ D2 x
= R g 0, (7)
d D 2
where the last equality holds only upon using the equations of motion. Hence
for physical motions the quantities J (x, x) are conserved. This is Noethers
theorem.

1.3 Generalised Killing Equations


Assuming that J (x, x) can be expanded in the four-velocity as

1 (2) 1 (3)
x x J (x) + x x x J (x) + ... (8)
J (x, x) = J (0) (x) + x J(1) (x) +
2 3!
we can compare terms with equal powers of the acceleration and velocity on the
left- and right-hand side of eq.(7), using the Ansatz (6) for x . This leads first
of all to an identification of the co-efficients in the expansions (6) and (8):
(1)
J (x) = R (x),

(2)
J (x) = K (x), (9)

(3)
J (x) = L (x), etc.

2
indicating that all covariant tensors on the right hand side of eq.(6) should be
taken to be completely symmetric. Secondly, the comparison shows that the
following differential equations have to be satisfied
(n)
J(1 ...n ;n+1 ) = 0, (10)
in which the parentheses denote full symmetrisation over all indices enclosed,
with total weight one. Eqs.(10) constitute a straightforward generalisation of
the Killing equation for the isometries of differentiable manifolds. Explicitly

(0)
J, = 0, (11)

R(;) = 0, (12)

K(;) = 0, etc. (13)

The first equation (11) implies that J (0) is an irrelevant constant which we
ignore from now on. The second equation (12) is the standard equation for
Killing vectors, whilst (13) and its higher-rank counterparts constitute tensorial
generalisations of this equation. Therefore one refers to K and higher-rank
tensors satisfying eq.(10) as Killing tensors.

1.4 Canonical Analysis


In terms of phase-space variables (x , p ) the conserved quantities of motion
read
1 1
J (x, p) = p R (x) + p p K (x) + p p p L (x) + ... (14)
2 3!
Suppose that there exist n independent Killing vectors Ra (x), a = 1, ..., n. Then
a general Killing vector is a linear combination

R [] = a Ra , (15)
where the a constitute a set of n linearly independent parameters. A similar
remark holds for Killing tensors. Hence a particular J , eq.(14), is in general
specified by the values of these parameters A = ( a , ...).
Introducing the fundamental Poisson brackets

{x , p } = , (16)
we can compute the Poisson bracket of two conserved Noether charges J (1)
J [1A ] and J (2) J [2A ]. The result is

3
= p R (1); R (2) R (2); R (1)

{J (1), J (2)}

1
+ p p K (1); R (2) K (2); R (1)
2

+ 2R (1); K (2) 2R (2); K (1)




1
+ p p p L (1); R (2) L (2); R (1) (17)
3!

3K ( (1)R ) (2); + 3K ( (2)R ) (1);



(
+ 3K; (1)K ) (2) 3K;
(
(2)K ) (1)

+ ...

The left-hand side being conserved for physical motions, the right-hand side
must again be of the form (14), modulo equations of motion. We now restrict
ourselves to off-shell closed algebras, i.e. the case in which the Poisson bracket
(17) itself takes the form of a Noether charge without explicit use of the equa-
tions of motion. Then

{J (1), J (2)} = J (3), (18)


where the parameters 3 are bilinear combinations of 1 and 2 :

3A = f BCA 2B 1C . (19)
From eqs.(17)-(19) it now follows, that

b
[LRa (R)] Rb; Ra Ra; Rb = f abc Rc , (20)

i a(
[LRa (K)] K i ; Ra 2R ; K
i )
= taij K j , (21)

[LRa (L)]r Lr; Ra 3Ra(


; L
r )
= g ars Ls , etc. (22)

Here the symbol L has been introduced to denote Lie derivatives. In addition,
we find relations involving higher-rank Killing tensors of the type

[LK j (K)]i K i(; K j ) K j(


; K
i )
= cijr Lr , (23)
and its further generalisations. Thus the notion of Lie derivative is extended to
higher-rank tensors.

