ArticlePDF Available

Zeta-Function Regularization is Uniquely Defined and Well

Authors:
  • Spanish National Research Council Universitat Autonoma de Barcelona

Abstract

Hawking's zeta function regularization procedure is shown to be rigorously and uniquely defined, thus putting and end to the spreading lore about different difficulties associated with it. Basic misconceptions, misunderstandings and errors which keep appearing in important scientific journals when dealing with this beautiful regularization method ---and other analytical procedures--- are clarified and corrected. Comment: 7 pages, LaTeX file
arXiv:hep-th/9308028v1 6 Aug 1993
ZETA-FUNCTION REGULARIZATION IS UNIQUELY
DEFINED AND WELL
E. Elizalde1
Division of Applied Mechanics, Norwegian Institute of Technology,
University of Trondheim, N-7034 Trondheim, Norway
Abstract
Hawking’s zeta function regularization procedure is shown to be rigorously and uniquely
defined, thus putting and end to the spreading lore about different difficulties associated
with it. Basic misconceptions, misunderstandings and errors which keep appearing in im-
portant scientific journals when dealing with this beautiful regularization method —and
other analytical procedures— are clarified and corrected.
PACS numbers: 03.70.+k, 04.20.Cv, 11.10.Gh
1On leave of absence from and permanent address: Department E.C.M., Faculty of Physics, Barcelona
University, Diagonal 647, 08028 Barcelona, Spain; e-mail: eli @ ebubecm1.bitnet
1
This letter is a defense of Hawking’s zeta function regularization method [1], against the
different criticisms which have been published in important scientific journals and which
seem to conclude (sometimes with exagerated emphasis) that the procedure is ambiguous,
ill defined, and that it possesses even more problems than the well known ones which afflict
e.g. dimensional regularization. Our main purpose is to clarify, once and for all, some basic
concepts, misunderstandings, and also errors which keep appearing in the physical litera-
ture, about this method of zeta function regularization and other analytical regularization
procedures. The situation is such that, what is in fact a most elegant, well defined, and
unique —in many respects— regularization method, may look now to the non-specialist as
just one more among many possible regularization procedures, plagued with difficulties and
illdefiniteness.
We shall not review here in detail the essentials of the method, nor give an exhaustive
list of the papers that have pointed out ‘difficulties with the zeta function procedure’. For
this purpose we refer the specialized reader to a book scheduled to appear towards the end
of 1993 [2]. Instead, we shall restrict ourselves to some specific points which are at the very
heart of the matter and which may be interesting to a much broader audience.
The method of zeta function regularization is uniquely defined in the following way. Take
the Hamiltonian, H, corresponding to our quantum system, plus boundary conditions, plus
possible background field and including a possibly non-trivial metric (because we may live
in a curved spacetime). In mathematical terms, all this boils down to a (second order,
elliptic) differential operator, A, plus corresponding boundary conditions. The spectrum of
this operator Amay or may not be calculable explicitly, and in the first case may or may
not exhibit a beautiful regularity in terms of powers of natural numbers. Whatever be the
situation, it is a well stablished mathematical theorem that to any such operator a zeta
function,ζA, can be defined in a rigorous way. The formal expression of this definition is:
ζA(s) = Tr eslnA.(1)
Let us stress that this is completely general. Moreover, the present is by no means the
only kind of zeta function known to the mathematicians, for whom this concept has a wider
character (see, for instance [3]). The zeta function ζA(s) is generally a meromorphic function
(develops only poles) in the complex plane, sC. Its calculation usually requires complex
integration around some circuit in the complex plane, the use of the Mellin transform, and
the calculation of invariants of the spacetime metric (the Hadamard-Minakshisundaram-
Pleyel-Seeley-DeWitt-Gilkey-... coefficients) [4].
2
In the particular case when the eigenvalues of the differential operator A—what is
equivalent, the eigenvalues of the Hamiltonian (with the boundary conditions, etc. taken
into acount)— can be calculated explicitly (let us call them λnand assume they form a
discrete set), the expression of the zeta function is given by:
ζA(s) = X
n
λs
n.(2)
Now, as a particular case of this (already particular) case, when the eigenvalues are of one
of the forms: (i) an, (ii) a(n+b) or (iii) a(n2+b2), we obtain, respectively, the so-called (i)
(ordinary) Riemann zeta function ζR(or simply ζ), (ii) Hurwitz (or generalized Riemann)
zeta function ζH, and (iii) (a specific case of the) Epstein-Hurwitz zeta function ζE H [3].
