ijmpa2013

Upload: gabriel-menezes

Post on 14-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/30/2019 IJMPA2013

    1/17

    International Journal of Modern Physics A

    Vol. 28, No. 1 (2013) 1350001 (17 pages)c World Scientific 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

    [email protected]

    G. MENEZES

    Instituto de Fsica Teorica, Universidade Estadual Paulista,

    Rua Dr. Bento Teobaldo Ferraz 271 Bloco II,

    Sao Paulo, SP 01140-070, Brazil

    [email protected]

    N. F. SVAITER

    Centro Brasileiro de Pesquisas Fsicas,

    Rua Dr. Xavier Sigaud 150, Rio de Janeiro, RJ 22290-180, Brazil

    [email protected]

    Received 26 September 2012

    Revised 21 November 2012

    Accepted 7 December 2012Published 3 January 2013

    We investigate the effects of light-cone fluctuations over the renormalized vacuum expec-tation value of the stressenergy tensor of a real massless minimally coupled scalar field

    defined in a (d+ 1)-dimensional flat spacetime with topology RTd. For modeling theinfluence of light-cone fluctuations over the quantum field, we consider a random Klein

    Gordon equation. We study the case of centered Gaussian processes. After taking intoaccount all the realizations of the random processes, we present the correction caused

    by random fluctuations. The averaged renormalized vacuum expectation value of the

    stressenergy associated with the scalar field is presented.

    Keywords: Quantum fields 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

    http://dx.doi.org/10.1142/S0217751X13500012mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]://dx.doi.org/10.1142/S0217751X13500012
  • 7/30/2019 IJMPA2013

    2/17

    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 fixed 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 fixed for all times; rather, it is dynamical. Such a geometrical picture

    of the gravitational field is a very successful classical field theory. There are many

    attempts to bring general relativity into the quantum domain.1,2 Despite such enor-

    mous efforts, 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 field obeys quantum-mechanical laws is that the structure of spacetime

    must undergo quantum fluctuations. Ford and collaborators developed this idea,

    39

    showing that the effects of fluctuations of the geometry of the spacetime caused by

    quantum mechanical fluctuations of the gravitational field is to smear out the light

    cone. In general, in addition to quantum mechanical metric fluctuations, there are

    induced metric fluctuations generated by quantum fluctuations of matter fields. In

    both scenarios the concept of light-cone structure has to be modified.

    On the other hand, it is well known that when quantum fields are forced to

    obey classical boundary conditions, interesting effects arise which are connected

    with vacuum fluctuations of the field. Even if the spacetime is unbounded, if the

    quantum field is constrained by the presence of material boundaries, nontrivial con-

    sequences will show up due to vacuum fluctuations. In this respect, the most well

    known physical manifestation of such fluctuations is the Casimir effect, which has

    been extensively discussed in the literature.1015 In such a context, an intriguing

    question would be how light-cone fluctuations could affect measurable effects asso-

    ciated with virtual processes in quantum theory. Here we provide a scenario where

    this issue is investigated.

    In this paper, we consider the effects of light-cone fluctuations upon the renor-

    malized vacuum expectation value of the stressenergy tensor of a quantum field. In

    such a direction, recently it was investigated the influence of light-cone fluctuations

    over the transition probability rate of a two-level system coupled to a massless scalarfield undergoing uniformly accelerated motion.16 The assumptions made in such a

    reference were that the light-cone fluctuations can be treated classically and their

    effects on the quantum fields can be described via random differential 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 field in the presence of a

    disordered medium. For modeling the influence of light-cone fluctuations over the

    quantum field, we consider a random KleinGordon equation.18 We study the case

    of centered Gaussian processes. We consider fields defined in a nonsimply connectedspacetime with topology of Td R1, with Td corresponding to a hypertorus, i.e.

    a flat 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

  • 7/30/2019 IJMPA2013

    3/17

    Light-Cone Fluctuations

    massless scalar field in disordered media.1921 One could also regard this situation

    as a simplified model for the more realistic case of the Casimir energy due to

    phonons2224 in a random fluid confined between two infinite walls. In this case the

    influence of the light-cone fluctuations 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 fields in the presence of classical macroscopic boundaries and

    also in curved spacetime. We discuss the modifications in a free scalar quantum

    field theory due to the presence of randomness in Sec. 3. In Sec. 4, we derive

    the correction caused by random fluctuations in the renormalized vacuum expec-

    tation value of the stress tensor associated with quantum fields in the case men-

    tioned above. Conclusions are given in Sec. 5. In this paper, we use 32G = =

    kB = c = 1.