4
Observe, that eq.(20) implies that the Killing vectors define an n-dimensional
Lie algebra. Equations like (21), (22) then assert that the Killing tensors trans-
form in linear representations of this Lie algebra defined by the structure con-
stants (taij , g ars , ...). Indeed, the Jacobi identities guarantee that these structure
constants realize the Lie algebra via their matrix commutators:
ij
t a , tb f abc tcij ,

=
(24)
rs
ga, gb f abc g crs ,

= etc.
From eq.(23) it folllows, that Killing tensors of rank n and m generate Killing
tensors of rank (n + m 1) via their tensorial Lie derivatives. In this way
one obtains in principle an infinite-dimensional algebra of conserved quantities,
unless the left-hand side of eq.(23) vanishes identically, as might happen in
special cases.

1.5 A Universal Solution


The generalised Killing equations (10) admit one solution which exists for any
arbitrary metric g . This solution is generated by the metric itself:

K (x) = g (x). (25)


It satisfies eq.(10) identically by virtue of the metric postulate g; = 0. The
constant of motion constructed from this Killing tensor is the world-line Hamil-
tonian
1
g (x) p p ,
H(x, p) = (26)
2
which is the generator of proper-time translations. From this solution we can
construct a whole tower of higher order Killing tensors by taking completely
symmetrised products of n metric tensors. The corresponding conserved quan-
tities are given by the n-th power of the Hamiltonian (26). Clearly, all these
Noether charges commute among themselves, and no other solutions are gen-
erated by their Poisson brackets. The existence of any further solutions to the
generalised Killing equations depends on the specific choice of g (x).
Note, that the method of constructing higher-rank Killing tensors out of
products of lower ones works quite generally. Indeed, from any two Killing
(n) (m)
tensors J1 ...n and J1 ...m one can construct a new Killing tensor
(n) (m)
J(n+m)
1 ...n+m
= J(1 ...n J 1 ...m ) . (27)
This satisfies the generalised Killing equation (10) because of Leibniz rule.

5
1.6 Example: Flat Space
In order to illustrate the general formalism presented above, we consider the
example of flat space, for which all conserved quantities of motion can be con-
structed explicitly. For

g (x) = , (28)
the Killing equation

R, + R, = 0 (29)
has the general solution

R [a, ] = a + x , = . (30)
Here a and are constant parameters labeling the various independent
Killing vectors (ten in four dimensions). The constants of motion corresponding
to these Killing vectors are:
1
J (1) [a, ] = a p + M , (31)
2
with M = x p x p . Clearly, the first term in eq.(31) generates a transla-
tion, whilst the second one generates Lorentz transformations.
At the level of second-rank tensors one finds in addition to symmetrised
products of Killing vectors the universal solution plus one new solution:

K [, ] = + x2 x x .

(32)
These solutions correspond to the quadratic Casimir invariants of the Poincare
algebra:
1
J (2) [, ] = p2 + M 2
. (33)
2
Since for scalar particles there exist no other independent higher order Casimir
invariants of the Poincare algebra, all other Killing tensors can now be expressed
as symmetrised products of the Killing vectors (30) and the second-rank Killing
tensors (32). For example, the Killing tensors of rank 3 which can be constructed
are of the form
 
(1) (1)
L [a(1) , (1) ] = ( a ) + ) x
(34)
  (2) (2)

L [a(2) , (2) ] = ( x2 x( x a ) + ) x .

To these tensors correspond the Noether charges

6
 
1 1  
J (3) [a(i) , (i) ] = p2 a(1) p + (1) M + M 2 a(2) p + (2) M . (35)
2 2

Clearly, the constants of motion (p , M ) form the building blocks for the
construction of the whole algebra, and the general form of the conserved charges
is
1
J (x, p) = c + a p + M, (36)
2
in which the co-efficients (c, a, ) are arbitrary functions of the quadratic Cas-
imir invariants (p2 , M 2 ).

