cálculo de las energías en el vacío

Upload: amalia-morales

Post on 03-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 Clculo de las energas en el vaco

    1/36

    arXiv:hep-th/0207120v21

    3Jul2002

    Calculating Vacuum Energies in Renormalizable Quantum FieldTheories: A New Approach to the Casimir Problem

    N. Grahama, R.L. Jaffeb, V. Khemanib, M. Quandtb, M. Scandurrab, H. Weigel1b,c

    aDepartment of Physics and Astronomy,University of California, Los Angeles

    Los Angeles, CA 90095

    bCenter for Theoretical Physics, Laboratory for Nuclear Scienceand Department of Physics, Massachusetts Institute of Technology

    Cambridge, Massachusetts 02139

    cInstitute for Theoretical Physics, Tubingen UniversityD-72076 Tubingen, Germany

    UCLA/02/TEP/14 MIT-CTP-3278 UNITU-HEP-11/2002hep-th/0207120

    AbstractThe Casimir problem is usually posed as the response of a fluctuating quantum field to externally

    imposed boundary conditions. In reality, however, no interaction is strong enough to enforce a

    boundary condition on all frequencies of a fluctuating field. We construct a more physical model of

    the situation by coupling the fluctuating field to a smooth background potential that implements the

    boundary condition in a certain limit. To study this problem, we develop general new methods to

    compute renormalized oneloop quantum energies and energy densities. We use analytic properties

    of scattering data to compute Greens functions in timeindependent background fields at imaginary

    momenta. Our calculational method is particularly useful for numerical studies of singular limits

    because it avoids terms that oscillate or require cancellation of exponentially growing and decaying

    factors. To renormalize, we identify potentially divergent contributions to the Casimir energy with

    low orders in the Born series to the Greens function. We subtract these contributions and add

    back the corresponding Feynman diagrams, which we combine with counterterms fixed by imposing

    standard renormalization conditions on loworder Greens functions. The resulting Casimir energy

    and energy density are finite functionals for smooth background potentials. In general, however,

    the Casimir energy diverges in the boundary condition limit. This divergence is real and reflects

    the infinite energy needed to constrain a fluctuating field on all energy scales; renormalizable

    quantum field theories have no place for ad hoc surface counterterms. We apply our methods tosimple examples to illustrate cases where these subtleties invalidate the conclusions of the boundary

    condition approach.

    Keywords: Energy densities, Greens functions, density of states, renormalization, Casimir effect

    PACS: 03.65.Nk, 03.70.+k, 11.10.Gh

    1Heisenberg Fellow

    1

    http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120v2http://arxiv.org/abs/hep-th/0207120http://arxiv.org/abs/hep-th/0207120http://arxiv.org/abs/hep-th/0207120v2
  • 8/12/2019 Clculo de las energas en el vaco

    2/36

    1. Introduction

    Long ago, Casimir predicted the existence of a force generated by quantum fluctuations ofthe electromagnetic field between uncharged, perfectly conducting metal plates in vacuo [1].Since then, many related calculations have been carried out (for reviews see [ 2,3,4, 5, 6]).

    The standard approach to problems of this type has been to consider the response of a freequantum field to timeindependent boundary conditions imposed on stationary surfaces.This calculation is an idealization of the more physical description of the quantum field asinteracting with a smooth external potential, which goes to zero away from the surfaces.This paper will provide the technical tools necessary for understanding precisely when thisidealization is justified, and when it is hazardous. We will demonstrate the use of these toolsin some simple models; the reader is also referred to [7] for a briefer, less technical, and morefocused discussion of a variety of Casimir problems.

    In the field theory picture, the dynamics of a fluctuating field in a background potentialis governed by the oneloop effective energy, which can be expressed either as an infinitesum of Feynman diagrams or as a Casimir sum over modes. Both expressions are formally

    divergent, so they can only be defined by introducing counterterms fixed with physical renor-malization conditions. A Dirichlet boundary condition emerges when a repulsive backgroundscalar field becomes strong and sharply peaked. We will see that the renormalized total en-ergy of the fluctuating field relative to its value in the absence of boundaries can diverge inthis limit. In such cases, the idealized Casimir energyper se does not exist. However, thesedivergences are associated with the boundaries and therefore do not affect the energy densityaway from the boundaries or the force between rigid surfaces. Nevertheless some geometricquantities, such as the Casimir stress [8] the force per unit area on the surface wherethe boundary conditions are applied are sensitive to these divergences. In such cases,there is no way to calculate meaningful finite results by considering boundary conditions.

    To demonstrate these results in detail, we require efficient and robust techniques forcomputing Casimir energies and energy densities, which we describe in Section 2. Thecalculation of Casimir energy densities is interesting in its own right, for example in thestudy of energy conditions in general relativity [2, 9, 10]. Our approach will phrase thecalculation in the language of conventional renormalizable quantum field theory, in a waythat is amenable to efficient numerical computation. We study the vacuum polarizationenergy and energy density of a fluctuating boson field coupled to a background (x),which is sharply peaked in the boundary condition limit. We introduce counterterms fixedby perturbative renormalization conditions, which define physical inputs to the theory suchas particle masses and coupling constants at particular values of the external momenta. Theresulting renormalization scheme is conventional, precise, and unambiguous. It can be related

    to any another conventional scheme by the analysis of loworder Feynman diagrams. We donot, however, see any relation between this scheme and analyses in which boundary surfacedependent counterterms are introduced in an ad hoc manner. It is a common procedure inthe literature to simply ignore divergent surface integrals and pole terms arising in the courseof the calculation [11,12]. Thefunction method, in particular, introduces counterterms ofany power of the geometrical parameter [13]. Such counterterms redefine classical parameters

    2

  • 8/12/2019 Clculo de las energas en el vaco

    3/36

    of the background, but have no justification in quantum field theory. As a result, theseapproaches do not appear to be reconcilable with ordinary renormalizable quantum fieldtheory.

    Our method is based on an extension of methods developed for the total energy [ 14,15],together with an extension of the analytic continuation methods introduced in Ref. [16].

    First, we relate the matrix element of the energy density operator to the scattering Greensfunction at coincident points,G(x, x, k). We regulate the largekbehavior ofGby subtractingthe first few terms in its Born expansion and adding back the contribution of the associatedloworder Feynman diagrams [14,15]. The Greens function is difficult to deal with at large(real) k when the background field approaches the boundary condition limit: it oscillateswith amplitude much larger than the net contribution to the energy density. To avoid suchoscillations we rotate k to the imaginary axis [16,17]. Before doing so, however, we rewritethe Greens function as a product of terms, each of which is bounded for k on the imaginaryaxis and each of which can be easily computed by integrating a Schrodingerlike equationsubject to simple boundary conditions. This formalism allows us to avoid exponentiallygrowing quantities anywhere in the calculation. As a result, we are able to write the Casimir

    energy density as an integral along the positive imaginary axis and perform the integraldirectly for imaginary k. Here we depart from Refs. [16,17], which formulate the problemas an integral along the imaginary axis but then use a dispersion relation to compute theintegrand in terms of scattering data for real k .

    The final ingredient we need for the calculation is the Feynman diagram contribution,which is where we implement the renormalization conditions. The energy density operatorT00generates insertions in Feynman diagrams of every order in the background field . Usinga functional approach allows us to keep careful track of orders in in this calculation, whichwe need to maintain consistency with the rest of the calculation.

    In Section 3 we show how to apply our methods to a simple example, the Casimir energy

    density for a boson field in one space dimension. We couple the fluctuating field to abackground (x) with an interaction of the form2(x). For(x) =(xa) + (x + a), weobtain Dirichlet boundary conditions at x = a in the limit where the coupling constant becomes infinite. We find that thetotalCasimir energy diverges in this limit. The divergenceis localized at the points where the background is nonzero. For all x=a, the energydensity remains finite and asapproaches infinity it smoothly approaches the result obtainedby imposing the boundary condition = 0 at x= a. To complete Section 3 we study thevacuum polarization energy in the presence of a smooth, strongly peaked background field.By taking to be the sum of Gauians centered ata, we can study the way that theenergy density behaves as (x) approaches the boundary condition limit. As expected, itapproaches its limiting form at every x

    =

    a, though the rate of approach depends on the

    proximity toa.In Section 4 we consider the case of a circle in two spatial dimensions. We consider

    Gauian backgrounds peaked around r = a and study the limit in which the Gauianapproaches a -function. When the strength of the coupling goes to infinity, we obtain aDirichlet boundary condition at r = a. However, the renormalized total energy divergesin the -function limit, even at finite coupling. This result is a simple consequence of the

    3

  • 8/12/2019 Clculo de las energas en el vaco

    4/36

    fact that

    d2x[(r)]2 in this limit. As a result, the surface tension on the circle the derivative of the energy with respect to a diverges. This example illustrates thegeneric problem we have found with attempts to define the Casimir energy in the context ofa quantum field theory: Even though the renormalized vacuum polarization energy is a finitefunctional of a smooth background, it diverges as the background field assumes the singular

    configuration necessary to implement the boundary condition. This divergence affects anyphysical quantity, such as the surface tension, whose measurement requires comparison ofconfigurations with different boundaries, for which the divergent contributions are different.We take this divergence as an indication that such quantities depend on the details of theinteraction between the fluctuating field and the material to which it couples. Thus theCasimir stress cannot be defined in a way that is independent of the other stresses to whichthe material is subject.

