Você está na página 1de 17

January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012

International Journal of Modern Physics A


Vol. 28, No. 1 (2013) 1350001 (17 pages)
c World Scientic Publishing Company
DOI: 10.1142/S0217751X13500012
LIGHT-CONE FLUCTUATIONS AND THE
RENORMALIZED STRESS TENSOR OF
A MASSLESS SCALAR FIELD
V. A. DE LORENCI
Instituto de Ciencias, Universidade Federal de Itajuba,
Itajuba, MG 37500-903, Brazil
delorenci@unifei.edu.br
G. MENEZES

Instituto de Fsica Teorica, Universidade Estadual Paulista,


Rua Dr. Bento Teobaldo Ferraz 271 Bloco II,
Sao Paulo, SP 01140-070, Brazil
gsm@ift.unesp.br
N. F. SVAITER
Centro Brasileiro de Pesquisas Fsicas,
Rua Dr. Xavier Sigaud 150, Rio de Janeiro, RJ 22290-180, Brazil
nfuxsvai@cbpf.br
Received 26 September 2012
Revised 21 November 2012
Accepted 7 December 2012
Published 3 January 2013
We investigate the eects of light-cone uctuations over the renormalized vacuum expec-
tation value of the stressenergy tensor of a real massless minimally coupled scalar eld
dened in a (d+1)-dimensional at spacetime with topology RT
d
. For modeling the
inuence of light-cone uctuations over the quantum eld, we consider a random Klein
Gordon equation. We study the case of centered Gaussian processes. After taking into
account all the realizations of the random processes, we present the correction caused
by random uctuations. The averaged renormalized vacuum expectation value of the
stressenergy associated with the scalar eld is presented.
Keywords: Quantum elds in curved spacetime; Casimir energy; wave propagation in
random media.
PACS numbers: 03.70.+k, 04.62.+v, 11.10.Gh, 42.25.Dd

Corresponding author.
1350001-1
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
1. Introduction
The concept of spacetime in special relativity is based on the geometric construc-
tion of xed light-cones, which divides spacetime into causally distinct regions.
On the other hand, within the general relativity framework, the causal structure of
events is not xed for all times; rather, it is dynamical. Such a geometrical picture
of the gravitational eld is a very successful classical eld theory. There are many
attempts to bring general relativity into the quantum domain.
1,2
Despite such enor-
mous eorts, so far there is no consensus on what would be a quantum theory of
gravity. One should expect that one of the consequences of assuming that the gravi-
tational eld obeys quantum-mechanical laws is that the structure of spacetime
must undergo quantum uctuations. Ford and collaborators developed this idea,
39
showing that the eects of uctuations of the geometry of the spacetime caused by
quantum mechanical uctuations of the gravitational eld is to smear out the light
cone. In general, in addition to quantum mechanical metric uctuations, there are
induced metric uctuations generated by quantum uctuations of matter elds. In
both scenarios the concept of light-cone structure has to be modied.
On the other hand, it is well known that when quantum elds are forced to
obey classical boundary conditions, interesting eects arise which are connected
with vacuum uctuations of the eld. Even if the spacetime is unbounded, if the
quantum eld is constrained by the presence of material boundaries, nontrivial con-
sequences will show up due to vacuum uctuations. In this respect, the most well
known physical manifestation of such uctuations is the Casimir eect, which has
been extensively discussed in the literature.
1015
In such a context, an intriguing
question would be how light-cone uctuations could aect measurable eects asso-
ciated with virtual processes in quantum theory. Here we provide a scenario where
this issue is investigated.
In this paper, we consider the eects of light-cone uctuations upon the renor-
malized vacuum expectation value of the stressenergy tensor of a quantum eld. In
such a direction, recently it was investigated the inuence of light-cone uctuations
over the transition probability rate of a two-level system coupled to a massless scalar
eld undergoing uniformly accelerated motion.
16
The assumptions made in such a
reference were that the light-cone uctuations can be treated classically and their
eects on the quantum elds can be described via random dierential equations.
For a similar situation involving massless fermions see Ref. 17.
Here we study the renormalized vacuum expectation value of the stress tensor
associated with a real massless minimally coupled scalar eld in the presence of a
disordered medium. For modeling the inuence of light-cone uctuations over the
quantum eld, we consider a random KleinGordon equation.
18
We study the case
of centered Gaussian processes. We consider elds dened in a nonsimply connected
spacetime with topology of T
d
!
1
, with T
d
corresponding to a hypertorus, i.e.
a at spacetime with periodicity in all of the d spatial dimensions. In order to
obtain our results we have to implement a perturbation theory associated with a
1350001-2
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
massless scalar eld in disordered media.
1921
One could also regard this situation
as a simplied model for the more realistic case of the Casimir energy due to
phonons
2224
in a random uid conned between two innite walls. In this case the
inuence of the light-cone uctuations over surface divergences can be investigated.
The organization of the paper is as follows. In Sec. 2, we present a brief review
of the point-splitting approach that can be used to obtain the renormalized vacuum
energy of quantum elds in the presence of classical macroscopic boundaries and
also in curved spacetime. We discuss the modications in a free scalar quantum
eld theory due to the presence of randomness in Sec. 3. In Sec. 4, we derive
the correction caused by random uctuations in the renormalized vacuum expec-
tation value of the stress tensor associated with quantum elds in the case men-
tioned above. Conclusions are given in Sec. 5. In this paper, we use 32G = =
k
B
= c = 1.
2. Renormalized Vacuum Expectation Value of the Stress Tensor
of Quantum Fields
The problem of renormalization of ill dened quantities leading to a physically
meaningful result is a fundamental question in quantum eld theory. Although for-
mally divergent, in the absence of gravity the dierence between the vacuum energy
of quantum elds at dierent physical congurations can be nite.
10
Being more
precise, in a global approach, the Casimir energy can be obtained adopting the
following prescription: the eigenfrequencies of the mode solutions of the classical
wave equation satisfying appropriate boundary conditions are found; the diver-
gent zero-point energy of the quantized eld is regularized and then renormalized
using auxiliary congurations which are added and subtracted. For example, the
introduction of a pair of conducting plates into the vacuum of the electromagnetic
eld alters the zero-point uctuations of the eld and thereby produces an attrac-
tion between the plates.
1115
In this paper, we are interested to study the eects
of light-cone uctuations over the renormalized vacuum expectation value of the
stress tensor.
There are dierent ways to nd the renormalized vacuum energy of quantum
elds dened in a at spacetime with classical boundary conditions or in generic
curved spacetimes.
2540
Of our particular interest is the point-splitting regular-
ization method. In this approach we compute the vacuum energy from the renor-
malized stress tensor. For a minimally coupled massless spin-0 eld dened in at
spacetime with a given interaction potential V (), the stress-tensor reads
T