2 Motions of Spinning Points in Curved Space-


Time
2.1 Spinning Space
In this chapter we extend the spaces considered previously with additional
fermionic dimensions, parametrised by vectorial Grassmann co-ordinates .
Following refs.[5]-[10], we take the extension in such a way, that a supersymme-
try is realised in these graded spaces; it acts on the co-ordinates as

x = i , = x . (37)
Such graded spaces were called spinning spaces in [11]. An action for the ex-
tremal trajectories (geodesics) of spinning space is
Z 2
D
 
1 i
S= d g (x)x x + g (x) , (38)
1 2 2 D
where the covariant derivative of is defined by
D
= + x . (39)
D
Under a general variation of the the co-ordinates (x , ) the action changes
by
Z 2
D2 x
  
i
S = d x g + R x
1 D 2 2
(40)

 
D d i
+ i g + x p g .
D d 2
Here the canonical momentum is

7
i
p = g x , (41)
2
whilst R is the Riemann curvature tensor. Moreover, we have simplified
the expression for S by introducing a covariantized variation of :

= + x . (42)

The action S is stationary under arbitrary variations x and vanishing at
the end points if the following equations of motion are satisfied

D2 x i
= R x , (43)
D 2 2
D
= 0. (44)
D
We briefly consider the physical interpretation of these equations. The quantity

S = i (45)
can formally be regarded as the spin-polarisation tensor of the particle [5]-
[8],[12], and correspondingly eqs.(43,44) describe the classical motion of a Dirac
particle. Eq.(44) implies that the spin tensor is covariantly constant. Equation
(43) then becomes

D2 x 1
= S R x . (46)
D 2 2
It implies, that there exist spin-dependent gravitational forces similar to the
electro-magnetic Lorentz force
q
x = F x , (47)
m
with the spin-polarisation tensor replacing the scalar electric charge [12, 13]
(here for unit mass). Such forces in principle allow the determination of spin
without any reference to the intrinsic electro-magnetic dipole moments which
are associated with it for charged particles.

2.2 Symmetries and Generalised Killing Equations


We now look for specific variations x and which leave the action off-shell
invariant modulo boundary terms. We take the variations to be of the form

8

X 1 1
x = R (x, x, ) = R(1) (x, ) + x ...xn R(n+1)
1 ...n (x, )
n=1
n!
(48)

X 1 1 (n)
= S (x, x, ) = S (0) (x, ) + x ...xn S1...n (x, )
n=1
n!

If the Lagrangian transforms into a total derivative


Z 2  
d i
S = d x p g J (x, x, ) , (49)
1 d 2
it follows that

D2 x D
 
dJ i
= R g + R x + i S g . (50)
d D 2 2 D
If the equations of motion are satisfied, the right-hand side vanishes and J is
conserved. Again, this is Noethers theorem. Otherwise, expanding J (x, x, )
in terms of the four-velocity

X 1 1
J (x, x, ) = J (0) (x, ) + x ...xn J(n)
1 ...n
(x, ) (51)
n=1
n!

and comparing the left- and right-hand side of eq.(50) with the Ansatz (48) for
x and , we find the following identities

J(n)
1 ...n
(x, ) = R(n)
1 ...n
(x, ), n 1; (52)
and
(n)
J1 ...n
S(n)
1 ...n
(x, ) = i (x, ), n 0. (53)

Moreover, these quantities have to satisfy a generalisation of the Killing equa-
tions of the form [11]

(n)
(n)
J(1 ...n i (n+1)
J(1 ...n ;n+1 ) + n+1 ) = R(n+1 J 1 ...n ) . (54)
2
(1) (2) (3)
Writing as before R = R , R = K , R = L , etc., and J (0) = B,
this reduces for the lowest components to

9
B i
B, + = R R ,
2

R( i
R(;) + = R( K ) , (55)
) 2

K( i
K(;) + ) = R( L ) , etc.
2
These equations have to hold independent of the equations of motion. The
purely bosonic (-independent) parts of these equations reduce to those we
found for the scalar particle, eqs.(11-13). In particular, the bosonic terms
in the Killing vectors R define a Lie-algebra by taking Lie-derivatives as in
eq.(20). Furthermore we note that, contrary to the bosonic case, the Killing
scalar B(x, ) = J (0) (x, ) is not always an irrelevant constant, because it can
depend non-trivially on x and , as follows from the first of eqs.(55).