    2. Renormalized Vacuum Expectation Value of the Stress Tensor

    of Quantum Fields

    The problem of renormalization of ill defined quantities leading to a physically

    meaningful result is a fundamental question in quantum field theory. Although for-

    mally divergent, in the absence of gravity the difference between the vacuum energy

    of quantum fields at different physical configurations can be finite.10

    Being moreprecise, 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 field is regularized and then renormalized

    using auxiliary configurations which are added and subtracted. For example, the

    introduction of a pair of conducting plates into the vacuum of the electromagnetic

    field alters the zero-point fluctuations of the field and thereby produces an attrac-

    tion between the plates.1115 In this paper, we are interested to study the effects

    of light-cone fluctuations over the renormalized vacuum expectation value of thestress tensor.

    There are different ways to find the renormalized vacuum energy of quantum

    fields defined in a flat 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 field defined in flat

    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 fields and field derivatives. Here

    is the usual Minkowski metric (we take the sign convention of Ref. 41). For

    a quantum field propagating in a medium with disorder, V() will represent the

    coupling between the field and random impurities. In this paper, we assume that

    1350001-3

  • 7/30/2019 IJMPA2013

    4/17

    V. A. De Lorenci, G. Menezes & N. F. Svaiter

    such a random potential has the following functional form:

    V() = 12

    (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) = limx,xx

    T(x, x) , (3)

    where

    T(x, x) =

    1

    4

    2G

    (1)(x, x) G

    (1)(x, x)

    + (x)00G(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 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

    flat 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 isgiven 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 : T00(x) : .

    We remark that such a procedure cannot be trusted when the background space

    time is curved. In nongravitational physics, only energy differences 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 field theory in the presence of stochastic

    fluctuations of the light cone.

    1350001-4

  • 7/30/2019 IJMPA2013

    5/17

    Light-Cone Fluctuations

    3. Scalar Quantum Field Theory in the Presence of

    Random Fluctuations

    In this section, we present the solution to the scalar field equation in the presence of

    light-cone fluctuations. As pointed out, for modeling the influence of such fluctua-

    tions over the quantum field, it is enough to consider a stochastic KleinGordon

    equation. The field equation obtained in this way cannot be solved in a closed

    form. However, assuming that stochastic fluctuations 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 fluctuations. 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 field equation for a massless minimally coupled scalar

    field in a spacetime with stochastic light-cone fluctuations. It reads21[1 + (x)]

    2

    t2 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 fields behave in the presence of random fluctuations 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 fluctuations 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 define a preferred reference frame, similarly to

    an external heat bath, which induces a breaking in Lorentz symmetry.

    The field equation (7) should be compared with the ones used in Refs. 18 and21. 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 field equation cannot be

    given in a closed form, one can employ a perturbative series expansion for Greens

    1350001-5

  • 7/30/2019 IJMPA2013

    6/17

    V. A. De Lorenci, G. Menezes & N. F. Svaiter

    functions. The propagator iG(x, x) = T((x)(x)) associated with the wave

    equation (7) satisfies[1 + (x)]

    2

    t2 2x

    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 modifications introduced by the presence of the

    random term, we proceed as follows. Define the operator K = K0 L, where

    K0(x, y) =

    2/t2 2x i

    (x y) and

    L(x) = (x)

    2

    t2 . (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 defined:

    G = G0 G0LG0 + G0LG0LG0 + , (13)

    where we used a formal operator notation in this last equation and [K0]1 = G0

    is the unperturbed propagator. Following Ref. 21, Eq. (13) can be written in termsof space and time variables:

    G(x, x) = G0(x, x) +

    n=1

    dy1 G0(x, y1)G

    (n)(y1, x) , (14)

    where

    G(n)(y1, x) = (1)n

    nj=1

    L(yj)

    dyj+1 G0(yj , yj+1) . (15)

    In Eq. (15), it is to be understood that yn+1 = x

    and that there is no integra-tion in yn+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 (x1)(x2) (xp) can be easily expressed

    as the sum of products of two-point correlation functions corresponding to all pos-

    sible partitions of x1, x2, . . . , xp.