(x) =

(x)

(x)
1
2

(x)

(x) +

V () . (1)
The potential V () may depend on products of elds and eld derivatives. Here

is the usual Minkowski metric (we take the sign convention of Ref. 41). For
a quantum eld propagating in a medium with disorder, V () will represent the
coupling between the eld and random impurities. In this paper, we assume that
1350001-3
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
such a random potential has the following functional form:
V () =
1
2
(x)
0
(x)
0
(x) . (2)
The statistical properties of the random variable (x) which describes randomness
will be given in due course. Hence, using the point-splitting method, the vacuum
expectation value of the stress tensor is found to be
T

(x) = lim
x

,x

x
T

(x

, x

) , (3)
where
T

(x

, x

) =
1
4
_
2

G
(1)
(x

, x

G
(1)
(x

, x

)
+

(x)
0

0
G
(1)
(x

, x

, (4)
where the Greens function G
(1)
(x, x

) is given by
G
(1)
(x, x

) = (x), (x

) = G
(+)
(x, x

) + G
()
(x, x

) , (5)
with the Wightman functions given by the expressions G
(+)
(x, x

) = (x)(x

)
and G
()
(x, x

) = (x

)(x). In Eq. (4), it is to be understood that

) acts
on x

(x

). We chose to write such an equation in a form which is symmetric under


the interchange x

. We remark that we are using the summation convention,


i.e. repeated indices are summed unless otherwise stated. Greek indices are referred
to spacetime components (e.g. , , , . . . = 0, 1, . . . , d) whereas latin indices stand
for space components (e.g. i, j, r, . . . = 1, 2, . . . , d).
The coincidence limits for G
(1)
(x, x

) and its derivatives yield formally diver-


gent expressions. Therefore (3) should be properly renormalized. Considering a
at background spacetime, this could be done in the usual way, i.e. by replacing
such a quantity by the renormalized expectation value of the stress tensor which is
given by
: T

(x): = T

(x) T

(x)
M
, (6)
where T

(x)
M
is the expectation value of the stress tensor in Minkowski space
time. In this way, the renormalized vacuum energy density is given by : T
00
(x): .
We remark that such a procedure cannot be trusted when the background space
time is curved. In nongravitational physics, only energy dierences are observable; in
this case the subtraction scheme can be carried out without further inconveniences.
When one considers gravity, this technique is not satisfactory since energy is a
source of gravity and therefore we are not free to rescale the zero-point energy.
In Sec. 3, we discuss a scalar quantum eld theory in the presence of stochastic
uctuations of the light cone.
1350001-4
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
3. Scalar Quantum Field Theory in the Presence of
Random Fluctuations
In this section, we present the solution to the scalar eld equation in the presence of
light-cone uctuations. As pointed out, for modeling the inuence of such uctua-
tions over the quantum eld, it is enough to consider a stochastic KleinGordon
equation. The eld equation obtained in this way cannot be solved in a closed
form. However, assuming that stochastic uctuations are small, one may introduce
a perturbation theory similar to the one discussed in Ref. 21. So, we will have an
expression for the Hadamard function that will contain the corrections due to light-
cone uctuations. This will enable us to calculate the vacuum expectation value of
the stress tensor and, therefore, the corrections to the Casimir energy.
Before we describe the physical situation which will lead us to the vacuum
energy, let us present the eld equation for a massless minimally coupled scalar
eld in a spacetime with stochastic light-cone uctuations. It reads
21
_
[1 + (x)]