2.3 Universal Solutions for Spinning Space


In contrast to the scalar particle, the spinning particle admits several conserved
quantities of motion in a general curved space-time with metric g (x) [11].
Specifically, we can construct the following four universal constants of motion:

1. Like in the bosonic case g itself is a Killing tensor:

K = g , (56)
with all other Killing vectors and tensors (bosonic as well as fermionic) equal
to zero. The corresponding constant of motion is the Hamiltonian
1
H(x, P ) =g (x)P P , (57)
2
where we have defined a covariant momentum
i
P = p + . (58)
2
2. A second obvious solution is provided by the Grassmann-odd Killing vectors

R = , T = i . (59)
Again all other Killing vectors and tensors are taken to vanish. This solution
gives us the supercharge

Q = P . (60)

10
3. In addition to ordinary supersymmetry, the spinning particle action has a
second non-linear supersymmetry, generated by Killing vectors
d
i[ 2 ]
R = g 1 ...d1 1 ... d1 ,
(d 1)!
(61)
d2
i[ 2 ]
T = g 1 ...d2 1 ... d2 .
(d 2)!
Obviously, the Grassmann parities of (R , T ) depend on the number of space-
time dimensions. The corresponding constant of motion is the dual supercharge,
given by
d
i[ 2 ]
Q = g 1 ...d P 1 2 ... d . (62)
(d 1)!
4. Finally, there exists a non-trivial Killing scalar
d
i[ 2 ]
J (0) = g 1 ...d 1 ... d . (63)
d!
This constant of motion acts as the Hodge star duality operator on . In
quantum mechanics it becomes the d+1 element of the Dirac algebra. For this
reason is refered to as the chiral charge.
From the fundamental Dirac brackets

{x , p } = ,
{ , } = ig ,
{p , } = 1
2 g g, ,

{p , p } = 4i g g, g, . (64)

we now find the following non-trivial Dirac-brackets between these universal


constants of motion:

{Q, Q} = 2i H, {Q, } = iQ . (65)


Observe, that d = 2 is an exceptional case: Q is linear and acts as an ordinary

supersymmetry:

{Q , Q } = 2i H, {Q , } = iQ. (66)
Hence in two dimensions the theory actually possesses an N = 2 supersymmetry.
For d 6= 2, the right-hand side of eqs.(66) is to be replaced by zero.

11
3 Spinning Particles in Schwarzschild Space-Time
3.1 Conservation Laws in Schwarzschild Space-Time
As an application of the generalised Killing equations for spinning space we
discuss the motion of a spinning particle in a static and spherically symmetric
gravitational field5 . The field is described by the Schwarzschild metric
  2 1
ds2 = 1  dr2 + r2 d2 + sin2 d2 ,

dt + (67)
r 1 r
where = 2M G, M being the total mass of the spherically symmetric object in
the centre of the field. It is well-known, that the Schwarzschild metric possesses
four Killing vector fields of the form

D() R() (x) ; = 0, ..., 3 (68)


where

D(0) = ,
t

D(1) = sin + cot cos ,

(69)
(2)
D = cos cot sin ,


D(3) = .

These Killing vector fields express the time-translation invariance and the rota-
tion symmetry of the gravitational field. They generate the corresponding Lie
algebra o(1,1) so(3):
 (i) (j) 
D ,D = ijk D(k) ,
 (0) (i)  (70)
D ,D = 0.
The first generalised Killing equation (55) shows, that with each Killing vec-
()
tor R there is an associated Killing scalar B () . These Killing scalars are
necessary to obtain the constants of motion

J () = B () + mx R() . (71)
The Killing scalars have a natural interpretation: the constants of motion rep-
resent the total angular momentum, which is the sum of the orbital and the
5 More details can be found in ref.[14].

12
spin angular momentum. In general, orbital angular momentum is no longer
separately conserved; therefore the Killing vector itself does not give a conserved
quantity of motion. The contribution of spin is contained in the Killing scalars,
and has to be added.
Inserting the expression for the connections and the Riemann curvature com-
ponents of the Schwarzschild space-time in eq.(55), we obtain for the Killing
scalars

i t r
B (0) = ,
2r2

B (1) = ir sin r + ir sin cos cos r ir2 sin2 cos ,

B (2) = ir cos r + ir sin cos sin r ir2 sin2 sin ,

B (3) = ir sin2 r ir2 sin cos .