    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

    ) throughthe 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

  • 7/30/2019 IJMPA2013

    7/17

    Light-Cone Fluctuations

    can use the decomposition property stated above. On the other hand, G(1)(x, x)

    can also be computed through the relation41

    G(x, x) +1

    2

    GR(x, x) + GA(x, x)

    =

    i

    2G(1)(x, x) , (16)

    where GR(x, x) = i(t t)[(x), (x)] (GA(x, x) = i(t t)[(x), (x)])

    is the retarded (advanced) Greens function which obeys [K]1 = GR(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 theStress 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 T00(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)

    + 00 (x)G(1)(x, x)

    . (17)

    Now we focus our attentions on calculating the components of the stress tensor con-

    sidering the effects of light-cone random fluctuations. We consider the case where

    the fields satisfy periodic boundary conditions in all spatial directions. In this way,

    since the background spacetime is flat, 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 first 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:

    : T00(x): 0 =1

    2a1

    +n1=

    1

    ad

    +nd=

    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, wewrite this multiple sum in terms of the Epstein zeta function. Using (A.11) and

    defining

    f(d) =d+12

    2(d+1)/2

    , (19)

    1350001-7

  • 7/30/2019 IJMPA2013

    8/17

    V. A. De Lorenci, G. Menezes & N. F. Svaiter

    the renormalized vacuum energy density becomes

    : T00(x) : 0 = f(d)Z(a1, . . . , ad; d + 1) . (20)

    The result (20) is finite 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,

    : Tij(x): 0 = 14a1

    +n1=

    1ad

    +nd=

    1k(n)

    2kikj ijrskrks ijk2(n)

    , (21)

    where kj = 2nja1j . For i = j, one has

    : Tjj (x): 0 =1

    2a1

    +n1=

    1

    ad

    +nd=

    k2jk(n)

    =1

    2da1

    +

    n1=

    1

    ad

    +

    nd=

    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

    : Tjj (x) : 0 = f(d)

    dZ(a1, . . . , ad; d + 1) . (23)

    Now consider i = j. One has

    : Tij(x) : 0 =1

    2a1

    +n1=

    1

    ad

    +nd=

    kikjk(n)

    . (24)

    Since we have sums over the integers of a product between an even function and

    an odd function, we have that : Tij(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),

    : T0j(x) : 0 = 12a1

    +n1=

    1ad

    +nd=

    kj . (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

  • 7/30/2019 IJMPA2013

    9/17

    Light-Cone Fluctuations

    continuation for such series except for a pole at p = 2s.43 It is

    ()s(s)Zp(a1, . . . , ap; 2s)

    = 1

    s+

    2

    p 2s+ s

    dx xs1

    0, . . . , 0; a21x , . . . , a2px

    1

    + 2sp

    2

    1/

    dx x(p2s)/21

    0, . . . , 0;

    x

    a21, . . . ,

    x

    a2p

    1

    , (26)

    where p/2 is the product of the ps parameters ai given by p/2 = a1 ap, and

    the generalized Jacobi function (z1, . . . , zp; x1, . . . , xp) is defined by

    (z1, . . . , zp; x1, . . . , xp) =

    pi=1

    (zi; xi) , (27)

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

    (z; x) =

    n=

    e(2nzn2x) . (28)

    Using this integral expression for the Epstein zeta-function, given by (26), we can

    find that

    Zp(a1, . . . , ap; 2s)|s=0 = 1 , (29)

    for any p 1. So

    : T0j(x) : 0 =1

    2a1 ad

    +nj=

    kj . (30)

    Such a summation is zero, as the reader can easily check. So : T0j(x) : 0 = 0.

    Now let us introduce the corrections due to light-cone random fluctuations.