2
t
2

2
_
(x) = 0 . (7)
We remark that such an equation can be derived from a Lagrangian containing the
random potential given by (2), assuming that the inhomogeneity and disorder are
smooth, i.e. (x) is a slowly-varying function in comparison with . In this way,
derivatives of the (x) are neglected. The important point to be observed is that
we are not considering a toy model for quantum gravity; rather, we are interested
in how quantum elds behave in the presence of random uctuations of the light
cone.
For simplicity, we consider the (local) random variable (x) to be a Gaussian
centered distribution:
(x) = 0 , (8)
with a white-noise correlation function given by
(x)(x

) =
2
(x x

) , (9)
where
2
gives the intensity of stochastic uctuations and (x x

) is the (d + 1)-
dimensional Dirac delta function. The symbol ( ) denotes an average over all
possible realization of the random variable. On the other hand, in principle it is
possible to extend the method to colored or non-Gaussian noise functions. In addi-
tion, observe that the noise sources dene a preferred reference frame, similarly to
an external heat bath, which induces a breaking in Lorentz symmetry.
The eld equation (7) should be compared with the ones used in Refs. 18 and
21. However, in contrast to these references, which employ a static noise, here
we assume that the random function is also time-dependent. This situation was
carefully analyzed in Ref. 42. Since the solution to the eld equation cannot be
given in a closed form, one can employ a perturbative series expansion for Greens
1350001-5
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
functions. The propagator iG(x, x

) = T((x)(x

)) associated with the wave


equation (7) satises
_
[1 + (x)]

2
t
2

2
x
_
G(x, x

) = (t t

)
d1
(x x

) . (10)
Suitable boundary conditions must be imposed on the solutions of the above equa-
tion in order for them to have the properties of a time-ordered product. In order
to accommodate the appropriate modications introduced by the presence of the
random term, we proceed as follows. Dene the operator K = K
0
L, where
K
0
(x, y) =
_

2
/t
2

2
x
i
_
(x y) and
L(x) = (x)

2
t
2
. (11)
This leads to the following formal relation:
[K(x, y)]
1
= G(x, y) . (12)
In this way, since the disorder is weak, a natural perturbative expansion for G in
form of a Dyson series can be dened:
G = G
0
G
0
LG
0
+ G
0
LG
0
LG
0
+ , (13)
where we used a formal operator notation in this last equation and [K
0
]
1
= G
0
is the unperturbed propagator. Following Ref. 21, Eq. (13) can be written in terms
of space and time variables:
G(x, x

) = G
0
(x, x

) +

n=1
_
dy
1
G
0
(x, y
1
)(
(n)
(y
1
, x

) , (14)
where
(
(n)
(y
1
, x

) = (1)
n
n

j=1
L(y
j
)
_
dy
j+1
G
0
(y
j
, y
j+1
) . (15)
In Eq. (15), it is to be understood that y
n+1
= x

and that there is no integra-


tion in y
n+1
. Details on the derivations of the above expressions can be found in
Ref. 21. Furthermore, due to the Gaussian nature of the noise averaging, higher-
order correlation functions of the form (x
1
)(x
2
) (x
p
) can be easily expressed
as the sum of products of two-point correlation functions corresponding to all pos-
sible partitions of x
1
, x
2
, . . . , x
p
.
Since we are working in the weak-disorder regime, we consider terms up to
second-order in of the above series and disregard any higher-order corrections.
The calculation for G is discussed at length in App. A. With an expression for
the propagator, one is able to calculate the Greens function G
(1)
(x, x

) through
the following formulae. The propagator can be written in terms of the Wightman
functions as iG(x, x

) = (t t

)G
(+)
(x, x

) + (t

t)G
()
(x, x

). In turn, since
G
(1)
(x, x

) is given by (5), one sees that the calculations of the propagator allows
one to reach expressions for the Greens function G
(1)
(x, x

) in situations where one


1350001-6
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
can use the decomposition property stated above. On the other hand, G
(1)
(x, x

)
can also be computed through the relation
41
G(x, x

) +
1
2
_
G
R
(x, x

) + G
A
(x, x

=
i
2
G
(1)
(x, x

) , (16)
where G
R
(x, x

) = i(t t

)[(x), (x

)] (G
A
(x, x

) = i(t

t)[(x), (x

)])
is the retarded (advanced) Greens function which obeys [K]
1
= G
R(A)
.
After this digression on Greens functions, we are able to calculate the correc-
tions to the expectation value of the vacuum energy due to randomness of the light
cone. This is the subject of Sec. 4.
4. The Averaged Renormalized Vacuum Expectation Value of the
Stress Tensor
The aim of this section is to determine the averaged components of the renormalized
vacuum expectation value of the stress tensor. Such quantities will be calculated in
the way prescribed in Sec. 2. As mentioned earlier, the vacuum energy density is
given by T
00
(x). Other stress-tensor components have well known physical inter-
pretation. We should average such quantities over all the realizations of the noise.
Therefore, after performing the stochastic averages of (4), the vacuum expectation
values of the stress-tensor components are given by the coincidence limit of the
following expression:
T