(72)
Upon subtitution in J () , eq.(71), and using the spin-tensor notation introduced
in eq.(45), one finds

  dt
J (0) E =m 1 2 S rt ,
r d 2r
 
d r
 
J (1) = r sin mr +S cos cot J (3) r2 S ,
d
  (73)
d r
 
J (2) = r cos mr +S sin cot J (3) r2 S ,
d
 
d
J (3) = r sin2 mr + S r + r2 sin cos S .
d

In addition to these constants of motion, the universal conserved charges such


as the world-line hamiltonian and supercharge also provide information about
the allowed orbits of the particle. Specifically, we consider motions for which

m2 c2
H= , Q = 0. (74)
2
The first equation implies geodesic motion (equality of proper time with the
geodesic interval):

d 2 = ds2 , (75)

13
cf. eqs.(2),(67). The second equation expresses the fact that spin represents
only three independent degrees of freedom. Indeed, we can now solve for t in
terms of the spatial components i :

 
  dt t 1 dr r d d
1 = + r2 + sin2 . (76)
1 r d

r d d d

As a result, the (classical) chiral charge vanishes as well.


From eqs.(73) one can derive a useful identity

r2 sin S = J (1) sin cos + J (2) sin sin + J (3) cos . (77)
In physical terms, this identity simply states that there is no orbital angular
momentum in the radial direction.
Combining eqs.(73),(74) one obtains a complete set of first integrals of mo-
tion, expressing the velocities as functions of the co-ordinates, the spatial spin
components and the constants of motion:
 
dt 1 E rt
= + S ,
1 r m 2mr2

d
"  2
dr  2 dt
= 1 1+
d r d r

( 
2  2 )#1/2

2  d 2 d (78)
r 1 + sin ,
r d d

d 1  (1) (2) r

= J sin + J cos rS ,
d mr2
d 1 (3) 1 r 1
= 2 J S cot S ,
d mr2 sin mr m
in which

mr2
 
rt d r 2 d r
S = S + sin S . (79)
E d d
Finally, the rate of change of the spins is determined by

14
d r
  
3 d d
= r 1 + sin2 ,
d 2r d d

d
 
1 dr d r d
= + + sin cos , (80)
d r d d d

d
 
1 dr d 1 d r d
= + cot cot .
d r d d r d d

Eqs.(78-80) can be integrated to give the full solution of the equations of motion
for all co-ordinates and spins.

3.2 Special Solutions


As an application of the results obtained in sect.(3.1) we study the special case of
motion in a plane, for which we choose = /2. Unlike for scalar point particles,
this is not the generic case, because in general orbital angular momentum is not
conserved separately.
For spinning particles, motion in a plane occurs in two kind of situations.
The first possibility is radial motion, for which = 0. In this case there is no
orbital angular momentum and spin is conserved independently. Clearly, either
the particle escapes to infinity or hits the centre of the potential after a finite
time.
The second possibility concerns motion for which 6= 0. In this case orbital
and spin angular momentum decouple if they are parallel. Hence we impose the
conditions

S = 0, S r = 0, (81)
(1,2)
from which it now follows that J vanish. In this case we have two constants
of motion:

L = mr2 , J (3) L = rS r . (82)


From the first of eqs.(78) we now find a formula for the gravitational redshift
given by
 
d E
dt = + L . (83)
1 r m 2mEr3


The first term corresponds to the usual time dilation in a gravitational field for
spinless particles. In this case, there is an additional contribution from spin-
orbit coupling. This shows, that time dilation is not a purely geometrical effect,
but also has a dynamical component [12].

15
From the second of eqs.(78) and the first eq.(82) we obtain the equation for
the orbit of the particle:
s
E 2 m2 2

m mr J (3)
 
1 dr
= r 1 + + . (84)
r d L2 L L mr
In terms of dimensionless quantities
E r
= , x = ,
m
(85)
L
= , = ,
m L
we find for the stationary points xm of the orbit, as defined by the vanishing of
the right-hand side of eq.(84):