    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

    T00(x, x) yields

    : T00(x): 1 = P + E , (31)

    where

    P = 2

    8a1

    +

    n1= 1

    ad

    +

    nd=11

    a1

    +

    m1= 1

    ad

    +

    md=k2(m) (32)

    and

    E = 2Q(a1, . . . , ad; d)1

    2a1

    +n1=

    1

    ad

    +nd=

    k(n) . (33)

    1350001-9

  • 7/30/2019 IJMPA2013

    10/17

    V. A. De Lorenci, G. Menezes & N. F. Svaiter

    Consider the quantity P. The sums over n can be evaluated considering that

    +n1=

    +

    nd=

    1 = Zd(a1, . . . , ad; 2s)

    s=0

    .

    So, with the help of (29), we get

    P =2

    8a1

    +m1=

    1

    ad

    +md=

    k2(m) . (34)

    The sum over m can also be expressed in terms of the Epstein zeta-function

    +m1=

    +

    md=

    k2(m) = (2)2Zd(a1, . . . , ad; 2) ,

    which vanishes, in virtue of the functional reflection formula (A.11). So P = 0.

    Consider now E. Comparing Eqs. (18) and (33), one has

    E = 2[f(d)Z(a1, . . . , ad; d + 1)]2 , (35)

    where we have used Eqs. (19) and (A.14). Therefore, collecting our results one has

    : T00(x): 1 = 2[f(d)Z(a1, . . . , ad; d + 1)]

    2 . (36)

    This is the correction to the renormalized vacuum energy due to light-cone fluc-

    tuations up to second order in the noise. The subscript 1 on the left-hand side

    of the above equation indicates the first-order correction after performing the ran-

    dom averages. Now let us calculate the corrections to the components : Tij(x): .

    Inserting Eqs. (A.17) and (A.20) in (17), one has

    : Tij(x): 1 =2

    8Q(a1, . . . , ad; d)

    1

    a1

    +n1=

    1

    ad

    +nd=

    1

    k(n)

    2kikj ijrskrks 3ijk

    2(n)

    + ijP . (37)

    Considering the results derived above one has, for i = j, : Tij(x) : 1 = 0. For i = j,

    one has (no summation over the index j)

    : Tjj (x): 1 =

    2

    4 Q(a1, . . . , ad; d)

    1

    a1

    +n1=

    1

    ad

    +nd=

    k2j + k2(n)

    k(n)

    =2

    4

    (d + 1)

    dQ(a1, . . . , ad; d)

    1

    a1

    +n1=

    1

    ad

    +nd=

    k(n) , (38)

    1350001-10

  • 7/30/2019 IJMPA2013

    11/17

    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

    : Tjj (x) : 1 = 2

    2

    1 +

    1

    d

    [f(d)Z(a1, . . . , ad; d + 1)]

    2 . (39)

    With similar considerations as before, one may show that all corrections to the

    momentum density vanish, : T0j(x) : 1 = 0. Consequently, the final form for the

    renormalized expectation value of stress-tensor components in a nonsimply con-

    nected spacetime subjected to light-cone fluctuations reads, up to second order in

    the perturbations,

    : T00(x) : = g(a, d)[1 + 2g(a, d)] (40)

    and

    : Tjj (x) : = g(a, d)

    1

    d+

    2

    2

    1 +

    1

    d

    g(a, d)

    , (41)

    where a = (a1, a2, . . . , ad) and

    g(a, d) = f(d)Z(a1, . . . , ad; 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 field theory in the presence of light-cone

    fluctuations. After performing the random averages over the noise function, the

    correction caused by randomness in the renormalized stress tensor associated with

    the quantum field was presented. We obtained a correction which is proportional

    to the square of the unperturbed contribution. Although calculations show that

    contributions of light-cone fluctuations to the renormalized stress tensor are quite

    small, one cannot exclude the possibility of the existence of some mechanism whichis able to amplify such fluctuations giving rise to nontrivial measurable effects.

    Although the Casimir effect is a well understood phenomenon, there are still

    some open questions related to this effect. 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 electromagneticfield. Here we did not consider such problems.

    We stress that here we present all the technical details concerning how to

    evaluate the effects of light-cone fluctuations over the renormalized stress tensor

    associated with a quantum field. A further step is to use the approach developed

    1350001-11

  • 7/30/2019 IJMPA2013

    12/17

    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 fluid confined between

    plane boundaries. Phonons share several properties with relativistic quantum fields.