    We close with a summary and some considerations for future work in Section 5 and givetechnical details in the Appendices.

    2. Method

    We begin by developing the method for computing the energy density of a fluctuatingquantum field coupled to a classical background. From the corresponding scattering andbound state wavefunctions, we construct a Fock decomposition of the quantum field andcompute the matrix element of the energy density operator T00. We express this matrixelement in terms of a Greens function with appropriate boundary conditions. The energydensity is given by a sum over bound states plus an integral over the continuum labeled bythe wave numberk. We develop a representation of the Greens function suitable for analyticcontinuation into the upper halfkplane, so the k integral can be deformed along a cut onthe positive imaginary axis. To regulate the ultraviolet divergences of the theory, which cor-responds to eliminating the contribution associated with the semicircle at infinite complex

    momenta, we subtract leading Born approximations to the Greens function, which we lateradd back in as Feynman diagrams. These diagrams are then regularized and renormalizedin ordinary Feynman perturbation theory.

    2.1 Formalism

    We consider a static, spherically symmetric background potential = (r) with r= |x|innspatial dimensions. The symmetric energy density operator for a real scalar field coupledto is

    T00(x) =1

    22 + ( )

    2 +m22 +(r)2=

    1

    2

    2 +

    2 +m2 +(r)

    +1

    42

    2

    (1)

    where we have rearranged the spatial derivative term in order to be able to use the Schr o-dinger equation to evaluate the expression in brackets. We assume that the background

    4

  • 8/12/2019 Clculo de las energas en el vaco

    5/36

    potential is smooth, with0 dr|(r)|(1 + r)< .2 We define the vacuum to be the state

    | of lowest energy in the background and the trivial vacuum to be the state|0 oflowest energy when 0. The vacuum energy density is the renormalized expectation valueofT00 with respect to the vacuum|,|T00(x)|ren, which includes the matrix elementsof the counterterms. Since we have spherical symmetry, we will define the energy density in

    a spherical shell,(r) =

    2n/2

    ( n2 )rn1|T00(x)|ren. (2)

    We first perform a partial wave decomposition using spherical symmetry,

    (t, x) ={}

    (t, r)Y{}(x) (3)

    whereY{}(x) are the ndimensional spherical harmonics{} refers to the set of all angularquantum numbers in n dimensions. The total angular momentum assumes integer values= 0, 1, 2, . . .in all dimensions except for n= 1, where = 0 and 1 only, corresponding to

    the symmetric and antisymmetric channels respectively. We make the Fock decomposition

    (t, r) = 1

    rn12

    0

    dk

    (k, r) e

    ita(k) + (k, r) e

    ita(k)

    + 1

    rn12

    j

    12j

    j (r) e

    ijtaj+j (r) eijtaj

    , (4)

    which we have split into scattering states with =

    k2 +m2 and bound states with j =m2 2j . Note that the latter wave functions are real. The vacuum | is annihilated by all

    of thea(k) andaj. The radial wavefunctions in eq. (4) are solutions to the Schrodingerlikeequation

    + 1

    r2

    1

    2

    +1

    2

    +(r) k2

    = 0 (5)

    where= 1 + n

    2. (6)

    In each channel, we normalize the wavefunctions to give the completeness relation

    2

    0

    dk (k, r)(k, r) +

    j

    j(r)j(r) =(r r) (7)

    which implies the orthonormality relations

    0 j (r)j

    (r) dr = jj

    for the bound states and0

    (k, r)(k, r) dr =

    2(k k) for the continuum states.

    (8)

    2In view of later applications with singular background fields, we can relax this condition to allow forall (r) for which a scattering problem in the usual sense can be defined. However, some of the standardbounds on the asymptotics of Jost functions no longer hold for a nonsmooth (r), giving rise to theadditional singularities discussed later.

    5

  • 8/12/2019 Clculo de las energas en el vaco

    6/36

    For 0, the normalized wavefunctions are (0) (k, r) =

    2krJ(kr). With the above con-

    ventions and normalizations, the standard equal time commutation relations for the quantumfield yield canonical commutation relations for the creation and annihilation operators,[a(k), a

    (k

    )] =(k k), and [aj , aj] =jj while all other commutators vanish.The vacuum expectation value of the energy density can now be computed by using

    eq. (4) in eq. (1). We obtain

    (r) =

    N

    0

    dk

    (k, r)(k, r) +

    j

    j2

    j(r)2

    +1

    4Dr

    N

    0

    dk

    (k, r)(k, r) +

    j

    1

    2jj(r)

    2

    (0)(r) +CT(r) . (9)

    where N = (n+2)(n1)(+1)

    (n+ 2 2) is the degeneracy factor in the th partial wave, Dr =

    r

    r

    n1r

    , CT(r) is the counterterm contribution, and

    (0)(r) indicates the subtractionof the energy in the trivial vacuum, which is just given by evaluating the first integralwith free wavefunctions (0) (k, r). This subtraction corresponds to a renormalization of thecosmological constant.

    We can identify the scattering state contribution with the Greens function by definingthe local spectral density

    (k, r) kiG(r,r,k) , (10)

    in the upper halfplane, so that for real k

    Re {(k, r)} = (k, r)(k, r) = Im{k G(r,r,k)} , (11)

    whereG(r, r

    , k) = 2

    0dq

    (q, r)(q, r)(k+i)2 q2

    j

    j(r)j (r)k2 +2j

    . (12)

    The i prescription has been chosen so that this Greens function is meromorphic in theupper halfplane, with simple poles at the imaginary momenta k = ij corresponding tobound states. Thus we have

    (r) =

    N

    dk

    2i

    1 + 1

    42Dr

    kG(r,r,k)

    +

    N j j

    1 + 1

    42jDr

    j (r)

    2 (0)(r) +CT(r) , (13)

    where we have made use of the fact that for real k, the imaginary part of the Greensfunction at r = r is an odd function of k, while the real part is even. We would like touse eq. (13) to compute (r) as a contour integral in the upper halfplane. First we musteliminate the contribution from the semicircular contour at large |k| with Im(k) 0. Usingthe techniques of Refs. [18] and [19] we see that subtracting sufficiently many terms in the

    6

  • 8/12/2019 Clculo de las energas en el vaco

    7/36

  • 8/12/2019 Clculo de las energas en el vaco

    8/36

    (r) +N

    i=1

    (i)FD(r) +CT(r) . (16)

    Here (i)FD(r) is the Feynman diagram contribution at order i in the background field. As

    emphasized in the Introduction, the representation of(r) as an integral along the imaginary

    axis is only useful if it can be computed efficiently. In the following subsections we showhow to do so.

    2.2 Computational Techniques

    In this subsection we concentrate on (r), the continuum integral in eq. (16). We developa representation of the partial wave Greens function,G(r, r, k) and its Born series, eq. (14),that can be computed efficiently for imaginary k. By working with pure imaginary momentawe avoid functions that would oscillate at large |k|. We will make extensive use of theanalytic properties of scattering data [21].

    We begin by introducing solutions to eq. (5) that obey a variety of different boundary

    conditions. First we define the free Jost solution

    w(kr) = (1)

    2kr [J(kr) +iY(kr)] , (17)

    where= 1 + n2 . The radial functionw(kr) is a solution to eq. (5) with 0 describingan outgoing spherical wave. Then we define

    TheJost solution,f(k, r). It behaves like an outgoing wave at r , with

    limr

    f(k, r)

    w(kr) = 1 . (18)

    The Jost function, F(k). It is obtained as the ratio of the interacting and free Jostsolutions at r= 0,

    F(k) = limr0

    f(k, r)

    w(kr) . (19)

    The regular solution,(k, r). It is defined by a kindependent boundary condition atthe origin,

    limr0

    (+ 1)

    r

    2

    (+ 12)

    (k, r) = 1 . (20)

    The physical scattering solution, (k, r). It is also regular at the origin (and thus

    proportional to(k, r)), but differently normalized3

    (k, r) = k+

    12

    F(k)(k, r) . (21)

    3Whenever a fractional power ofk appears, we define it to be the limit as k approaches the real axis fromabove.

    8

  • 8/12/2019 Clculo de las energas en el vaco

    9/36

    The reason for distinguishing two regular solutions is that has a simple boundarycondition at r = 0, while has a physical boundary condition at r , corresponding toincoming and outgoing spherical waves. With these definitions, the Greens function has asimple representation

    G(r, r, k) =(k, r)

    F(k) (k)

    1

    2 , (22)

    where r> (r

  • 8/12/2019 Clculo de las energas en el vaco

    10/36

    where a prime indicates a derivative with respect to the radial coordinate r. Using theseboundary conditions, we integrate this differential equation numerically forg(it,r), startingat r = and proceeding to r= 0. Similarly, h(it,r) obeys

    h (it,r) =

    2t(tr)h

    (it,r) + (r)

    2t2 d()

    d

    =trh(it,r) (27)

    with the boundary conditions

    h(it, 0) = 0 and h(it, 0) = 1 (28)

    and we integrate numerically from r = 0 to r = . For real,

    () dd

    ln [w(i)] (29)

    is real with lim () = 1, so the two functions h(it,r) andg(it,r) are manifestly real.