(x

, x

) =
1
4
_
2

G
(1)
(x

, x

G
(1)
(x

, x

)
+


0

0
(x)G
(1)
(x

, x

)
_
. (17)
Now we focus our attentions on calculating the components of the stress tensor con-
sidering the eects of light-cone random uctuations. We consider the case where
the elds satisfy periodic boundary conditions in all spatial directions. In this way,
since the background spacetime is at, we may use the subtraction scheme dis-
cussed in Sec. 2. The corrections to the Greens function G
(1)
(x, x

) up to second-
order in the perturbations are given in App. A. Let us rst present the zero-order
contribution. Inserting expression (A.16) in (17) and remembering Eqs. (3) and (6)
lead to the following expression for the renormalized vacuum energy density:
: T
00
(x):
0
=
1
2a
1
+

n
1
=

1
a
d
+

n
d
=
k(n) , (18)
where the n = 0 term is excluded from the above sum as it is just the contribution
from the Minkowski vacuum. Following the discussion presented in App. A, we
write this multiple sum in terms of the Epstein zeta function. Using (A.11) and
dening
f(d) =

_
d+1
2
_
2
(d+1)/2
, (19)
1350001-7
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
the renormalized vacuum energy density becomes
: T
00
(x):
0
= f(d)Z(a
1
, . . . , a
d
; d + 1) . (20)
The result (20) is nite for all d > 0 and is always negative. To reach such an
expression we must remember to introduce an arbitrary mass parameter in the
summations above in order to keep the Epstein zeta function a dimensionless quan-
tity. This procedure is necessary in order to enable one to implement the analytic
procedure described in App. A.
Similarly,
: T
ij
(x):
0
=
1
4a
1
+

n
1
=

1
a
d
+

n
d
=
1
k(n)
_
2k
i
k
j

ij

rs
k
r
k
s

ij
k
2
(n)

, (21)
where k
j
= 2n
j
a
1
j
. For i = j, one has
: T
jj
(x):
0
=
1
2a
1
+

n
1
=

1
a
d
+

n
d
=
k
2
j
k(n)
=
1
2da
1
+

n
1
=

1
a
d
+

n
d
=
k(n) , (22)
where the last equality follows by symmetry. In the above equation it is to be
understood that there is no summation over the repeated index j. Proceeding as
earlier, one has
: T
jj
(x):
0
=
f(d)
d
Z(a
1
, . . . , a
d
; d + 1) . (23)
Now consider i ,= j. One has
: T
ij
(x):
0
=
1
2a
1
+

n
1
=

1
a
d
+

n
d
=
k
i
k
j
k(n)
. (24)
Since we have sums over the integers of a product between an even function and
an odd function, we have that : T
ij
(x):
0
= 0 for i ,= j.
As for the zero-order contribution to momentum density, one has, after con-
sidering the coincidence limit and using (6),
: T
0j
(x):
0
=
1
2a
1
+

n
1
=

1
a
d
+

n
d
=
k
j
. (25)
The sums over indices other than j in the above expression can be expressed with the
help of a integral representation for the Epstein zeta-function. It gives an analytic
1350001-8
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
continuation for such series except for a pole at p = 2s.
43
It is
()
s
(s)Z
p
(a
1
, . . . , a
p
; 2s)
=
1
s
+
2
p 2s
+
s
_

dxx
s1
_

_
0, . . . , 0; a
2
1
x, . . . , a
2
p
x
_
1
_
+
2sp
2
_

1/
dxx
(p2s)/21
_

_
0, . . . , 0;
x
a
2
1
, . . . ,
x
a
2
p
_
1
_
, (26)
where
p/2
is the product of the ps parameters a
i
given by
p/2
= a
1
a
p
, and
the generalized Jacobi function (z
1
, . . . , z
p
; x
1
, . . . , x
p
) is dened by
(z
1
, . . . , z
p
; x
1
, . . . , x
p
) =
p

i=1
(z
i
; x
i
) , (27)
with (z; x) being the Jacobi function, i.e.
(z; x) =

n=
e
(2nzn
2
x)
. (28)
Using this integral expression for the Epstein zeta-function, given by (26), we can
nd that
Z
p
(a
1
, . . . , a
p
; 2s)[
s=0
= 1 , (29)
for any p 1. So
: T
0j
(x):
0
=
1
2a
1
a
d
+

n
j
=
k
j
. (30)
Such a summation is zero, as the reader can easily check. So : T
0j
(x):
0
= 0.
Now let us introduce the corrections due to light-cone random uctuations.
First consider the corrections to the renormalized vacuum energy density. Inserting
Eqs. (A.17) and (A.20) in (17) and taking the coincidence limit for the component
T
00
(x