Finally, depending on the physical magnitude to be calculated, the zeta function must be
evaluated at a certain particular value of s. For instance, if we are interested in the vacuum
or Casimir energy, which is simply given as the sum over the spectrum:
EC=¯h
2X
n
λn,(3)
this will be given by the corresponding zeta function evaluated at s=1:
EC=¯h
2ζA(1).(4)
Normally, the series (3) will be divergent, and this will involve an analytic continuation
through the zeta function. That is why such regularization can be termed as a particular
case of analytic continuation procedure. In summary, the zeta function of the quantum
system is a very general, uniquely defined, rigorous mathematical concept, which does not
admit either interpretations nor ambiguities.
Let us now come down to the concrete situations which have motivated this article. In
Ref. [5], when calculating the Casimir energy of a piecewise uniform closed string, Brevik
and Nielsen where confronted with the following expression ([5], Eq. (52))
X
n=0
(n+β),(5)
which is clearly infinite. Here, the zeta-function regularization procedure consists in the
following. This expressions comes about as the sum over the eigenvalues n+βof the
Hamiltonian of a certain quantum system (here the transverse oscillations of the mentioned
string), i.e. λn=n+β. There is little doubt about what to do: as clearly stated above, the
corresponding zeta function is
ζA(s) =
X
n=0
(n+β)s.(6)
3
Now, for Re s > 1 this is the expression of the Hurwitz zeta function ζH(s;β), which can be
analytically continued as a meromorphic function to the whole complex plane. Thus, the zeta
function regularization method unambiguously prescribes that the sum under consideration
should be assigned the following value
X
n=0
(n+β) = ζH(1; β).(7)
The wrong alternative (for obvious reasons, after all what has been said before), would be
to argue that ‘we might as well have written
X
n=0
(n+β) = ζR(1) + βζR(0),(8)
what gives a different result’. In fact, that ‘the Hurwitz zeta function (and not the ordinary
Riemann)’ was the one ‘to be used’ was recognized by Li et al. [6], who reproduced in this
way the correct result obtained by Brevik and Nielsen by means of a (more conventional)
exponential cutoff regularization. However, the authors of [6] were again misunderstanding
the main issue when they considered their method as being a generalization of the zeta reg-
ularization procedure (maybe just because the generalized Riemann zeta function appears!).
Quite on the contrary, this is just a specific and particularly simple application of the zeta
function regularization procedure .
Of course, the method can be viewed as just one of the many possibilities of analytic
continuation in order to give sense to (i.e., to regularize) infinite expressions. From this
point of view, it is very much related with the standard dimensional regularization method.
In a very recent paper, [7], Svaiter and Svaiter have argued that, being so close relatives,
these two procedures even share the same type of diseases. But precisely to cure the problem
of the dependence of the regularized result on the kind of the extra dimensions (artificially
introduced in dimensional regularization) was —let us recall— one of the main motivations
of Hawking for the introduction of a new procedure, i.e. zeta function regularization, in
physics [1]. So we seem to have been caught in a devil’s staircase.
The solution to this paradox is the following. Actually, there is no error in the examples
of Ref. [7] and the authors know perfectly what they are doing, but their interpretation
of the results may originate a big deal of confusion among non-specialists. To begin with,
it might look at first sight as if the concept itself of analytical continuation would not be
uniquely defined. Given a function in some domain of the complex plane (here, normally, a
part of the real line or the half plane Re s > a, being asome abscisa of convergence), its
4
analytic continuation to the rest of the complex plane (in our case, usually as a meromorphic
function, but this need not in general be so) is uniquely defined. Put it plain, a function
cannot have two different analytic continuations. What Svaiter and Svaiter do in their
examples is simply to start in each case from two different functions of sand then continue
each of them analytically. Of course, the result is different. In particular, these functions
are
f1(s) =
X
n=0
ns(9)
vs.
f2(s) =
X
n=0
nn
a+ 1(s+1)
(10)
continued to s=1, in the first example, which corresponds to a Hermitian massless
conformal scalar field in 2d Minkowski spacetime with a compactified dimension, and
g1(s) =
X
n=0
n3s(11)
continued to s=1, vs.
g2(s) =
X
n=0
n3n
a+ 1s
,(12)
continued to s= 0, in the second example, in which the vacuum energy corresponding
to a conformally coupled scalar field in an Einstein universe is studied. Needless to say,
the number of posibilites to define ‘different analytic continuations’ in this way is literally
infinite. What use can one make of them remains to be seen.