    They are also holomorphic in the upper kplane and, most importantly, they are boundedaccording to

    |g(k, r)| const|h(k, r)| const 2r

    1 + |k|r (30)

    so that neither g nor h grows exponentially during the numerical integration. Thus therepresentation of the partial wave Greens function in terms of g and h is smooth andnumerically tractable on the positive imaginary axis. We have avoided oscillating functions orexponentially growing functions that eventually would cancel against exponentially decaying

    functions.Finally, the computation of the Born series, eq. (14), is also straightforward in thisformalism. We expand the solutions to the differential equations eq. (25) and eq. (27) aboutthe free solutions,

    g(it,r) = 1 +g(1) (it,r) +g

    (2) (it,r) +. . .

    h(it,r) = 2rI(tr)K(tr) +h(1) (it,r) +h

    (2) (it,r) +. . . , (31)

    where the superscript labels the order of the background potential . The higher ordercomponents obey inhomogeneous linear differential equations with the boundary conditions

    limr g

    (j)

    (it,r) = 0 and limr g

    (j)

    (it,r) = 0,h(j) (it, 0) = 0 and h

    (j) (it, 0) = 0 . (32)

    In these equations is the source term for g (1), g (1) is the source term for g(2), and so on.We then obtain the Born series for the local spectral density by substituting these solutionsin the expansion of eq. (24) with respect to the order of the background potential. Thus

    10

  • 8/12/2019 Clculo de las energas en el vaco

    11/36

    we have developed a computationally robust representation for the Born subtracted localspectral density,

    [(it,r)]N=

    t

    h(it,r)g(it,r)

    2g(it, 0)

    N

    . (33)

    2.3 Feynman Diagram Contribution

    In this subsection we develop methods to compute the contribution from the Feynman

    diagrams and counterterms,N

    i=0(i)FD(r) + CT(r), in eq. (16). To oneloop order, the Feynman

    diagrams of interest are generated by the expansion of

    0|T00(x)|0 i2

    Tr

    Tx2 m2

    1 (34)

    to order N in the background. Here Tx is the coordinate space operator corresponding tothe insertion of the energy density defined by eq. (1) at the spacetime point x, and the traceincludes spacetime integration. Since we are doing ordinary perturbation theory, we takethe matrix elements with respect to the trivial vacuum, which is annihilated by the planewave annihilation operators. The energy density operator has pieces of order 0 and 1,

    T00 =1

    2

    2 +

    2+m22 +2

    =

    1

    2

    ddy(y)

    T(0)x + T

    (1)x

    (y) (35)

    where d= n + 1 is the number of spacetime dimensions. We have separated the differentorders in the external background field,

    T(0)x =

    t (x

    y)

    t+

    (x

    y)

    +m2(x

    y)

    T(1)x =(x)(x y) . (36)

    The-functions are understood asddimensional. The arrows denote the direction in whichthe derivatives act when substituting this representation into the functional trace in eq. ( 34)and y is the coordinate space position operator, with y| = |. However, it is moreconvenient to perform the computation in momentum space. The relevant matrix elementsare

    k|T(0)x |k =ei(kk)x

    k0k0 + k k+m2

    k|T(1)x |k =(x)ei(k

    k)x . (37)

    Here we will explicitly consider the contributions toT00that are linear in because theseare the only contributions required in the examples studied in Sections 3 and 4. The firstpiece comes directly from T(1)x as shown in the left panel of Fig.2,

    i

    2Tr

    1

    2 m2 T(1)

    x

    =

    i

    2(x)

    ddk

    (2)d1

    k2 m2. (38)

    11

  • 8/12/2019 Clculo de las energas en el vaco

    12/36

    Tx^ (0)

    T^ (1)x

    Figure 2: First order Feynman diagrams contributing to the energy densityT00(x). The solidcircle represents the loop and the operator insertions from eq. (36) are denoted by T

    (0,1)

    x . Inthe right panel, is the local insertion of the background field from expanding the propagator,cf. eq. (39).

    This local contribution is ultraviolet divergent for d 2. It is canceled identically by thecounterterm in the notadpole renormalization scheme, since the righthand side of eq. (38)equals minus the counterterm Lagrangian evaluated for the background field . For furtherdiscussion of this renormalization condition and its consequences, see Ref. [15]. The secondcontribution at order originates from T(0)x and the firstorder expansion of the propagatoras shown in Fig.2,

    i

    2 Tr 12 m2 T

    (0)

    x

    1

    2 m2 . (39)For this diagram we find

    i

    2

    ddk

    (2)d

    ddk

    (2)d1

    k2 m2 ei(kk)x

    k k +m2

    1k2 m2 (k k

    )

    = i

    2

    ddq

    (2)d(q)eiqx

    10

    d

    ddl

    (2)dl l+m2 (1 )q q[l2 m2 +(1 )q2]2 , (40)

    where (q) = 2(q0)(q) is the Fourier transform of the (time independent) backgroundfield. We have defined p= (p0, p) for the Lorentz-vectorp = (p0, p). The leading ultravioletdivergences cancel in eq. (40) between the temporal and spatial components. We thus obtain

    the expression for the Feynman diagram and counterterm contribution through first orderin,

    (1)FD(r) +CT(r) =

    2n2 rn1

    ( n2 )

    i

    2Tr

    1

    2 m2 T(0)

    x

    1

    2 m2

    =2

    n2 rn1

    ( n2 )

    2 d2

    (4)d/2

    dnq

    (2)n(q)eiqx

    10

    d (1 )q2

    m2 +(1 )q22d/2.(41)

    We note that this piece does not contribute to the total energy because it vanishes whenintegrated over space. The extension to higher order Feynman diagrams in is straightfor-ward. For example, the contribution

    O(2) is obtained from the two terms

    i

    2Tr

    T(1)x

    1

    2 +m2

    1

    2 +m2

    i

    4Tr

    T(0)x

    1

    2 +m2

    1

    2 +m2

    1

    2 +m2

    . (42)

    We use the matrix elements, eq. (37), to write this in terms of ordinary loop integrals

    i

    2(x)

    ddq

    (2)d(q)eiqx

    10

    d ddl

    (2)d1

    [l2 m2 +(1 )q2]2

    12

  • 8/12/2019 Clculo de las energas en el vaco

    13/36

    +i

    2

    ddq

    (2)d

    ddp

    (2)d(qp)(p)eiqx

    10

    d 10

    d

    ddl

    (2)dl l+ [q+ p] [(1 )q p]

    [l2 m2 +(1 )q2 +(1 )p2]3. (43)

    As a consequence of eq. (37), the term involving T(1)

    x will always carry an explicit factor of(x).

    2.4 Total Energy

    We can obtain the total energy simply by integrating the energy density in eq. (16),

    E[] =0

    (r)dr . (44)

    Since both the t integral and the sum over channels in eq. (16) are absolutely convergent,we can interchange the order of integration and obtain

    E[] =

    N

    m

    dt

    t2 m2

    0

    dr [(it,r)]N+N

    i=1

    E(i)FD+ECT, (45)

    where the total derivative term has integrated to zero. As explained in the previous subsec-tion, we have to subtract a sufficient number of Born approximations to the local spectraldensity(it,r) to render the t integral convergent. These subtractions are then added backin as the contribution to the total energy from the Feynman diagrams

    N

    i=1 E(i)FD=

    N

    i=1

    0dr

    (i)FD(r) , (46)

    which we then combine with the contribution from the counterterms

    ECT =0

    dr CT(r) (47)

    to obtain a finite result. In practice, these terms can be more efficiently computed directlyfrom the perturbation series of the total energy. For the first term in eq. (45), we employthe relation

    20

    dr [(k, r)]0=2k

    i

    0

    drG(r,r,k) G(0) (r,r,k)

    = i

    d

    dkln F(k) (48)

    which we prove in Appendix A for all k in the upper halfplane, Imk >0. Using eq. (48) onthe imaginary axisk = it, we find in particular

    20

    dr [(it,r)]0= d

    dtln F(it) =

    d

    dtln lim

    r0

    f(it,r)

    w(itr) =

    d

    dtln g(it, 0) . (49)

    13

  • 8/12/2019 Clculo de las energas en el vaco

    14/36

  • 8/12/2019 Clculo de las energas en el vaco

    15/36

    We can also see this relation directly by integrating eq. (9) over r, with NBorn termssubtracted and added back in as diagrams,

    E[] =

    N

    +i0+i

    dk

    k2 +m2

    0

    dr [ (k, r)(k, r)]N+

    j

    j2

    +

    Ni=1

    E(i)FD+ECT ,

    (57)where the total derivative term has integrated to zero and the bound state wavefunctions haveintegrated to unity. We can then use eq. (A.14) to see that eq. (57) is equivalent to eq. (56).While this relation is useful for building an intuitive understanding of the expressions for thetotal energy and the energy density, eq. (57) is of no practical use in numerical calculations;it only converges because eq. (54) has introduced the standard i prescription to control theoscillations of the wavefunctions at spatial infinity. Any direct numerical calculation usingeq. (57) will be hopelessly obscured by these oscillations.