, x

) yields
: T
00
(x):
1
= P + E , (31)
where
P =

2
8a
1
+

n
1
=

1
a
d
+

n
d
=
1
1
a
1
+

m
1
=

1
a
d
+

m
d
=
k
2
(m) (32)
and
E =
2
Q(a
1
, . . . , a
d
; d)
1
2a
1
+

n
1
=

1
a
d
+

n
d
=
k(n) . (33)
1350001-9
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
Consider the quantity P. The sums over n can be evaluated considering that
+

n
1
=

+

n
d
=
1 = Z
d
(a
1
, . . . , a
d
; 2s)

s=0
.
So, with the help of (29), we get
P =

2
8a
1
+

m
1
=

1
a
d
+

m
d
=
k
2
(m) . (34)
The sum over m can also be expressed in terms of the Epstein zeta-function
+

m
1
=

+

m
d
=
k
2
(m) = (2)
2
Z
d
(a
1
, . . . , a
d
; 2) ,
which vanishes, in virtue of the functional reection formula (A.11). So P = 0.
Consider now E. Comparing Eqs. (18) and (33), one has
E =
2
[f(d)Z(a
1
, . . . , a
d
; d + 1)]
2
, (35)
where we have used Eqs. (19) and (A.14). Therefore, collecting our results one has
: T
00
(x):
1
=
2
[f(d)Z(a
1
, . . . , a
d
; d + 1)]
2
. (36)
This is the correction to the renormalized vacuum energy due to light-cone uc-
tuations up to second order in the noise. The subscript 1 on the left-hand side
of the above equation indicates the rst-order correction after performing the ran-
dom averages. Now let us calculate the corrections to the components : T
ij
(x): .
Inserting Eqs. (A.17) and (A.20) in (17), one has
: T
ij
(x):
1
=

2
8
Q(a
1
, . . . , a
d
; d)
1
a
1
+

n
1
=

1
a
d
+

n
d
=
1
k(n)

_
2k
i
k
j

ij

rs
k
r
k
s
3
ij
k
2
(n)

+
ij
P . (37)
Considering the results derived above one has, for i ,= j, : T
ij
(x):
1
= 0. For i = j,
one has (no summation over the index j)
: T
jj
(x):
1
=

2
4
Q(a
1
, . . . , a
d
; d)
1
a
1
+

n
1
=

1
a
d
+

n
d
=
k
2
j
+ k
2
(n)
k(n)
=

2
4
(d + 1)
d
Q(a
1
, . . . , a
d
; d)
1
a
1
+

n
1
=

1
a
d
+

n
d
=
k(n) , (38)
1350001-10
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
where the last line of the right-hand side of the above expression follows by sym-
metry. Hence, using the same technique as earlier, one has
: T
jj
(x):
1
=

2
2
_
1 +
1
d
_
[f(d)Z(a
1
, . . . , a
d
; d + 1)]
2
. (39)
With similar considerations as before, one may show that all corrections to the
momentum density vanish, : T
0j
(x):
1
= 0. Consequently, the nal form for the
renormalized expectation value of stress-tensor components in a nonsimply con-
nected spacetime subjected to light-cone uctuations reads, up to second order in
the perturbations,
: T
00
(x): = g(a, d)[1 +
2
g(a, d)] (40)
and
: T
jj
(x): = g(a, d)
_
1
d
+