However, what is absolutely misleading is to conclude from those examples that analytical
regularizations ‘suffer from the same problem as dimensional regularization’, precisely the
one that Hawking wanted to cure!. This has no meaning. In the end, also dimensional
regularization is an analytical procedure! One must realize that zeta function regularization
is perfectly well defined, and has little to do with these arbitrary analytic continuations ‘`a
la ζ’ in which one changes at will any exponent at any place with the only restriction that
one recovers the starting expression for a particular value of the exponent s.
The facts are as follows. (i) There exist infinitely many different analytic regularization
procedures, being dimensional regularization and zeta function regularization just two of
them. (ii) Zeta function regularization is, as we have seen, a specifically defined procedure,
provides a unique analytical continuation and (sometimes) a finite result. (iii) Therefore,
zeta function regularization does not suffer, in any way, from the same kind of problem (or
a related one) as dimensional regularization. (iv) This does not mean, however, that zeta
5
function regularization has no problems, but they are of a different kind; the first appears
already when it turns out that the point (let say s=1 or s= 0) at which the zeta function
must be evaluated turns out to be precisely a pole of the analytic continuation. This and
similar difficulties can be solved, as discussed in detail in Ref. [8]. Eventually, as a final step
one has to resort to renormalization group techniques [9]. (v) Zeta function regularization
has been extended to higher loop order by McKeon and Sherry under the name of operator
regularization and there also some difficulties (concerning the breaking of gauge invariance)
appear [10]. (vi) But, in the end, the fundamental question is: which of the regularizations
that are being used is the one choosen by nature? In practice, one always tries to avoid
answering this question, by cheking the finite results obtained with different regularizations
and by comparing them with classical limits which provide well-known, physically meaningful
values. However, one would be led to believe that in view of its uniqueness, naturalness and
mathematical elegance, zeta function regularization could well be the one. Those properties
are certainly to be counted among their main virtues, but (oddly enough) in some sense
also as its drawbacks: we do not manage to see clearly how and what infinites are thrown
away, something that is evident in other more pedestrial regularizations (which are actually
equivalent in some cases to the zeta one, as pointed out, e.g., in [7]).
The final issue of this paper will concern the practical application of the procedure.
Actually, aside from some very simple cases (among those, the ones reviewed here), the use
of the procedure of analytic continuation through the zeta function requires a good deal of
mathematical work [2]. It is no surprise that has been so often associated with mistakes and
errors [11]. One which often repeates itself can be traced back to Eq. (1.70) of the celebrated
book by Mostepanenko and Trunov [12] on the Casimir effect:
a2
π2
X
n=1 n2+a2m2
π2!1
=1
2m21 + am
πcoth am
π,(13)
in other words (for a=πand m=c),
X
n=1 n2+c21=1
2c2(1 + ccoth c).(14)
That Eq. (13) is not right can be observed by simple inspection. The corrected formula
reads
X
n=1 π2n2+c21=1
2c2(1 + ccoth c).(15)
The integrated version of this equality, namely,
X
n=−∞
ln n2+c22= 2c+ 2 ln 1e2c,(16)
6
under the specific form
T
X
n=−∞
ln h(ωn)2+ (ql)2i=ql+ 2Tln 1eql/T ,(17)
with ωn= 2πnT and ql=πl/R, has been used by Antill´on and Germ´an in a very recent paper
([13], Eq. (2.20)), when studying the Nambu-Goto string model at finite length and non-zero
temperature. Now this equality is again formal. It involves an analytic continuation, since
it has no sense to integrate the lhs term by term: we get a divergent series.