    3. Examples in One Space Dimension

    To illustrate our method, we apply it first to simple Casimir problems in one spacedimension. There are only two channels (symmetric and antisymmetric) in this case. Theboundary condition in the symmetric channel, d

    dx|x=0 = 0, requires slight modifications of

    the formalism developed so far. We study two examples: a background field consisting of-functions, which allows us to make contact with treatments already in the literature [3,4],and a Gauian background, to demonstrate the efficient use of our methods for numericalcomputation.

    3.1 Greens Function

    In one spatial dimension, the Smatrix for a symmetric potential is diagonalized in a

    basis of symmetric and antisymmetric wavefunctions. In the antisymmetric channel, we canproceed with the general techniques described in Section 2. From eqs. (25) and (27) wecompute g(it,x) and h(it,x) with = 1. Note thatg(it,x) is the same in both channelsbecause the channels only differ by the boundary conditions at the origin, which do notaffectg . At equal spatial arguments we have

    G(x,x,it) =h(it,x)g(it,x)

    g(it, 0) . (58)

    However, the boundary conditions in the symmetric channel,

    +(0) = 1 and +(0) = 0 , (59)

    are not in the class defined by eq. (20). Constructing the Jost solution f(k, x) with theboundary condition limx f(k, x)e

    ikx = 1 and working out the Wronskian for + and fyields the Greens function in the symmetric channel

    G+(x,y,k) = +(k, x)f(k, 0)

    , (60)

    15

  • 8/12/2019 Clculo de las energas en el vaco

    16/36

    where f(k, 0) = xf(k, x)|x=0. By symmetry we have G+(x, y, k) = G+(x, y, k). Wefactor out the asymptotic behavior,

    +(k, x) =eikxh+(k, x) (61)

    and obtain h+(it,x) 2th+(it,x) +(x)h+(it,x) = 0 . (62)

    which is the same differential equation obeyed by h(it,x). The boundary conditions ineq. (59) then become

    h+(it, 0) = 1 and h+(it, 0) = t . (63)

    With this solution, we can now construct the Greens function in the symmetric channel. Tocompute densities it again suffices to consider x= y, where we have

    G+(x,x,it) = g(it,x)h+(it,x)

    tg(it, 0) g

    (it, 0)

    (64)

    and the complete Greens function is the sum G= G++G, giving

    (it,x) =tg(it,x)h(it,x)

    g(it, 0) +

    g(it,x)h+(it,x)

    g(it, 0) g(it, 0)/t. (65)

    The Born series is then obtained by iteration of the differential equations ( 25), (27), and(62).

    3.2 Total Energy

    The contribution of the symmetric channel to the total energy cannot be cast in thegeneral form of eq. (50) because, as for the Greens function, the boundary condition for thecorresponding wavefunction is not of the type of eq. (20). Since the derivative of+vanishesat x = 0, the Jost function becomes

    F+(k) = limx0

    f(k, x)

    ik =g(k, 0) +

    g (k, 0)

    ik (66)

    which is normalized so that F+(k) = 1 when 0. Otherwise, the same analyticityarguments apply as in channels whose wavefunctions vanish at r = 0. In particular, the kintegration contour can be closed around the discontinuity along the positive imaginary axis.

    We obtain

    E[] =

    m

    dt

    2

    tt2 m2

    ln g(it, 0) + ln

    g(it, 0) 1

    tg(it, 0)

    1

    +E(1)FD+ECT, (67)

    where as before the subscript indicates that the first Born term has been removed fromthe expression in brackets. One subtraction suffices in one dimension. By our notadpole

    16

  • 8/12/2019 Clculo de las energas en el vaco

    17/36

    renormalization condition, the counterterm cancels the tadpole diagram identically, so thatE

    (1)FD+ECT 0.

    3.3 DeltaFunction Background

    We first consider the case of two -functions separated by a distance 2a,

    (x) = [(x+a) +(x a)]. (68)

    At each -function, the derivative of the field jumps by . For modes withk , thewavefunction (k, x) vanishes ata, so that the limit gives a Dirichlet boundarycondition ata. Since the counterterms are local functionals of the background field, theyare identically zero in regions where the background fields vanish. Hence in a renormalizabletheory, local observables such as the energy density cannot have ultraviolet divergences inthese regions, as we will see explicitly below.

    By integration of the respective differential equations we obtain

    g(it,x) =

    1 |x| > a1 +

    2t

    1 e2t(xa)

    0 |x| < a

    h+(it,x) =1

    2

    1 +e

    2tx +

    2t

    1 +e2ta e2tx e2t(xa)

    |x| > a

    1 +e2tx 0 |x| < a(69)

    h(it,x) = 1

    2t

    1 e

    2tx +

    2t

    1 e2ta +e2tx e2t(xa)

    |x| > a

    1 e2tx 0 |x| < a

    It is clear that g, h+, and h are all wellbehaved for t >0. We substitute g(it, 0) and itsderivative into eq. (67) and find the total energy

    E2(a, ) =

    m

    dt

    2

    1t2 m2

    t ln

    1 +

    t +

    2

    4t2

    1 e4ta

    . (70)

    The last term in curly brackets is the N = 1 Born subtraction. Carrying out a similarcalculation for a single -function, we obtain

    E1() =

    m

    dt

    2

    t ln1 +

    2t

    2t2 m2 (71)

    = m

    4 +

    m

    2

    1 24m2arccos 2m 2m 2

    4m2 1 ln

    2m

    2

    4m2 1

    >2m .

    (72)

    From eq. (72) we can verify thatE1 andE2 obey the consistency conditions lima

    E2(a, ) =

    2E1() and lima0

    E2(a, ) = E1(2).

    17

  • 8/12/2019 Clculo de las energas en el vaco

    18/36

    We observe that in the limit of large the total energy associated with a single -functionapproaches minus infinity as ln , which cannot be canceled by any available counterterm.Thus in this limit the energy is infinite. In this sense the DirichletCasimir problem is notwell defined in the context of renormalizable quantum field theory. However, we can computethe force between two -functions for finite and subsequently take the limit,

    K(a) = lim

    E2(a, )

    (2a) =

    m

    dt

    t2t2 m2 (e4ta 1) . (73)

    which is equal to the result found using boundary conditions [3]. The massless limit does notexist for the vacuum polarization energy in eq. (72), even for finite. This is to be expectedsince massless scalar field theories are infrared divergent in 1 + 1 dimensions. However, itexists for the force, yielding the wellknown result

    limm0

    K(a) = 96a2

    . (74)

    Next we turn to the energy density. In this Section, we consider the energy density(x) = T00 without factoring out the surface area factor as we did in the previous Section(which would just give a factor of 2 here). The resulting energy density (x) is normalizedsuch that

    (x)dxgives the total energy. It is convenient to decompose the energy density

    into (x), the integral over the subtracted local spectral density, and (1)FD(x) + CT(x), the

    contribution of the Feynman diagram plus counterterms.First we study (x). From the radial functions in eq. (69), we find the local spectral

    density as a function ofx andk = it,

    [(it,x)]0 =

    2

    (2t+)2

    e4ta

    2t + (2t+)e4ta

    e2t(a|x|) |x| > a

    2(2t+)e2ta cosh(2tx) 0 |x| < a ,

    (75)

    where only the free local spectral density has been subtracted. For |x| =a the local spectraldensity decays exponentially as t increases, so it yields a finite energy density without anysubtractions, as we expected since the counterterm vanishes away from x = a.

    To prepare for more general problems involving smooth potentials, where the energy willdiverge at allx, we carry out the full Born subtraction procedure even though in this case theBorn subtraction and the diagram contribution are both finite and equal for all |x| =a, andso their contributions cancel. We will thus obtain the energy density everywhere includingthe pointsx = a. The first Born approximation is given by

    (1)(it,x) = 2t

    e2t(|x|a) +e2t(|x|+a), |x| > a

    e2t(|x|a) +e2t(|x|+a), 0 |x| < a(76)

    and we obtain

    (x,a,) =

    m

    dt

    (it,x)t2 m2 (77)

    18

  • 8/12/2019 Clculo de las energas en el vaco

    19/36

    where

    (it,x) =2

    2

    2t + (2t+)e4ta2 (2t+)2e4ta (|x| a)

    + m22/t

    (2t+)2

    e4ta

    2

    12[6t e4ta + (2t+)e4ta] e2t(a|x|) |x| > a

    2tm2 (m2 t2) + [(2t+)e2ta e2ta] cosh(2tx) 0 |x| < a .