2
2
_
1 +
1
d
_
g(a, d)
_
, (41)
where a = (a
1
, a
2
, . . . , a
d
) and
g(a, d) = f(d)Z(a
1
, . . . , a
d
; d + 1) . (42)
These expressions summarize the main results of the paper. Let us now discuss the
results presented here.
5. Discussions and Conclusions
In this paper, we studied a massless scalar eld theory in the presence of light-cone
uctuations. After performing the random averages over the noise function, the
correction caused by randomness in the renormalized stress tensor associated with
the quantum eld was presented. We obtained a correction which is proportional
to the square of the unperturbed contribution. Although calculations show that
contributions of light-cone uctuations to the renormalized stress tensor are quite
small, one cannot exclude the possibility of the existence of some mechanism which
is able to amplify such uctuations giving rise to nontrivial measurable eects.
Although the Casimir eect is a well understood phenomenon, there are still
some open questions related to this eect. A interesting question is how the sign of
the Casimir force depends on the topology, dimensionality of the spacetime, the
shape of bounding geometry or others physical properties of the system.
4346
This
problem is still unsolved in the literature. There are also some controversies in the
literature that inspired many recent papers. For example, questions concerning the
temperature dependence of real materials and also how to obtain closed-form results
for the interaction of bodies that alters the zero-point energy of the electromagnetic
eld. Here we did not consider such problems.
We stress that here we present all the technical details concerning how to
evaluate the eects of light-cone uctuations over the renormalized stress tensor
associated with a quantum eld. A further step is to use the approach developed
1350001-11
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
in this paper in a more realistic physical situation. An interesting example is the
renormalized vacuum energy due to phonons in a disordered uid conned between
plane boundaries. Phonons share several properties with relativistic quantum elds.
Quantized acoustic perturbations in the presence of disorder and boundaries lead
us to the phononic Casimir eect with randomness. Other intriguing situation is
the case of a quantized electromagnetic eld in the presence of boundaries. Ford
and Sopova calculated the stress tensor for the quantized electromagnetic eld in
the region between a pair of dispersive, dielectric half-spaces.
47
They showed that
although the vacuum energy density is nite between perfectly reecting plates, it
will diverge near plates of nite reectivity. It is interesting to investigate whether
such behavior can also be obtained in the presence of light-cone uctuations. These
subjects are under investigation by the authors.
Acknowledgments
We thank E. Arias for useful discussions. This paper was supported by the Brazilian
agencies CAPES, CNPq and FAPEMIG.
Appendix A. Perturbative Corrections to the Feynman
Propagator and the Hadamard Function for a
Nonsimply Connected Space Time
Our aim is to present an expression for the Hadamard function from which we
can calculate corrections to T

due to light-cone uctuations. Let us present the


propagator up to second order in (x). From (14), we have
G(x, x

) = G
0
(x, x

)
_
dy G
0
(x, y)L(y)G
0
(y, x

)
+
__
dy
1
dy
2
G
0
(x, y
1
)L(y
1
)G
0
(y
1
, y
2
)L(y
2
)G
0
(y
2
, x

) . (A.1)
We consider a topology of the background spacetime such that the elds must
satisfy periodic boundary conditions in all spatial directions. For a general hyper-
cuboidal space, with sides of nite length a
1
, . . . , a
d
, this corresponds to the com-
pactication of the space dimensions to a hypertorus T
d
. The modes of the eld
then consist of a simple product of modes analogous to the usual Minkowski space.
In this way, the unperturbed propagator reads
G
0
(x, x

) =
1
a
1
+

n
1
=

1
a
d
+

n
d
=
_
d
(2)
e
i[k(n)(xx

(tt

)]

2
k
2
(n) + i
, (A.2)
with
k(n) = 2N = 2
_
n
1
a
1
,
n
2
a
2
, . . . ,
n
d
a
d
_
(A.3)
1350001-12
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
and
k
2
(n) = k
2
(n) = k
2
1
+ + k
2
d
=
_
2n
1
a
1
_
2
+ +
_
2n
d
a
d
_
2
, (A.4)
where use has been made of the notation n
2
= n
2
1
+ n
2
2
+ + n
2
d
. In order to
perform the integration we may resort to contour integrals. We choose the usual
contour for the Feynman propagator (see for instance Ref. 41). We get
G
0
(x, x

) =
i
a
1
+

n
1
=

1
a
d
+

n
d
=
1
2k(n)

_
e
i[k(n)(xx

)k(n)(tt

)]
(t t

)
+ e
i[k(n)(xx

)k(n)(tt

)]
(t

t)
_
. (A.5)
Now let us introduce the corrections due to light-cone random uctuations. In virtue
of Eqs. (8) and (9), the rst correction to G will be given by the third term on the
right-hand side of (A.1). Then
G
2
(x, x

) =
__
dy
1
dy
2
G
0
(x, y
1
)L(y
1
)G
0
(y
1
, y
2
)L(y
2
)G
0
(y
2
, x

) , (A.6)
where the subscript in G stands for nth-order in (x). Inserting Eqs. (11) and (A.2)
in the last expression and then using (9) allow us to write
G
2
(x, x

) =
1
a
1
+

n
1
=

1
a
d
+

n
d
=
_
d
2
e
i[k(n)(xx

)(tt

)]

1
(
2
k
2
(n) + i)
()
1
(
2
k
2
(n) + i)
, (A.7)
where
() = lim
0

2
1
a
1
+

n
1
=

1
a
d
+

n
d
=
_
d

2
e
i

2
k
2
(n) + i
. (A.8)
Performing the integration as earlier we get, after taking 0,
() =
i
2

2
2
1
a
1
+

n
1
=

1
a
d
+

n
d
=
k(n) . (A.9)
Such a multiple sum may be written in terms of the Epstein zeta function:
37
Z(a
1
1
, . . . , a
1
p
; s) =
+

n
1
,...,n
p
=
__
n
1
a
1
_
2
+
_
n
2
a
2
_
2
+ +
_
n
p
a
p
_
2
_
s/2
(A.10)
for s > p and it should be understood that the term for which all n
i
= 0 is to be
omitted. This function obeys the reection formula
43