    Quantized acoustic perturbations in the presence of disorder and boundaries lead

    us to the phononic Casimir effect with randomness. Other intriguing situation is

    the case of a quantized electromagnetic field in the presence of boundaries. Ford

    and Sopova calculated the stress tensor for the quantized electromagnetic field in

    the region between a pair of dispersive, dielectric half-spaces.47 They showed that

    although the vacuum energy density is finite between perfectly reflecting plates, it

    will diverge near plates of finite reflectivity. It is interesting to investigate whether

    such behavior can also be obtained in the presence of light-cone fluctuations. 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 fluctuations. Let us present the

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

    G(x, x) = G0(x, x)

    dy G0(x, y)L(y)G0(y, x

    )

    + dy1 dy2 G0(x, y1)L(y1)G0(y1, y2)L(y2)G0(y2, x) . (A.1)We consider a topology of the background spacetime such that the fields must

    satisfy periodic boundary conditions in all spatial directions. For a general hyper-

    cuboidal space, with sides of finite length a1, . . . , ad, this corresponds to the com-

    pactification of the space dimensions to a hypertorus Td. The modes of the field

    then consist of a simple product of modes analogous to the usual Minkowski space.

    In this way, the unperturbed propagator reads

    G0(x, x) =

    1

    a1

    +

    n1=

    1

    ad

    +

    nd=

    d(2)

    ei[k(n)(xx)(tt

    )]

    2

    k2

    (n) + i

    , (A.2)

    with

    k(n) = 2N = 2

    n1a1

    ,n2a2

    , . . . ,ndad

    (A.3)

    1350001-12

  • 7/30/2019 IJMPA2013

    13/17

    Light-Cone Fluctuations

    and

    k2(n) = k2(n) = k21 + + k2d =

    2n1a12

    + +

    2ndad2

    , (A.4)

    where use has been made of the notation n2 = n21 + n22 + + n

    2d. 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

    G0(x, x) =

    i

    a1

    +n1=

    1

    ad

    +nd=

    1

    2k(n)

    ei[k(n)(xx)k(n)(tt)](t t)+ ei[k(n)(xx

    )k(n)(tt)](t t)

    . (A.5)

    Now let us introduce the corrections due to light-cone random fluctuations. In virtue

    of Eqs. (8) and (9), the first correction to G will be given by the third term on the

    right-hand side of (A.1). Then

    G2(x, x) =

    dy1 dy2 G0(x, y1)L(y1)G0(y1, y2)L(y2)G0(y2, 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

    G2(x, x) =1

    a1

    +n1=

    1

    ad

    +nd=

    d

    2ei[k(n)(xx

    )(tt)]

    1

    (2 k2(n) + i)()

    1

    (2 k2(n) + i), (A.7)

    where

    () = lim0

    221

    a1

    +

    n1=

    1

    ad

    +

    nd=

    d

    2

    2ei

    2 k2(n) + i. (A.8)

    Performing the integration as earlier we get, after taking 0,

    () = i22

    2

    1

    a1

    +n1=

    1

    ad

    +nd=

    k(n) . (A.9)

    Such a multiple sum may be written in terms of the Epstein zeta function:37

    Z(a11 , . . . , a1p ; s) =

    +

    n1,...,np=

    n1a1

    2+

    n2a2

    2+ +

    npap

    2s/2

    (A.10)

    for s > p and it should be understood that the term for which all ni = 0 is to be

    omitted. This function obeys the reflection formula43

    s

    2

    s/2Z(a1, . . . , ap; s) =

    (sp)/2

    a1 ap

    p s

    2

    Z(a11 , . . . , a

    1p ;p s) . (A.11)

    1350001-13

  • 7/30/2019 IJMPA2013

    14/17

    V. A. De Lorenci, G. Menezes & N. F. Svaiter

    Using Eqs. (A.10) and (A.11), we have

    () = i22Z(a1, . . . , ad; d + 1)d+12

    2(d+1)/2

    . (A.12)

    Then, after performing the integration,

    G2(x, x) =2

    2Q(a1, . . . , ad; d)

    1

    a1

    +n1=

    1

    ad

    +nd=

    1

    2k(n)

    i(t t)

    2 ei[k(n)(xx)k(n)(tt)](t t)

    +

    1

    2k(n)

    i(t t)