A rigorous way to proceed is as follows. The expression on the left hand side happens
to be the most simple form of the inhomogeneous Epstein zeta function (called usually
Epstein-Hurwitz zeta function [4]). This function is quite involved and different expresions
for it (including asymptotical expansions very useful for accurate numerical calculations)
have been given in [14] (see also [15]). In particular
ζEH (s;c2) =
X
n=1 n2+c2s
=c2s
2+πΓ(s1/2)
2Γ(s)c2s+1 +2πscs+1/2
Γ(s)
X
n=1
ns1/2Ks1/2(2πnc),(18)
which is reminiscent of the famous Chowla-Selberg formula (see [3], p. 1379). Derivatives
can be taken here and the analytical continuation in spresents again no problem.
The usefulness of zeta function regularization is without question [16,2,4]. It can give
immediate sense to expressions such as 1+1 + 1+···=1/2, which turn out to be invaluable
for the construction of new physical theories, as different as Pauli-Villars regularization with
infinite constants (advocated by Slavnov [17]) and mass generation in cosmology through
Landau poles (used by Yndurain [18]). The Riemann zeta function was termed by Hilbert in
his famous 1900 lecture as the most important function of whole mathematics [19]. Probably
it will remain so in the Paris Congress of AD 2000, but now maybe with quantum field physics
adhered to.
Acknowledgments
I am very grateful to Prof. I. Brevik and to Prof. K. Olaussen for many illuminating
discussions and also to them and to Prof. L. Brink for the hospitality extended to me at
the Universities of Trondheim and G¨oteborg, respectively. This work has been supported by
DGICYT (Spain) and by CIRIT (Generalitat de Catalunya).
7
References
[1] S.W. Hawking, Commun. Math Phys. 55, 133 (1977); J.S. Dowker and R. Critchley,
Phys. Rev. D13, 3224 (1976); L.S. Brown and G.J. MacLay, Phys. Rev. 184, 1272
(1969).
[2] E. Elizalde, S.D. Odintsov, A. Romeo, A.A. Bytsenko, and S. Zerbini, Zeta regularization
techniques with applications, World Sci., to be published.
[3] S. Iyanaga and Y. Kawada, Eds., Encyclopedic dictionary of mathematics, Vol. II (The
MIT press, Cambridge, 1977), p. 1372 ff.
[4] E. Elizalde, S. Leseduarte, and S. Zerbini, Mellin transform techniques for zeta-function
resummations, Univ. of Barcelona preprint, UB-ECM-PF 93/7 (1993).
[5] I. Brevik and H.B. Nielsen, Phys. Rev. D41, 1185 (1990).
[6] X. Li, X, Shi, and J. Zhang, Phys. Rev. D44, 560 (1991).
[7] B.F. Svaiter and N.F. Svaiter, Phys. Rev. D47, 4581 (1993).
[8] S.K. Blau, M. Visser, and A. Wipf, Nucl. Phys. B310, 163 (1988).
[9] E. Elizalde and K. Kirsten, to be published.
[10] D.G.C. McKeon and T.N. Sherry, Phys. Rev. Lett. 59, 532 (1987); Phys. Rev. D35,
3854 (1987); A. Rheban, Phys. Rev. D39, 3101 (1989).
[11] E. Elizalde and A. Romeo, Phys. Rev. D40, 436 (1989).
[12] Mostepanenko and Trunov, The casimir effect (in russian), Atomiadz, Moscow, 1991.
[13] A. Antill´on and G. Germ´an, Phys. Rev. D47, 4567 (1993).
[14] E. Elizalde, J. Math. Phys. 31, 170 (1990).
[15] I. Brevik and I. Clausen, PÆS3hys. Rev. D39, 603 (1989).
[16] A. Actor, J. Phys. A24, 3741 (1991); D.M. McAvity and H. Osborn, Cambridge preprint
DAMTP/92-31 (1992), to be published in Nucl. Phys. B.
[17] A. Slavnov, to be published.
8
[18] F. Yndurain, to be published.
[19] C. Reid, Hilbert (Springer, Berlin, 1970), p. 82.
9
... yields, with the same method (namely, analytic continuation by means of the zeta function), the value −1/12 [11][12][13][14] 1 + 2 + 3 + 4 + 5 + 6 + 7 + . . . ≈ −1/12. ...