    (78)

    For the Feynman diagram and the counterterm, we use eq. (41) and split off a-functioncontribution,

    (1)FD(x,a,) +CT(x,a,) =

    2(|x| a)

    2

    0

    dqcos(qa) cos(qx) 10

    d m2

    m2 +(1 )q2. (79)

    We recast the second term in a form which makes its identification with the first Borncontribution, eq. (76), manifest,

    (1)FD(x,a,) +CT(x,a,) =

    2(|x| a) 4m

    2

    2

    0

    dqcos(qa)cos(qx)

    q

    q2 + 4m2 sinh1

    q

    2m

    =

    2(|x| a) m

    2

    2

    m

    dt

    t

    t2 m2e2|xa|t +e2|x+a|t

    , (80)

    which confirms that our Feynman diagram representation of theO() contribution to(x,a,) agrees with the k integral representation for values of x where the countertermsvanish.

    In Fig.3we display the energy density for two delta functions separated by 2a,2(x,a,)as functions ofx. Note that there are-function contributions to 2(x,a,) at x=

    a not

    drawn in the figure. For reference we also show the sum of the energy densities for deltafunctions at x = a and x =a separately, which we denote 1(x a, ). The differencebetween2(x,a,) and the sum1(x a, ) +1(x+a, ) will result in a force between thetwo points.

    As , we expect to obtain the same energy density as if we had imposed theboundary conditions (a) = (a) = 0 from the outset. No subtraction is necessary awayfrom x=a, so we can study the energy density in this region directly from []0, given ineq. (75),

    lim

    2(x) = 1

    m

    dtt2

    m2

    t2 m2 +m2e2ta cosh(2tx)e4at

    1

    for |x| < a . (81)

    To compare with Ref. [3], we separate the xdependent part and transform the t integralinto a sum over the poles of the function e2ta/(e4at 1) = 1/(2sinh2ta) at t = in/2a,

    lim

    2(x) = 1

    m

    dtt2 m2

    t2 m2e4at 1

    m

    8am

    2

    4a

    n=1

    cos

    na (|x| a)

    n2a

    2+m2

    for |x| < a . (82)

    19

  • 8/12/2019 Clculo de las energas en el vaco

    20/36

    -0.4 -0.2 0.0 0.2 0.4

    xm

    -0.50

    -0.25

    0.00

    0.25

    0.50

    (x)/m2

    2

    FD

    -0.4 -0.2 0.0 0.2 0.4

    x m

    -0.30

    -0.20

    -0.10

    0.00

    (x)/m2

    2

    1(a)+

    1(-a)

    Figure 3: Left panel: The total energy density 2(x,a,) and the contributions (x,a,) and

    (1)FD(x,a,), defined in eqs. (77) and (80). Note that the displayed energy densities are supplemented

    by-function contributions located atx = a and x = a. Right panel: The total density2(x,a,)compared to the sum1(x a, ) +1(x+a, ) of the energy densities of two isolated -functions.The parameters are = 3m and a= 0.2/m.

    This expression agrees with the Dirichlet result given in Ref. [3].This simple example allows us to study the nature of the divergence in the Dirichlet

    Casimir energy in the limit . For any fixed x away from the -functions, the energydensity reaches a finite limit as , but the limit is approached nonuniformly. Thiscan be seen clearly in Fig.4, where we display the energy density for large couplings andcompare it to the limiting function, eq. (81). The closerx approaches to a, the slower theenergy density converges to its limiting form. Thus the energy density away from the sourceon the boundary remains finite and well defined in the boundary condition limit even thoughthe total energy diverges.

    3.4 Gauian Background Field

    In this subsection we demonstrate how to calculate the energy density associated witha Gauian background using our method. This problem is a warmup for the next Section,where we will use a tall, narrow Gauian background as an approximation to the -function inhigher dimensions, where the energy of a delta function diverges even at finite coupling. Wechoose a Gauian background that approximates the -function potential from the previoussubsection,

    G(x) =

    A

    2

    e (

    xa)2

    2w2

    +e (

    x+a)2

    2w2 , (83)

    in the limit w 0 whereA is chosen so that the area under the Gauian is fixed to 2. Wesubstitute this background field into the differential equations (25), (27) and (62) and obtainthe wavefunctionsg,handh+ numerically. As promised, these functions are wellbehavedon the positive imaginary k axis. At large momentat, we expect h(it,x) to approach thefree solution, which decays asymptotically as (1 +et|x|)/2t. This behavior can be seen

    20

  • 8/12/2019 Clculo de las energas en el vaco

    21/36

    100.0 1000.0 10000.0

    /m

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    (x,

    )/(x,

    8)

    = 100

    = 101

    = 102

    = 104

    Figure 4: Ratio of the energy density for two -functions with finite to the energy density for with a = 0.2/m, plotted as a function of for several values of = (a x)/a.

    0 10 20 30 40 50

    t / m

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    g(it,x)/g(it,

    0)

    x = 0.2/m

    x = 0.3/m

    x = 0.4/m

    0 5 10 15 20 25

    t / m

    0

    0.1

    0.2

    0.3

    0.4

    h

    (it,x

    )

    x=0.2/m

    x=0.3/m

    x=0.4/m

    Figure 5: Plots of g(it,x)/g(it, 0) and h(it,x) as functions of the imaginary momentum t atseveral positions x. The parameters are a= 0.3/m, w= 0.03/m and = 3m.

    21

  • 8/12/2019 Clculo de las energas en el vaco

    22/36

  • 8/12/2019 Clculo de las energas en el vaco

    23/36

    0.20 0.25 0.30 0.35 0.40

    x m

    -0.1

    -1.0

    -10.0

    2/m

    2

    2 , G

    2,

    Figure 7: Comparison of the energy densities for a narrow Gauian, G with w = 0.003/m, anda -function background, , located at a = 0.3/m. The coupling is = 3m. Note that we haveomitted the (x a) piece for . Its coefficient is such that when it is added to the dotted line

    the areas under both curves are equal.

    of width w centered on a circle of radius a, which illustrates our general conclusions aboutthe nature of the Casimir problem.

    4.1 Boundary Conditions for Radial Functions

    In two space dimensions the angular momentum can be restricted to 0 with degen-eracyN = 2 for = 0 and N = 1 for = 0.6 In each channel with >0, we can solve theradial problem by following the steps described in Section 2.2. Specifically, 1 + n

    2 =

    and

    (z) = ddz

    lnz K(z)= 12

    z +K1(z)

    K(z) ( 0) , (84)

    where K1 K1. These functions are smooth for positive arguments z > 0 and quicklyapproach unity for z 1. For > 0 they have a simple pole at z= 0, (z) ( 12)/z.However, the functions g andh are regular near r = 0,

    g(it,r) = g(it, 0) + O(r2) h(it,r) =r+ O(r3) (85)

    canceling the potential singularities at r = 0 in eqs. (25) and (27) coming from (z).For = 0 it appears from eqs. (23) and (24) that both h0 and the Greens function G0

    become illdefined for= = 0. Ultimately, the source of the problem can be traced to the

    subleading logarithmic term in the Bessel functions,

    w0(kr) =i

    2krH

    (1)0 (kr) and 0(z) =

    1

    2z 1

    zln z+ O(z) . (86)

    6In the formula for N, given after eq. (9), we first have to take = 0 and subsequentlyn 2 for theproper treatment of the singularity in (n+ 2) for this case [23].

    23

  • 8/12/2019 Clculo de las energas en el vaco

    24/36

    The subleading term shows up as a singularity in 0(z), while it is harmless in 0(z) itself.The solution is to integrate the differential equation (25) forg0 fromr = down to a pointarbitrarily close to r= 0 and then getg0(it, 0) from eq. (85).

    For the regular solution, we simply omit the normalization 1/2 in the ansatz(23) andset

    0(k, r) =

    kh0(k, r)w0(kr)

    . (87)

    Then eq. (24) becomes

    G0(r,r,k) =g0(k, r)h0(k, r)

    g0(k, 0) . (88)

    The resulting differential equation for h0 is still eq. (27), but the boundary conditions atr= 0 must be modified: It is straightforward to verify that

    h0(it,r) = r ln tr and h0(it,r) = [1 + ln tr] as r 0 (89)

    removes7 all small r singularities in eq. (27). We use eq. (89) to integrate the differentialequation (27) numerically starting at an arbitrarily small but nonzero value r .

    It is then also straightforward, although tedious, to expand g and h in powers of thebackground field according to eq. (31). We thus obtain the Born subtracted local spectraldensity, eq. (14), for each angular momentum quantum number .

    The final computational ingredient in eq. (16) is the renormalized contribution from theFeynman diagrams. In two space dimensions one Born subtraction suffices. Then we have toadd back in the two Feynman diagrams in Fig.2. In the notadpole renormalization schemethe local diagram, eq. (38), is exactly canceled by the counterterm and we only need to addback the twopoint contribution, eq. (41). The corresponding Feynman parameter integralis readily computed to be

    FD(r) =m2r

    16

    0

    dpJ0(pr)(p)

    p+

    p2 4m22m

    arctan p

    2m

    , (90)

    where(p) = 2

    0

    drrJ0(pr)(r). (91)

    4.2 Numerical Results for a Gauian Background Field

    We now apply this method and discuss the numerical results for the total energy andenergy density of a Gauian background field in two dimensions. We consider

    (r) =Ae(ra)2

    2w2 . (92)

    7The logarithmic behavior is actually expected from the free solution: rI0(tr)K0(tr) ln 2 r ln(tr)as r 0, cf. eq. (31).