    2

    ei[k(n)(xx

    )k(n)(tt)](t t)

    , (A.13)

    where

    Q(a1, . . . , ad; d) = Z(a1, . . . , ad; d + 1)d+12

    2(d+1)/2

    . (A.14)

    Employing the decomposition property of the propagator in terms of the Wightman

    functions, one has

    iG0(x, x) + G2(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

    a1

    +n1=

    1

    ad

    +nd=

    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

    2Q(a1, . . . , ad; d)

    1

    a1

    +n1=

    1

    ad

    +nd=

    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

  • 7/30/2019 IJMPA2013

    15/17

    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)G1(x, x) = 2 1

    a1

    +n1=

    1

    ad

    +nd=

    d

    (2)

    ei[k(n)(xx)

    (tt)]

    2 k2(n) + i

    1

    a1

    +m1=

    1

    ad

    +md=

    d

    (2)

    ei[k(m)(xx)(tt

    )]

    2 k2(m) + i2 .

    (A.18)

    We see that we get a sort of product between two unperturbed propagators. Per-forming the integration over yields

    (x)G1(x, x) =2

    a21

    +n1,m1=

    1

    a2d

    +nd,md=

    k(m)

    4k(n)

    ei[k(n)(xx)k(n)(tt)](t t)

    + ei[k(n)(xx)k(n)(tt)](t t)

    ei[k(m)(xx)k(m)(tt)](t t)

    + ei[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 GR,A have similar perturbation expansions as G, Eq. (14). In

    this way, using the appropriate contour for GR,A, one gets

    (x)G

    (1)

    1 (x

    , x

    ) = 2

    (x)G1(x

    , x

    ) +

    1

    2

    (x)G

    R

    1 (x

    , x

    ) + (x)G

    A

    1 (x

    , x

    )

    ,

    (A.20)

    where

    (x)GN1 (x, x) =

    2

    a21

    +n1,m1=

    1

    a2d

    +nd,md=

    k(m)

    4k(n)N(t, t, t)

    ei[k(n)(x

    x)k(n)(tt)] ei[k(n)(x

    x)k(n)(tt)]

    ei[k(m)(xx)k(m)(tt)] ei[k(m)(xx

    )k(m)(tt)]

    ,

    (A.21)

    with N = R, A, R(t, t, t) = (tt)(tt) and A(t, t, t) = R(t, t, t).

    1350001-15

  • 7/30/2019 IJMPA2013

    16/17

    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. D51, 1692 (1995).4. L. H. Ford and N. F. Svaiter, Phys. Rev. D54, 2640 (1996).5. L. H. Ford and N. F. Svaiter, Phys. Rev. D56, 2226 (1997).6. H. Yu and L. H. Ford, Phys. Rev. D60, 084023 (1999).7. R. T. Thompson and L. H. Ford, Phys. Rev. D78, 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. D80, 124019 (2009).

    10. H. B. G. Casimir, Proc. Kon. Ned. Akad. Wekf. 51, 793 (1948).11. G. Plunien, B. Muller and W. Greiner, Phys. Rep. 134, 87 (1986).

    12. A. A. Grib, S. G. Mamayev and V. M. Mostepanenko, Vacuum Quantum Effects inStrong 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 Effect: Physical Manifestation of Zero-Point Energy

    (World Scientific, 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. Duenas 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, NewYork, 1978).

    20. P. Sheng, Scattering and Localization of Classical Waves in Random Media (WorldScientific, 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. D80, 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. D14, 2622 (1976).

    27. D. Deutsch and P. Candelas, Phys. Rev. D20, 3063 (1979).28. P. Candelas, Phys. Rev. D21, 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. B310, 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. D47, 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 Scientific, 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. D58, 065007-1 (1998).

    1350001-16

  • 7/30/2019 IJMPA2013

    17/17

    Light-Cone Fluctuations

    41. N. D. Birrell and P. C. Davis, Quantum Fields in Curved Space(Cambridge UniversityPress, New York, 1982).

    42. M. J. Stephen, Phys. Rev. B37, 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. D43, 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. D72, 125007 (2005).47. V. Sopova and L. H. Ford, Phys. Rev. D66, 045026 (2002).

    1350001-17