... and it happens that, when the most advanced physical theories make an appropriate use of such a discovery, this apparent "foolishness" or "arbitrariness" (once well worked out in the complex field [11][12][13][14]) acquires an extraordinary value in describing with incredible precision the phenomena of the real world. Leonhard Euler devoted great efforts to studying infinite series. ...
... At the value s = −1 it yields 1/24. The deviation (a factor of 4) could be crucial, if this would correspond to an actual physical experiment [12]. ...
Article
Full-text available
It is advisable to avoid and, even better, demystify such grandiose terms as “infinity” or “singularity” in the description of the cosmos. Its proliferation does not positively contribute to the understanding of key concepts that are essential for an updated account of its origin and evolutionary history. It will be here argued that, as a matter of fact, there are no infinities in physics, in the real world: all that appears, in any given formulation of nature by means of mathematical equations, actually arises from extrapolations, which are made beyond the bounds of validity of the equations themselves. Such a crucial point is rather well known, but too often forgotten, and is discussed in this paper with several examples; namely, the famous Big Bang singularity and others, which appeared before in classical mechanics and electrodynamics, and notably in the quantization of field theories. A brief description of the Universe’s history and evolution follows. Special emphasis is put on what is presently known, from detailed observations of the cosmos and, complementarily, from advanced experiments of very high-energy physics. To conclude, a future perspective on how this knowledge might soon improve is given.
... yields, with the same method (namely, analytic continuation by means of the zeta function), the value −1/12 [11][12][13][14] 1 + 2 + 3 + 4 + 5 + 6 + 7 + . . . ≈ −1/12. ...
... and it happens that, when the most advanced physical theories make an appropriate use of such a discovery, this apparent "foolishness" or "arbitrariness" (once well worked out in the complex field [11][12][13][14]) acquires an extraordinary value in describing with incredible precision the phenomena of the real world. Leonhard Euler devoted great efforts to studying infinite series. ...
... At the value s = −1 it yields 1/24. The deviation (a factor of 4) could be crucial, if this would correspond to an actual physical experiment [12]. ...
Preprint
Full-text available
It is advisable to avoid and, even better, demystify such grandiose terms as "infinity" or "singularity" in the description of the cosmos. Its proliferation does not positively contribute to the understanding of key concepts that are essential for an updated account of its origin and evolutionary history. It will be here argued that, as a matter of fact, there are no infinities in physics, in the real world: all which appear, in any given formulation of nature by means of mathematical equations, actually arise from extrapolations, which are made beyond the bounds of validity of the equations themselves. Such crucial point is rather well-known, but too often forgotten, as discussed in this paper with several examples; namely, the famous Big Bang singularity and others, which appeared before in classical mechanics and electrodynamics, and notably in the quantization of field theories. A brief description of the Universe’s history and evolution follows. Special emphasis is put on what is presently known, from detailed observations of the cosmos and, complementarily, from advanced experiments of very-high-energy physics. To conclude, a future perspective on how this knowledge might soon improve is given.
... In this chapter we review the method of zeta function regularization [23] in two different cases. First, the case when the fluctuation spectrum is known, which leads to analytic continuations of divergent sums. ...
... The parameter µ with dimension of mass is introduced in order that the energy has the correct dimension for all values of s. One can show [23], that ζ D (s) has a well-defined analytic continuation as meromorphic function over the whole complex plane s ∈ C. The 1-loop contribution to the energy of a classical field configuration in Zeta function regularization is then defined as the value of the analytic continuation of ζ D (s) at s = − 1 2 : ...
... Therefore, we will now focus on the computation of the aforementioned object det(L) . Naively, one approach is to extract the spectrum of the L operator and compute its determinant using zeta-function regularization [72][73][74][75][76][77][78][79]. However, due to the complexity of the operator, this procedure is not straightforward, as it is not simple to solve the spectral problem. ...