_
s
2
_

s/2
Z(a
1
, . . . , a
p
; s) =

(sp)/2
a
1
a
p

_
p s
2
_
Z(a
1
1
, . . . , a
1
p
; p s) . (A.11)
1350001-13
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
Using Eqs. (A.10) and (A.11), we have
() = i
2

2
Z(a
1
, . . . , a
d
; d + 1)

_
d+1
2
_
2
(d+1)/2
. (A.12)
Then, after performing the integration,
G
2
(x, x

) =

2
2
Q(a
1
, . . . , a
d
; d)
1
a
1
+

n
1
=

1
a
d
+

n
d
=

__
1
2k(n)

i(t t

)
2
_
e
i[k(n)(xx

)k(n)(tt

)]
(t t

)
+
_
1
2k(n)

i(t

t)
2
_
e
i[k(n)(xx

)k(n)(tt

)]
(t

t)
_
, (A.13)
where
Q(a
1
, . . . , a
d
; d) = Z(a
1
, . . . , a
d
; d + 1)

_
d+1
2
_
2
(d+1)/2
. (A.14)
Employing the decomposition property of the propagator in terms of the Wightman
functions, one has
iG
0
(x, x

) + G
2
(x, x

) = T((x)(x

))
= (t t

)G
(+)
(x, x

) + (t

t)G
()
(x, x

) .
Hence, employing (5), one sees that
G
(1)
(x, x

) = G
(1)
0
(x, x

) + G
(1)
2
(x, x

) , (A.15)
with the unperturbed Greens function G
(1)
(x, x

) given by
G
(1)
0
(x, x

) =
1
a
1
+

n
1
=

1
a
d
+

n
d
=
cos[k(n) (x x

) k(n)(t t

)]
k(n)
. (A.16)
Therefore, the respective correction to G
(1)
(x, x

) is
G
(1)
2
(x, x

) =

2
2
Q(a
1
, . . . , a
d
; d)
1
a
1
+

n
1
=

1
a
d
+

n
d
=

_
cos[k(n) (x x

) k(n)(t t

)]
k(n)
+ (t t

) sin[k(n) (x x

) k(n)(t t

)]
_
. (A.17)
1350001-14
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
Note that, from (17), the second term on the right-hand side of (A.1) should
give contributions to the renormalized stress tensor. Inserting (11) in (A.1) and
using the Fourier representation (A.2) as well as the noise correlation (9), we have
(x)G
1
(x

, x

) =
2
1
a
1
+

n
1
=

1
a
d
+

n
d
=
_
d

(2)
e
i[k(n)(x

x)

(t

t)]

2
k
2
(n) + i

1
a
1
+

m
1
=

1
a
d
+

m
d
=
_
d
(2)
e
i[k(m)(xx

(tt

)]

2
k
2
(m) + i

2
.
(A.18)
We see that we get a sort of product between two unperturbed propagators. Per-
forming the integration over yields
(x)G
1
(x

, x

) =

2
a
2
1
+

n
1
,m
1
=

1
a
2
d
+

n
d
,m
d
=
k(m)
4k(n)

_
e
i[k(n)(x

x)k(n)(t

t)]
(t

t)
+ e
i[k(n)(x

x)k(n)(t

t)]
(t t

)
_

_
e
i[k(m)(xx

)k(m)(tt

)]
(t t

)
+ e
i[k(m)(xx

)k(m)(tt

)]
(t

t)
_
. (A.19)
To obtain the respective correction to G
(1)
(x, x

), one must employ the relation


(16) and notice that G
R,A
have similar perturbation expansions as G, Eq. (14). In
this way, using the appropriate contour for G
R,A
, one gets
(x)G
(1)
1
(x

, x

) = 2
_
(x)G
1
(x

, x

) +
1
2
_
(x)G
R
1
(x

, x

) + (x)G
A
1
(x

, x

)
_
_
,
(A.20)
where
(x)G
N
1
(x

, x

) =

2
a
2
1
+

n
1
,m
1
=

1
a
2
d
+

n
d
,m
d
=
k(m)
4k(n)