    24

  • 8/12/2019 Clculo de las energas en el vaco

    25/36

    0 10 20 30 40 50

    channel l

    0.20

    0.15

    0.10

    0.05

    0.00

    Energy:E(l)/m

    a = 1.0/m

    = 3.0w = 0.1/m

    0 1 2 3

    log l

    5

    4

    3

    2

    1

    0

    log(

    E(l)/m)

    Fit:E(l) = c l

    = 1.98

    Figure 8: Contributions E(,a,w) to the total energy from various angular momentum channels. The logarithmic plot on the right is consistent with the asymptotic behavior E 1/2. Theparameters are a = 1.0/m, w= 0.1/mand = 3m.

    This profile describes a ring of radius aand thicknessw. The normalizationAis chosen suchthat

    0 (r)dr= , giving (r) (r a) as w 0. Sending the coupling strength to

    infinity subsequently imposes the Dirichlet boundary condition on a circle of radius a.We first discuss the total energy E[] = E(,a,w) computed from eq. (50). Note that

    no Feynman diagram contribution has to be added to the total energy because in the no-tadpole scheme E

    (1)FD +ECT = 0. Fig. 8 shows the total energy contribution E(,a,w)

    of the angular momentum channel , for a Gauian ring of radius a = 1.0/m and widthw = 0.1/m. Note that the = 0 channel is suppressed relative to >0 by the degeneracyfactorN0/N =

    12

    . Unlike the case in three dimensions, however, higher angular momentumchannels are not favored by a further increasing degeneracy, so the contributions decayrapidly with increasing 1. The largest contribution always comes from = 1. Thelogarithmic plot in Fig. 8confirms that the asymptotic region is reached quickly and thatthe decay of the individual contributions is consistent with E 1/2 for 5. While thisclearly shows the finiteness of the renormalized quantum energy (for fixed a and w), theconvergence is too slow to allow us to sum the -series directly. Instead, we use a speeduptechnique related to Richardsons approach [24], which is particularly powerful in the presentcase, where the asymptotic behavior sets in quickly. We have checked that the results of thismethod are very robust and independent of the details of the extrapolation technique used. Asufficiently high precision in the energy and density calculation can then be achieved by usingonly approximately 30 channels. For details on this summation technique, see Appendix C.

    The computation of the total energy can be sped up even further by subtracting thesecond Born approximation in eq. (16) and adding back the corresponding secondorderFeynman diagram. With this (over)subtraction, the large behavior ofE(,a,w) changesto 1/3. The relevant Feynman diagram is easily computed,

    E(2)FD[] =

    1

    322

    0

    dp(p)2 arctan p

    2m (93)

    25

  • 8/12/2019 Clculo de las energas en el vaco

    26/36

    0 10 20 30 40 50

    channel l

    0.30

    0.25

    0.20

    0.15

    0.10

    0.05

    0.00

    Energy:E(l)/m

    a = 1.0/m

    = 3.0

    w = 0.008/m

    0 1 2 3

    log l

    5

    4

    3

    2

    1

    0

    log(

    E(l)/m)

    Fit:E(l) = c l

    = 1.94

    Figure 9: Contributions E to the total energy from various angular momentum channels . TheGauian background is much narrower than in Fig.8and the asymptotic decayE 1/2 is reachedfor much larger channels only. The parameters are a= 1.0/m, w= 0.008/m and = 3m.

    where denotes the Fourier transform, eq. (91). We have verified that this method givesthe same result as before.

    As the background field approaches a -function we find by explicit numerical calculationthat the total Casimir energy diverges. To illustrate the problem we consider a very narrowconfiguration with w= 0.008/m. As can be seen by comparing Figs. 9and8, E decreasesmore slowly with when w is smaller. For any finite width w the energy contributions Eeventually decrease steeply enough to yield a convergent sum and a finite total energy, butthis asymptotic regime is only reached for much larger when the width is small. Thisis a typical nonuniform behavior: For any fixed width, there is always a channel 0 largeenough such that the energies E decay as 1/

    2 for

    0. On the other hand, for everyfixed channel we can always find a width w small enough such that we are far from theasymptotic region. Eventually, as w 0 the asymptotic region is never reached and thetotal energy diverges.

    One can also understand this divergence in terms of the second Born approximation,which we separated out in the oversubtraction procedure above. For the singular background,(r) =(r a), this diagram contributes

    E(2)FD[(r a)] =

    2a2

    8

    0

    dpJ20 (ap) arctan p

    2m (94)

    which diverges logarithmically at large p just as the sum over channels did without oversub-

    traction. Oversubtraction shifts the divergence from the sum into the momentum integralin the Feynman diagrams. Whichever choice we make, the divergence cannot be renormal-ized away, despite the claims of [12]; it is a physical effect. Its implications for the usualCasimir calculations will be studied elsewhere [7].

    Next we examine the radial energy density (r). As a numerical check, we have verifiedthat its integral overragrees with the total energy computed directly from eq. (50). In Fig.10

    26

  • 8/12/2019 Clculo de las energas en el vaco

    27/36

    0 0.5 1 1.5 2

    r m

    2.0

    1.5

    1.0

    0.5

    0.0

    0.5

    (r)

    /m

    2

    l= 0 (x2)

    l= 1

    l= 2

    l= 5

    0 0.5 1 1.5 2

    r m

    4

    2

    0

    2

    4

    6

    8

    10

    FD

    (r)/m

    2

    Figure 10: Left panel: Contributions of various angular momentum channels to the Born sub-tracted energy density (r). Note that we have multiplied the = 0 contribution by a factor of2 to better display the secular decrease with . Right panel: The Feynman diagram contribution

    FD(r). The parameters for both panels area= 1.0/m, w = 0.1/m and = 3m. See also Fig. 11for the total energy density(r) found by combining the components displayed here.

    we display the contributions of various angular momentum channels to the Born subtractedenergy density, (r), for a relatively wide profile, w = 0.1/a. Generically the contributionsto the energy density from various angular momentum channels vary in magnitude buthave similar radial shapes. As for the total energy, the = 0 channel is suppressed by adegeneracy factor 1/2; the largest contribution comes from = 1 with all higher channelsdecaying quickly with increasing . For narrower widths the decay with is slower, as weexpect from the preceding discussion of the total energy. The smooth asymptotic behaviorallows us again to use an extrapolation method in order to speed up the sum over channels.

    Figure10also shows the renormalized Feynman diagram contribution in the right panel.Since the integral over all radii r is zero, the peak at the location of the background profilemust be compensated by negative densities on both sides. The sign and width of the peakessentially reproduce the Gauian background itself, while the similar peak in the density(r) has the opposite sign and reduces the overall magnitude of the density. As can be seenfrom the total energy density (r) in Fig. 11, combining all the contributions still yields apronounced peak at the location of the background. As we decrease the width of(r), thepeak increases in magnitude and becomes narrower. In the -function limit it develops anonintegrable singularity, unlike the case in one dimension.

    Figs.11and12display the dependence of the energy density for various values ofr on

    the width of the Gauian background field. We see the nonuniform approach to the limitw 0. In particular, it is clear that the energy density approaches a finite limit at any fixedr asw 0, even though the limiting function diverges as r a. Sufficiently far away fromthe location of the Gauian, around|r a| > 3w, the energy density quickly approachesthe limiting form, corresponding to that of a -function background. In contrast, at r = a,no such limiting value exits. This behavior can also be seen in Fig. 12, where we display

    27

  • 8/12/2019 Clculo de las energas en el vaco

    28/36

    0.8 0.9 1 1.1 1.2

    r m

    100

    0

    100

    200

    (r)/m

    2

    w = 0.02 m1

    w = 0.03 m1

    w = 0.05 m1

    w = 0.08 m1

    0.6 0.7 0.8 0.9

    r m

    30

    20

    10

    0

    (r)/m

    2

    w = 0.02 m1

    w = 0.03 m1

    w = 0.05 m1

    w = 0.08 m1

    Figure 11: The energy density in units ofm2 for the Gauian background field located ata= 0.1/mwith strength= 3.0m for various values of the width w. The left panel shows the energy densityfor all radii. The right panel focuses on the region where the densities can be seen to converge

    toward the limiting form.

    the energy density at various values of r as function of the width. The limiting form isapproached more rapidly the farther one is away from the Gauian peak.

    The Dirichlet boundary condition is achieved only by taking (r) (ra), followed bythe limit . However, since the renormalized Casimir energy diverges as approachesa -function, it is impossible to define the Casimir energy for any sharp background field,much less one with infinite coupling. Thus the DirichletCasimir energy of a circle in twodimensions does not exist. But all is not lost: the energy density away from the sourceapproaches a finite limit as the background becomes sharp.