Preprint
Full-text available
We provide non-trivial checks of the recently proposed duality between double-scaled SYK and a 2d dilaton gravity model with sine potential, studying the path integral at one-loop level. Specifically, we compute the logarithmic correction to the free energy of sine-dilaton gravity and, up to potential ordering ambiguities, we find a match with the corresponding quantity in double-scaled SYK. The computation relies on the description of sine-dilaton gravity in terms of a version of the q-Schwarzian theory, the quantum deformation of the standard Schwarzian model dual to JT gravity. A crucial aspect of the calculation is selecting the correct Hartle-Hawking vacuum for the gravitational theory, which implies a specific choice of boundary conditions for the one-loop determinant, computed using a generalization of the Gel'fand-Yaglom's theorem. We also evaluate the gravitational one-loop correction to the boundary to boundary propagator of a non-minimally coupled matter field in the bulk theory, showing a perfect agreement with the corresponding quantum correction of matter correlators in double-scaled SYK.
... These authors were very happy to have obtained such formula, as we can see if we read their original articles. In Ref. [39], a first attempt was carried out by the author of the present paper to try to generalize this formula to the case of nonhomogeneous zeta functions, since this is crucial for physical applications, see Ref. [40]. This extension was still in dimension two, what was then believed to be a constraint of the CS expression (see, e.g., Ref. [41]). ...
Article
Full-text available
This is a very basic and pedagogical review of the concepts of zeta function and of the associated zeta regularization method, starting from the notions of harmonic series and of divergent sums in general. By way of very simple examples, it is shown how these powerful methods are used for the regularization of physical quantities, such as quantum vacuum fluctuations in various contexts. In special, in Casimir effect setups, with a note on the dynamical Casimir effect, and mainly concerning its application in quantum theories in curved spaces, subsequently used in gravity theories and cosmology. The second part of this work starts with an essential introduction to large scale cosmology, in search of the observational foundations of the Friedmann-Lemaître-Robertson-Walker (FLRW) model, and the cosmological constant issue, with the very hard problems associated with it. In short, a concise summary of all these interrelated subjects and applications, involving zeta functions and the cosmos, and an updated list of the pioneering and more influential works (according to Google Scholar citation counts) published on all these matters to date, are provided.
Article
This paper shows how an application of zeta function regularisation to a physical model of quantum measurement yields a solution to the problem of wavefunction collapse. Realistic measurement dynamics based on a particle becoming non-isolated are introduced and, based on this, an outcome function is derived using the method of maximum entropy. It is shown how regularisation of an information theoretic quantity related to this outcome function leads to apparent collapse of the wavefunction. The physical principles and key assumptions that underlie this theory are discussed. Some possible experimental approaches are described.
Article
Two singularity theorems can be proven if one attempts to let a Lorentzian cobordism interpolate between two topologically distinct manifolds. On the other hand, Cartier and DeWitt-Morette have given a rigorous definition for quantum field theories (QFTs) by means of path integrals. This paper uses their results to study whether QFTs can be made compatible with topology changes. We show that path integrals over metrics need a finite norm for the latter and for degenerate metrics, this problem can sometimes be resolved with tetrads. We prove that already in the neighborhood of some cuspidal singularities, difficulties can arise to define certain QFTs. On the other hand, we show that simple QFTs can be defined around conical singularities that result from a topology change in a simple setup. We argue that the ground state of many theories of quantum gravity will imply a small cosmological constant and, during the expansion of the universe, will cause frequent topology changes. Unfortunately, it is difficult to describe the transition amplitudes consistently due to the aforementioned problems. We argue that one needs to describe QFTs by stochastic differential equations, and in the case of gravity, by Regge calculus in order to resolve this problem.
Chapter
In this chapter the following applications of the method of zeta-function regularization are described. Firstly, some aspects of the comparison that was established by Fujikawa between the generalized Pauli–Villars regularization method and of the covariant regularization of composite current operators are investigated. Secondly, a calculation of the Casimir energy for the transverse oscillations of a piecewise uniform closed string is performed in detail. The string consists of two parts, each having in general different tension and mass density, but adjusted in such a way that the velocity of sound always equals the velocity of light. This model was introduced by I. Brevik and H.B. Nielsen. For the calculation, an interesting modification of the zeta function method as described up to now needs be performed, in the sense that it must be combined with some basic theorems of complex analysis (as the Cauchy or argument theorems). Also, a comparison with the results obtained by means of the introduction of a cut-off will be established which provides additional physical insight to the zeta function procedure. Hadamard regularization is also discussed, as a very useful auxiliary tool to the zeta method, in dealing with additional infinities and physical cut-offs. This aspect of comparing zeta-function’s analytic continuation with other regularization procedures is the main point in common in the examples studied here.