1 2 2
2 = 1 + 2 3 (1 + ) . (86)
xm xm xm
Of course, all calculations presented here are rather formal, as is not a pure
number. However, the equations might be applicable to realistic physics situa-
tions if it is allowed to replace in certain limiting cases by a real number. Since
the pseudo-classical equations acquire physical meaning when averaged over in
functional integrals [5, 15] (i.e. in the path integral of the quantum Dirac parti-
cle), such a limit might arise in the semi-classical regime of the quantum theory,
as implied by the correspondence principle. In the following we assume that
such a numerical value of has been obtained and leads to valid results, at
least in expansions to first order in (where the fact that 2 = 0 plays no
role).
For 1 we have open orbits for which there is at most one point of closest
approach, the perihelion. If for fixed the energy exceeds a critical value, the
particle can cross into the central region of the potential (x < 1). The critical
value is given by

2crit = 2 (xm
()
), (87)
where
r !
() 3(1 + )
xm = 2 1 1 . (88)
2
For < 1 there are bound states corresponding to quasi-periodic orbits, which
have both a perihelion and an aphelion. The stationary points of the orbit are
again determined by6 eq.(86). Solutions of this equation exist only for
6 For the special case of circular motion, this equation determines the radius of the orbit,

which now is of course independent of .

16
2 3 (1 + ) . (89)
In particular, for circular orbits there exists a minimal radius given by

xm = 2 = 3 (1 + ) . (90)
For this critical orbit the energy is to first order in :
1
2crit = (8 + ) , (91)
9
whilst the time-dilation factor in the critical orbit is given by

   
dt 3
= 2 1 . (92)
d crit 8
For non-circular motion one finds that the perihelion of the orbit precesses as for
a spinless particle, but at a different, spin dependent rate. In the weak-coupling
limit (slow precession of the perihelion), we obtain for the precession angle after
one period
3
= (1 + ) , (93)
k
where k is the semilatus rectum of the elliptical orbit, and we have neglected
terms of O(k 2 ). Hence in this approximation spin-dependent effects disappear
for = O(k p ) whenever p 1. Note, that spin-dependent gravitational effects
can be larger or smaller than for a spinless particle depending on the sign of ,
i.e. the relative orientation of L and . This we interpret as a classical analogue
of fine splitting.
As a final remark, we observe that even if an a priori numerical value for
cannot be assigned, its appearance in various places like in eqs.(91-93) still
allows the pseudo-classical theory to make quantitative predictions by com-
paring different physical processes in the regime where the semi-classical limit
applies. For a discussion of some related issues in the case of spinning particles
in electro-magnetic fields we also refer to [12].

References
[1] C.W. Misner, K.S. Thorne and J.A. Wheeler, Gravitation (W.H. Freeman
Co., San Francisco, 1973)
[2] A. Papapetrou, Lectures on General Relativity (Reidel, Dordrecht, 1974)
[3] D. Kramer, H. Stephani, M. MacCallum and E. Herlt, Exact solutions of
Einsteins field equations (Cambridge Univ. Press, 1980)

17
[4] B. de Wit and D.Z. Freedman, Phys. Rev. D21 (1980), 358
[5] F.A. Berezin and M.S. Marinov, Ann. Phys. (N.Y.) 104 (1977), 336
[6] R. Casalbuoni, Phys. Lett. B62 (1976), 49
[7] A. Barducci, R. Casalbuoni and L. Lusanna, Nuov. Cim. 35A (1976), 377

[8] A. Barducci, R. Casalbuoni and L. Lusanna, Nucl. Phys. B124 (1977), 521
[9] L. Brink, S. Deser, B. Zumino, P. Di Vecchia and P. Howe, Phys. Lett. 64B
(1976), 43
[10] L. Brink, P. Di Vecchia and P. Howe, Nucl. Phys. B118 (1977), 76
[11] R.H. Rietdijk and J.W. van Holten, Class. Quantum Grav. 7 (1990), 247
[12] J.W. van Holten, Nucl. Phys. B356 (1991), 3; Physica A 182 (1992), 279
[13] I.B. Khriplovich, Novosibirsk preprint INR-89-1
[14] R.H. Rietdijk and J.W. van Holten, Spinning Particles in Schwarzschild
Space-Time; NIKHEF-H preprint (to appear)
[15] A. Barducci, R. Casalbuoni and L. Lusanna, Nucl. Phys. B180 [FS2] (1981),
141

18

Das könnte Ihnen auch gefallen