N
(t, t

, t

_
e
i[k(n)(x

x)k(n)(t

t)]
e
i[k(n)(x

x)k(n)(t

t)]
_

_
e
i[k(m)(xx

)k(m)(tt

)]
e
i[k(m)(xx

)k(m)(tt

)]
_
,
(A.21)
with N = R, A,
R
(t, t

, t

) = (t

t)(tt

) and
A
(t, t

, t

) =
R
(t, t

, t

).
1350001-15
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
V. A. De Lorenci, G. Menezes & N. F. Svaiter
References
1. S. Carlip, Rep. Prog. Phys. 64, 885 (2001).
2. A. Ashtekar, New. J. Phys. 7, 198 (2005).
3. L. H. Ford, Phys. Rev. D 51, 1692 (1995).
4. L. H. Ford and N. F. Svaiter, Phys. Rev. D 54, 2640 (1996).
5. L. H. Ford and N. F. Svaiter, Phys. Rev. D 56, 2226 (1997).
6. H. Yu and L. H. Ford, Phys. Rev. D 60, 084023 (1999).
7. R. T. Thompson and L. H. Ford, Phys. Rev. D 78, 024014 (2008).
8. R. T. Thompson and L. H. Ford, Class. Quantum Grav. 25, 154006 (2008).
9. H. Yu, N. F. Svaiter and L. H. Ford, Phys. Rev. D 80, 124019 (2009).
10. H. B. G. Casimir, Proc. Kon. Ned. Akad. Wekf. 51, 793 (1948).
11. G. Plunien, B. M uller and W. Greiner, Phys. Rep. 134, 87 (1986).
12. A. A. Grib, S. G. Mamayev and V. M. Mostepanenko, Vacuum Quantum Eects in
Strong Fields (Friedman Laboratory Publishing, St. Petesburg, 1994).
13. M. Bordag, U. Mohideen and V. M. Mostepanenko, Phys. Rep. 353, 1 (2001).
14. K. A. Milton, The Casimir Eect: Physical Manifestation of Zero-Point Energy
(World Scientic, Singapore, 2001).
15. K. A. Milton, J. Phys. A 37, R209 (2004).
16. E. Arias, G. Krein, G. Menezes and N. F. Svaiter, Int. J. Mod. Phys. A 27, 1250129
(2012).
17. C. H. G. Bessa, J. G. Due nas and N. F. Svaiter, Class. Quantum Grav. 29, 215011
(2012).
18. G. Krein, G. Menezes and N. F. Svaiter, Phys. Rev. Lett. 105, 131301 (2010).
19. A. Ishimaru, Wave Propagation and Scattering in Random Media (Academic, New
York, 1978).
20. P. Sheng, Scattering and Localization of Classical Waves in Random Media (World
Scientic, Singapore, 1990).
21. E. Arias, E. Goulart, G. Krein, G. Menezes and N. F. Svaiter, Phys. Rev. D 83,
125022 (2011).
22. A. Edery, J. Math. Phys. 44, 599 (2003).
23. L. Ford and N. F. Svaiter, Phys. Rev. D 80, 065034 (2009).
24. L. H. Ford and N. F. Svaiter, J. Phys. Conf. Ser. 161, 012034 (2009).
25. L. H. Brown and G. J. Maclay, Phys. Rev. 184, 1272 (1969).
26. C. M. Bender and P. Hays, Phys. Rev. D 14, 2622 (1976).
27. D. Deutsch and P. Candelas, Phys. Rev. D 20, 3063 (1979).
28. P. Candelas, Phys. Rev. D 21, 2185 (1980).
29. J. R. Ruggiero, A. H. Zimerman and A. Villani, Rev. Bras. Fis. 7, 663 (1977).
30. J. R. Ruggiero, A. Villani and A. H. Zimerman, J. Phys. A 13, 767 (1980).
31. S. K. Blau, M. Visser and A. Wipf, Nucl. Phys. B 310, 163 (1988).
32. N. F. Svaiter and B. F. Svaiter, J. Phys. A 25, 979 (1992).
33. B. F. Svaiter and N. F. Svaiter, Phys. Rev. D 47, 4581 (1993).
34. B. F. Svaiter and N. F. Svaiter, J. Math. Phys. 35, 1840 (1994).
35. S. W. Hawking, Commun. Math. Phys. 55, 133 (1977).
36. A. Voros, Commun. Math. Phys. 110, 439 (1987).
37. E. Elizalde, S. D. Odintsov, A. Romeo, A. A. Bytsenko and S. Zerbini, Zeta Regular-
ization Techniques and Applications (World Scientic, Singapore, 1994).
38. K. Kirsten, Spectral Functions in Mathematics and Physics (Chapman and Hall/CRC,
Florida, 2002).
39. S. Fulling, J. Phys. A 36, 6857 (2003).
40. L. H. Ford and N. F. Svaiter, Phys. Rev. D 58, 065007-1 (1998).
1350001-16
January 11, 2013 10:47 WSPC/Guidelines-IJMPA S0217751X13500012
Light-Cone Fluctuations
41. N. D. Birrell and P. C. Davis, Quantum Fields in Curved Space (Cambridge University
Press, New York, 1982).
42. M. J. Stephen, Phys. Rev. B 37, 1 (1988).
43. J. Ambjorn and S. Wolfram, Ann. Phys. (N.Y.) 147, 1 (1983).
44. F. Caruso, N. P. Neto, B. F. Svaiter and N. F. Svaiter, Phys. Rev. D 43, 1300 (1991).
45. R. D. M. De Paola, R. B. Rodrigues and N. F. Svaiter, Mod. Phys. Lett. A 34, 2353
(1999).
46. L. E. Oxman, N. F. Svaiter and R. L. P. G. Amaral, Phys. Rev. D 72, 125007 (2005).
47. V. Sopova and L. H. Ford, Phys. Rev. D 66, 045026 (2002).
1350001-17

Você também pode gostar