Article
Full-text available
The Casimir effect for rectangular boxes has been studied for several decades. But there are still some unclear points. Recently, there are new developments related to this topic, including the demonstration of the equivalence of the regularization methods and the clarification of the ambiguity in the regularization of the temperature-dependent free energy. Also, the interesting quantum spring was raised stemming from the topological Casimir effect of the helix boundary conditions. We review these developments together with the general derivation of the Casimir energy of the p-dimensional cavity in (D + 1)-dimensional space–time, paying special attention to the sign of the Casimir force in a cavity with unequal edges. In addition, we also review the Casimir piston, which is a configuration related to rectangular cavity.
Article
Zeta function regularization is an effective method to extract physical significant quantities from infinite ones. It is regarded as mathematically simple and elegant but the isolation of the physical divergency is hidden in its analytic continuation. By contrast, Abel-Plana formula method permits explicit separation of divergent terms. In regularizing the Casimir energy for a massless scalar field in a D-dimensional rectangular box, we give the rigorous proof of the equivalence of the two methods by deriving the reflection formula of Epstein zeta function from repeatedly application of Abel-plana formula and giving the physical interpretation of the infinite integrals. Our study may help with the confidence of choosing any regularization method at convenience among the frequently used ones, especially the zeta function method, without the doubts of physical meanings or mathematical consistency.
Article
Full-text available
As a further step in the general program of zeta‐function regularization of multiseries expressions, some original formulas are provided for the analytic continuation, to any value of s, of two‐dimensional series of Epstein–Hurwitz type, namely, ∑∞n1,n2=0[a1(n1+c1)2 +a2(n2+c2)2]−s, where the aj are positive reals and the cj are not simultaneously nonpositive integers. They come out from a generalization to Hurwitz functions of the zeta‐function regularization theorem of the author and Romeo [Phys. Rev. D 40, 436 (1989)] for ordinary zeta functions. For s=−k,0,2, with k=1,2,3,..., the final results are, in fact, expressed in terms of Hurwitz zeta functions only. For general s they also involve Bessel functions. A partial numerical investigation of the different terms of the exact, algebraic equations is also carried out. As a by‐product, the series ∑∞n=0exp[−a(n+c)2], a,c>0, is conveniently calculated in terms of them.
Article
The zero-point fluctuations of the electromagnetic field give rise to an attractive force between two perfectly conducting parallel plates, the Casimir force. We discuss the structure of the electromagnetic stress-energy tensor in the region between the plates for finite temperatures as well as for the zero-temperature limit, and we describe the relationship of its components to the thermodynamic variables of the radiation field. The stress-energy tensor is defined so that infinite quantities never appear, and it is explicitly computed with the aid of an image-source construction of the Green's function. The finite-temperature case involves both an infinite set of spatial images and an infinite sum of temperature-dependent images.
Article
The Casimir surface force on a solid ball is calculated, assuming the material to be dispersive and to be satisfying the condition ε(ω)μ(ω)=1, ε(ω) being the spectral permittivity and μ(ω) the spectral permeability. A recent paper by Brevik and Einevoll [Phys. Rev. D 37, 2977 (1988)] discussed this problem at zero temperature. In this paper, the case of finite temperatures is covered. The analysis is based upon use of the Debye expansion of the modified Bessel functions. This expansion, being an asymptotic one, is limited in accuracy. At general temperatures, the expansion shows an oscillatory variation on the second-order level. Such a variation is unphysical, so that the physical information that one can extract from second-order theory is limited. On the zeroth-order level, the formalism is, however, found to work well. We discuss the zeroth-order approximation in complete form and comment upon the second-order correction. Although numerical methods are in general necessary, useful analytic approximations can be obtained in the limiting cases of very low, or very high, nondimensional temperatures. At low temperatures, the dispersive effect induces a strong, attractive contribution to the force. At high temperatures, the force becomes small.
Article
The effective Lagrangian and vacuum energy-momentum tensor 〈Tμν〉 due to a scalar field in a de Sitter-space background are calculated using the dimensional-regularization method. For generality the scalar field equation is chosen in the form (□2+ξR+m2)ϕ=0. If ξ=1/6 and m=0, the renormalized 〈Tμν〉 equals gμν(960π2a4)-1, where a is the radius of de Sitter space. More formally, a general zeta-function method is developed. It yields the renormalized effective Lagrangian as the derivative of the zeta function on the curved space. This method is shown to be virtually identical to a method of dimensional regularization applicable to any Riemann space.