    5. Summary and Outlook

    In this paper we have developed an efficient method to compute oneloop energies andenergy densities in renormalizable quantum field theories. Starting from a Greens functionformalism, we employ Born subtractions to implement definite renormalization conditions,and express the result in a form that allows for direct, efficient calculation after analyticcontinuation to the imaginary k axis. This method allows us to study the limit in whichsmooth background fields approach idealized boundary conditions, illuminating subtleties ofthe process by which standard Casimir calculations are realized in renormalized quantumfield theory. We find that although the Casimir force between rigid bodies and energydensities away from the boundaries are welldefined observables, quantities like the Casimirenergy have divergences that cannot consistently be renormalized. These divergences havephysical consequences in examples such as the stress on a circle, where we have to comparetwo configurations for which the divergent contributions differ. In these cases, the physicaleffects depend strongly on the details of the source.

    We have also shown rigorously that the total energy, as obtained from the integratedenergy density, can be expressed as the integral over single particle energies weighted by the

    28

  • 8/12/2019 Clculo de las energas en el vaco

    29/36

    0 0.02 0.04 0.06 0.08 0.1

    w m

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    (r)/m

    2

    r = 0.8/m

    r = 0.7/m

    r = 0.5/m

    0 0.02 0.04 0.06 0.08 0.1

    w m

    15

    10

    5

    0

    5

    (r)/m

    2

    r = 0.9/m

    Figure 12: The energy density of the Gauian background field located at a = 1.0/m and withstrength = 3.0m. The figure shows the energy densities at various values ofr as functions of thewidth, w. The densities are scaled to (r) atw= 0.1/m.

    change in the density of states, which is given by the derivative of the phase shifts.There are many possible applications of our method. It can be generalized to fermions [25]

    and to situations where the energy spectrum is asymmetric about zero. In the latter case wewould have to add the different contributions from the two sides of the cut in Fig. 1. In suchsystems the background field can carry nonzero charge density, since the contribution to thecharge density from positive and negative energy modes will no longer cancel. It could alsobe calculated with our approach. We could consider the electromagnetic field in the presenceof conductors by decomposing it into a scalar field obeying Neumann boundary conditionsplus a scalar field obeying Dirichlet boundary conditions. We have seen in this paper howto implement Dirichlet boundary conditions with a background potential, and the Neumannboundary condition can be treated in a similar way [26]. Therefore we are equipped toreconsider problems such as the surface tension of a conducting sphere in QED [8]. Ourcalculational method is well suited to soliton calculations, generalizing [27] beyond the caseof reflectionless potentials. Finally, in cases like the GrossNeveu or NambuJonaLasiniomodels, where one introduces the background potential as a nondynamical auxiliary fieldcoupled to a dynamical field, one can search for solitons as selfconsistent solutions to theauxiliary field equations of motion [28,29]. These equations of motion are obtained by firstintegrating out the dynamical field, leaving an effective action for the auxiliary field. Thisprocedure generates expressions involving the full Greens function similar to those we haveconsidered here, which could be computed efficiently with our methods.

    Acknowledgments

    We gratefully acknowledge discussions with E. Farhi and K. D. Olum. N. G. and R. L. J. aresupported in part by the U.S. Department of Energy (D.O.E.) under cooperative researchagreements #DE-FG03-91ER40662 and #DF-FC02-94ER40818. M. Q. and H. W. are sup-ported by the Deutsche Forschungsgemeinschaft under contracts Qu 137/1-1 and We 1254/3-2.

    29

  • 8/12/2019 Clculo de las energas en el vaco

    30/36

    Appendix A: Density of States and the Jost Function

    In this Appendix, we demonstrate the relation (48) between the integral over r of thelocal spectral density and the Jost function

    20 dr [(k, r)]0 =

    2k

    i0 dr

    G(r,r,k) G

    (0)

    (r,r,k)

    = i

    d

    dkln F(k) , (A.1)

    which is valid everywhere in the upper halfplane Imk >0. This formula can be used bothto establish an efficient computation of the total energy on the imaginary axis, and also torelate the density of states for real k to the phase shift.

    We start by differentiating the Wronskian of the Jost solution, f(k, r), and the regularsolution,(k, r) [30],

    d

    drW[f(k, r), (k

    , r)] =

    k2 k2

    f(k, r)(k, r) (A.2)

    where all quantities in this relation are analytic for Im

    k >0. We integrate both sides fromr = 0 to r = R. Since the regular solution (k, r) becomes kindependent at small r, wecan compute the boundary term at r = 0 by replacing k with k and using the standardWronskian,

    W[f(k, r), (k, r)] = (k)12 F(k) , (A.3)

    giving

    W[f(k, R), (k, R)] = (k) 12 F(k) +

    k2 k2

    R0

    dr f(k, r)(k, r) . (A.4)

    Next we differentiate with respect to k, set k = k , and use the representation (22) for the

    Greens function to obtain

    (k) 12F(k)

    W

    f(k, R), (k, R)

    = 12

    k +

    F(k)

    F(k)+ 2k

    R0

    dr G(r,r,k) , (A.5)

    where f(k, R) ddk f(k, R) and F(k) ddk F(k). To eliminate the first term on the righthand side, we subtract the same equation for the noninteracting case, giving

    (k) 12F(k)

    W

    f(k, R), (k, R) (k) 12 W

    f(0) (k, R),

    (0) (k, R)

    =

    F(k)

    F(k)+ 2k R0 dr

    G(r,r,k) G

    (0)

    (r,r,k)

    . (A.6)

    To complete the proof, we have to show that the left hand side of eq. ( A.6) vanishes asR . To see this, we write the boundary condition (18) for the Jost solution in the form

    f(k, R) = w(kR)1 + O(R1)

    , R (A.7)

    30

  • 8/12/2019 Clculo de las energas en el vaco

    31/36

    which can also be inferred from the integral equation obeyed by f(k, r). Differentiating withrespect tok and using the asymptotics of the free Jost solution w(kR),

    w(kR) = d

    dkw(kR) =iR w(kR)

    1 + O(R2)

    ,

    it is easy to show the asymptotic behavior

    f(k, R) =iR f(k, R)1 + O(R2)

    . (A.8)

    The first term on the lefthand side of eq. (A.6) can thus be estimated by

    (k)12F(k)

    W[f(k, R), (k, R)] =

    iR(k) 12

    F(k) W[f(k, R), (k, R)] i(k)

    12

    F(k) f(k, R)(k, r)

    1 + O(R2)

    = i [R+G(R,R,k)] 1 + O(R2) , (A.9)where we have used the Wronskian off and and the definition of the Greens function,eq. (22). Subtracting the analogous equation in the free case, the term proportional to Rdrops out and we are left with

    (k) 12F(k)

    W

    f(k, R), (k, R) (k) 12 W

    f(0) (k, R),

    (0) (k, R)

    = i

    G(R,R,k) G(0) (R,R,k)

    1 + O(R1)

    . (A.10)

    We estimate the largeR behavior of the difference (k, R) G(R,R,k) G(0) (R,R,k)

    from eqs. (A.9) and (A.6),

    i (k, R) [1 + O(R1)] = F(k)F(k)

    + 2k R0

    dr (k, r) . (A.11)

    It is then straightforward to solve for (k, R) i [C(k) +O(1)R] exp(2ikR) at largeR, where C(k) is an Rindependent integration constant. The lefthand side vanishesexponentially for all Imk > 0 as R , so in this limit we obtain eq. (A.1), whichcompletes the proof.

    It is also instructive to explore the consequences of this result on the real axis. If we let k

    approach the real axis from above, we see that to leading order in R1, the integral equation(A.11) allows for a solution going like R exp[2ikR], which violates the bounds in eq. (30) ifk is real. Thus the coefficient of this term must actually be zero, leaving

    F(k)

    F(k)+ 2k

    R0

    drG(r,r,k) G(0) (r,r,k)

    = C(k) exp(2ikR)

    1 + O(R1)

    (A.12)

    31

  • 8/12/2019 Clculo de las energas en el vaco

    32/36

    which oscillates as R when k is real. As is typical for continuum problems, we mustspecify that the limit wherek becomes real is taken after the integral over r to eliminate thecontribution from these oscillations at the upper limit of integration.

    On the real axis, we can relate the Jost function to the phase shift by eq. (52). Takingthe imaginary part of eq. (A.1) and using eq. (54) yields the relationship between the density

    of states and the phase shift,1

    ddk

    =2k

    Im

    0

    G(r,r,k+i) G(0) (r,r,k+i)

    dr= (k) (0) (k) . (A.13)

    We can also rewrite eq. (A.13) as

    2

    0

    dr

    (k, r)(k, r) (0) (k, r)(0) (k, r)

    = 1

    ddk

    (A.14)

    where the momentum on the lefthand side is understood to be defined with the i pre-scription to eliminate the contribution from the oscillations of the wavefunctions at spatialinfinity.