Article
The energy due to zero point fluctuations of the dual string is calculated and shown to be divergent. We make it finite by introducing a cut-off. The result is non-covariant but invariant under reparametrization, so one has to modify the classical Lagrangian so as to get a covariant result. It is then shown that this leads to the well-known relation relating the mass squared of the lowest state to the number of dimensions space-time. This result is independent of the cut-off and shows clearly the physical significance of the critical dimension of space-time. Similar considerations for the Neveu-Schwarz model are also discussed.
Article
A framework allowing for perturbative calculations to be carried out for quantum field theories with arbitrary smoothly curved boundaries is described. It is based on an expansion of the heat kernel derived earlier for arbitrary mixed Dirichlet and Neumann boundary conditions. The method is applied to a general renormalisable scalar field theory in four dimensions using dimensional regularisation to two loops and expanding about arbitrary background fields. Detailed results are also specialised to an O(n) symmetric model with a single coupling constant. Extra boundary terms are introduced into the action which give rise to either Dirichlet or generalised Neumann boundary conditions for the quantum fields. For plane boundaries the resulting renormalisation group functions are in accord with earlier results but here the additional terms depending on the extrinsic curvature of the boundary are found. Various consistency relations are also checked and the implications of conformal invariance at the critical point where the β\beta function vanishes are also derived. The local Scr\"odinger equation for the wave functional defined by the functional integral under deformations of the boundary is also verified to two loops. Its consistency with the renormalisation group to all orders in perturbation theory is discussed.
Article
We use zeta function techniques to give a finite definition for the Casimir energy of an arbitrary ultrastatic spacetime with or without boundaries. We find that the Casimir energy is intimately related to, but not identical to, the one-loop effective energy. We show that in general the Casimir energy depends on a normalization scale. This phenomenon has relevance to applications of the Casimir energy in bag models of QCD.Within the framework of Kaluza-Klein theories we discuss the one-loop corrections to the induced cosmological and Newton constants in terms of a Casimir like effect. We can calculate the dependence of these constants on the radius of the compact dimensions, without having to resort to detailed calculations.
Article
This paper describes a technique for regularizing quadratic path integrals on a curved background spacetime. One forms a generalized zeta function from the eigenvalues of the differential operator that appears in the action integral. The zeta function is a meromorphic function and its gradient at the origin is defined to be the determinant of the operator. This technique agrees with dimensional regularization where one generalises ton dimensions by adding extra flat dimensions. The generalized zeta function can be expressed as a Mellin transform of the kernel of the heat equation which describes diffusion over the four dimensional spacetime manifold in a fith dimension of parameter time. Using the asymptotic expansion for the heat kernel, one can deduce the behaviour of the path integral under scale transformations of the background metric. This suggests that there may be a natural cut off in the integral over all black hole background metrics. By functionally differentiating the path integral one obtains an energy momentum tensor which is finite even on the horizon of a black hole. This energy momentum tensor has an anomalous trace.
Article
In this paper we present an alternate way of computing amplitudes in quantum field theory in the context of background-field quantization. We concentrate mainly on one-loop effects. The Feynman diagrams of the usual perturbation series are avoided by first performing the functional integration and then using a perturbative expansion due to Schwinger. In this approach we regulate operators rather than the initial Lagrangian. To one-loop order our scheme reduces to a perturbative expansion of the well-known ζ function associated with the superdeterminant of an operator. This technique preserves all symmetries present in the initial theory and does not lead to any explicit divergences as the regulating parameter approaches its limiting value. For illustration, we apply our approach to a toy (φ3)6 scalar theory, to Yang-Mills theory in the covariant gauge, and to quantum electrodynamics. This method reproduces the usual axial anomaly in the three-point functions VVA and AAA. Operator regularization is used in a dimensionally regulated theory reproducing the usual results obtained in the dimensionally regulated Feynman-diagram approach. An outline of how operator regularization is applied beyond one-loop order is provided. Other possible applications of operator regularization are discussed.