    Appendix B: Two Jost functions in One Dimension

    In this Appendix, we discuss a technique special to one dimension, which applies evenwhen the background field is not symmetric. Scattering in one dimension can be describedin terms of two Jost solutions f(k, x), which obey the boundary conditions

    limx

    f(k, x)eikx = 1 . (B.1)

    The Greens function then is

    G(x,y,k) = iT(k)

    2ik f+(k, x>)f(k, x

  • 8/12/2019 Clculo de las energas en el vaco

    33/36

    The transmission coefficient is most conveniently computed using the Wronskian for f,yielding

    G(x,x,it) = g+(it,x)g(it,x)

    g+(it, 0)g(it, 0) g(it, 0)g+(it, 0) + 2tg+(it, 0)g(it, 0), (B.7)

    for the Greens function at equal spatial arguments. In this form, we can compute the Greensfunction without encountering any oscillatory or exponentially growing contributions.

    Appendix C: Slowly Convergent Series

    In the examples studied in the main text, the sum over angular momentum channels isslowly converging, typically going as 1/2 for large . From a numerical point of view, sucha decay is too slow for a direct summation as it requires to computeO(1000) channels toget a relative precision of 1%. Even worse, the narrow Gauian profiles relevant for Casimirproblems have a tendency to fall like 1/before they eventually turn over to a 1/2 behaviorat channels with large angular momenta,

    0

    1.

    To sum such series efficiently using only a small number of channels, we employ a tech-nique related to Richardsons method [24]. Given a slowly converging sum

    =0

    a, we define

    a function f(t) at the particular points tn by

    f(tn) n

    =0

    a, tn 1n+ 1

    (n= 0, 1, 2, . . . .) (C.1)

    which gives the value of the full infinite series in the limit t 0. We assume thatf(0) can beapproximated by a smooth extrapolation using the points tn, which accumulate aroundt= 0.Thus given values a0, . . . , aN from the first N+ 1 channels, we construct an interpolatingfunction fof a fixed order through the last + 1 points t

    N, t

    N1, . . . , t

    N (for the first

    few channels N < , so in those cases we take =N). The extrapolation to t = 0 gives anapproximation f(0) for the sum of the series and an error estimate based on the assumptionof a smooth underlying function f(t). While new channels are computed, we monitor f(0)and terminate if we have a satisfying error estimate. For our purposes, the best results areobtained from a rational function (or Pade) approximation, since the resulting f(t) has thetendency to anticipate the smooth behavior of the underlying f(t). To illustrate the method,let us look at the example

    =1

    0(+0)

    (C.2)

    which imitates the behavior of our typical sums over channels: It starts with a

    1/ andturns into the converging 1/2 behavior for > 0. In the table below, we show the results ofour method with the standard parameters used in the main text. We also give the numberof channels actually summed and compare with the results obtained by direct summation ofthe same number of channels.

    As we can see from the table, our method improves the accuracy dramatically. For smallturning points 0, the first 10 channels are sufficient to find the exact value. As we go to

    33

  • 8/12/2019 Clculo de las energas en el vaco

    34/36

  • 8/12/2019 Clculo de las energas en el vaco

    35/36

    References

    [1] H. B. G. Casimir, Kon. Ned. Akad. Wetensch. Proc.51 (1948) 793.

    [2] A. A. Grib, S. G. Mamayev and V. M. Mostepanenko, Vacuum Quantum Effects inStrong Fields, St. Petersburg, 1994.

    [3] V.M. Mostepanenko and N.N. Trunov,The Casimir Effect and its Applications, Claren-don Press, Oxford (1997), M. Bordag, U. Mohideen and V. M. Mostepanenko, Phys.Rept.353(2001) 1 [arXiv:quant-ph/0106045],

    [4] K. A. Milton,The Casimir Effect: Physical Manifestations Of Zero-Point Energy, WorldScientific (2001).

    [5] G. Plunien, B. Muller and W. Greiner, Phys. Rept. 134(1986) 87.

    [6] M. Bordag, ed., Proceedings of the Fifth Workshop on Quantum Field Theory Under

    the Influence of External Conditions, Intl. J. Mod. Phys.A17

    (2002) No. 6 & 7.[7] N. Graham R.L. Jaffe, V. Khemani, M. Quandt, M. Scandurra, and H. Weigel,Casimir

    Energies in the Light of Quantum Field Theory, in preparation.

    [8] T. H. Boyer, Phys. Rev.174(1968) 1764.

    [9] N. D. Birrel and P. C. W. Davies,Quantum Fields in Curved Space, Cambridge, Cam-bridge University Press, 1982.

    [10] K. D. Olum and N. Graham,arXiv:gr-qc/0205134.

    [11] S. G. Mamayev and N. N. Trunov, Theor. Math. Phys.38 (1979) 228 [Teor. Mat. Fiz.38(1979) 345].

    [12] K. A. Milton and Y. J. Ng, Phys. Rev. D 46(1992) 842, S. Leseduarte and A. Romeo,Ann. Phys. 250(1996) 448 [arXiv:hep-th/9605022].

    [13] S. Blau, M. Visser and A. Wipf, Nucl. Phys. B 310(1988) 163.

    [14] E. Farhi, N. Graham, P. Haagensen and R. L. Jaffe, Phys. Lett. B 427 (1998) 334[arXiv:hep-th/9802015], E. Farhi, N. Graham, R.L. Jaffe, and H. Weigel, Phys. Lett.B 475 (2000) 335 [arXiv:hep-th/9912283], Nucl. Phys. B 585 (2000) 443 [arXiv:hep-th/0003144], Nucl. Phys. B 630 (2002) 241 [arXiv:hep-th/0112217].

    [15] For a comprehensive review see: N. Graham, R. L. Jaffe and H. Weigel, Int. J. Mod.Phys. A 17 (2002) 846 [arXiv:hep-th/0201148].

    [16] M. Bordag and K. Kirsten, Phys. Rev. D 53(1996) 5753 [arXiv:hep-th/9608070].

    [17] M. Bordag and J. Lindig, J. Phys. A 29(1996) 4481.

    35

    http://arxiv.org/abs/quant-ph/0106045http://arxiv.org/abs/gr-qc/0205134http://arxiv.org/abs/hep-th/9605022http://arxiv.org/abs/hep-th/9802015http://arxiv.org/abs/hep-th/9912283http://arxiv.org/abs/hep-th/0003144http://arxiv.org/abs/hep-th/0003144http://arxiv.org/abs/hep-th/0112217http://arxiv.org/abs/hep-th/0201148http://arxiv.org/abs/hep-th/9608070http://arxiv.org/abs/hep-th/9608070http://arxiv.org/abs/hep-th/0201148http://arxiv.org/abs/hep-th/0112217http://arxiv.org/abs/hep-th/0003144http://arxiv.org/abs/hep-th/0003144http://arxiv.org/abs/hep-th/9912283http://arxiv.org/abs/hep-th/9802015http://arxiv.org/abs/hep-th/9605022http://arxiv.org/abs/gr-qc/0205134http://arxiv.org/abs/quant-ph/0106045
  • 8/12/2019 Clculo de las energas en el vaco

    36/36

    [18] N. Graham, R. L. Jaffe, M. Quandt and H. Weigel, Phys. Rev. Lett. 87(2001) 131601[arXiv:hep-th/0103010].

    [19] N. Graham, R. L. Jaffe, M. Quandt and H. Weigel, Ann. Phys. 293 (2001) 240[arXiv:quant-ph/0104136].

    [20] M. Bordag, J. Phys. A 28 (1995) 755.

    [21] K. Chadan and P.C. Sabatier, Inverse Problems in Quantum Scattering Theory,SpringerVerlag (1977).

    [22] J. Baacke and S. Junker, Phys. Rev. D 49(1994) 2055 [arXiv:hep-ph/9308310].

    [23] E. Farhi, N. Graham, R. L. Jaffe and H. Weigel, Nucl. Phys. B 595 (2001) 536[arXiv:hep-th/0007189].

    [24] H. Press et. al., Numerical Recipes in C++, 2nd edition, Cambridge University Press(2002).

    [25] P. Sundberg and R. L. Jaffe, in preparation.

    [26] E. Farhi, private communication.

    [27] A. S. Goldhaber, A. Litvintsev and P. van Nieuwenhuizen,arXiv:hep-th/0109110.

    [28] R. Alkofer, H. Reinhardt and H. Weigel, Phys. Rept. 265 (1996) 139 [arXiv:hep-ph/9501213].

    [29] J. Baacke and H. Sprenger, Phys. Rev. D 60(1999) 054017 [arXiv:hep-ph/9809428].

    [30] R. Newton, Scattering Theory of Waves and Particles, McGraw-Hill (1966), Section12.1.

    http://arxiv.org/abs/hep-th/0103010http://arxiv.org/abs/quant-ph/0104136http://arxiv.org/abs/hep-ph/9308310http://arxiv.org/abs/hep-th/0007189http://arxiv.org/abs/hep-th/0109110http://arxiv.org/abs/hep-ph/9501213http://arxiv.org/abs/hep-ph/9501213http://arxiv.org/abs/hep-ph/9809428http://arxiv.org/abs/hep-ph/9809428http://arxiv.org/abs/hep-ph/9501213http://arxiv.org/abs/hep-ph/9501213http://arxiv.org/abs/hep-th/0109110http://arxiv.org/abs/hep-th/0007189http://arxiv.org/abs/hep-ph/9308310http://arxiv.org/abs/quant-ph/0104136http://arxiv.org/abs/hep-th/0103010