ebook img

The Heat-Kernel in a Schwarzschild Geometry and the Casimir Energy PDF

29 Pages·0.23 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview The Heat-Kernel in a Schwarzschild Geometry and the Casimir Energy

The Heat-Kernel in a Schwarzschild Geometry and the Casimir Energy 8 9 9 1 Frank Antonsen n University of Copenhagen a J Niels Bohr Institute 9 Blegdamsvej 17, DK-2100 Copenhagen Ø, Denmark 1 2 v 0 Abstract 0 1 We obtain an hybrid expression for the heat-kernel, and from that 0 the density of the free energy, for a minimally coupled scalar field in 1 7 a Schwarzschild geometry at finite temperature. This gives us the 9 zero-point energy density as a function of the distance from the mas- / c sive object generating the gravitational field. The contribution to the q zero-pointenergyduetothecurvatureisextracted too, inthisway ar- - r rivingatarenormalisedexpressionfortheenergydensity(theCasimir g : energy density). We use this to find an expression for other physical v i quantities: internal energy, pressure and entropy. It turns out that X the disturbance of the surrounding vacuum generates entropy. For β r a smalltheentropyispositiveforr > 2M. Wealsofindthattheinternal energy can be negative outside the horizon pointing to the existence of bound states. The total internal energy inside the horizon turns out to be finite but complex, the imaginary part to be interpreted as responsible for particle creation. 1 Introduction One of the most important quantum physical quantities is probably the zero- pointenergy. Itisneeded, forinstance, when wewant tostudy back-reaction, i.e., the influence the matter fields moving in a curved back-ground assert 1 on the back-ground geometry itself. This would be done by solving the Ein- stein equations with the expectation value of the energy-momentum tensor as source: 1 G = R g R = T , µν µν µν µν − 2 h i where we have chosen units such that κ = ~ = c = 1. It is also important for the study of renormalisation properties of a quantum field theory in a curved space-time. In order to find this quantity, we evaluate the heat-kernel corresponding to the equations of motion for a scalar field. This calculation is carried out at finite temperature. The integral of this heat-kernel with respect to some fictitious fifth coordinate, σ, will give the density of Helmholtz’ free energy (its integral over the entire space-time will give the zeta-function, which is essentially just Helmholtz’ free energy). The resulting expression is regularised and renormalised in the subsequent sections. We do this by using the spherical symmetry of the spacetime in order to collect all the unknown bits of the heat kernel into a function g (r,r′;σ) nl depending only upon the radial coordinates and σ. A recursive relation for an asympototic expansion of this unknow function can be found and solved, thereby allowing us to find the heat kernel. From the freen energy one can derive expressions for the entropy and the pressure by using standard thermodynamic relations. We find these and comment on their meaning. We also show that, to lowest order in the mass generating the Schwarzschild geometry, the free energy is that of an infinite family of particles moving in one dimension with an r-dependent mass. 2 Set-Up We consider a minimally coupled scalar field φ moving in a Schwarzschild back-ground, hence the action is 1 S = (∂ φ∂ φgµν µ2φ2) g d4x, (1) µ ν 2 − | | Z p where the metric is given by the standard expression −1 2M 2M ds2 = 1 dt2 1 dr2 r2dθ2 r2sin2θdφ2, − r − − r − − (cid:18) (cid:19) (cid:18) (cid:19) 2 whereM isthemassofthe(classical)massiveobjectgeneratingtheSchwarzschild geometry, a black hole or a star, say. We will put the mass, µ, of the scalar field equal to zero. We can later reinsert it if we find it desirable. The d’Alembertian becomes 1 ∂2 1 ∂ ∂ L2 2 = r2h , h∂t2 − r2∂r ∂r − r2 where h(r) = (1 2M/r) and L2 is the square of the angular momentum − operator. A finite temperature can be included by complexifying the time coordinate t, t τ. As φ describes a Bose field, it becomes periodic in τ, 7→ φ(τ+β) = φ(τ). This implies that the time direction is not only complexified but also compactified – by appropriate scaling we can take τ to lie on the unit circle S1 (see e.g. Ramond [1] or Itzykson and Zuber, [3], for further details on this). The action, furthermore, becomes “euclideanised”. Now, the partition function Z = e−S φ = e− 21φ2φ√|g|d3xdτ φ, D D Z Z R is simply a Gaussian and hence the functional integral can be carried out, the result being Z = (det2)−1/2. (2) Seeforinstance[1]. Thequantityweareparticularlyinterestedin,isHelmholtz’ free energy, F, which is defined as 1 1 F = lnZ = lndet2. (3) −β 2β The internal energy and pressure, which appear in the energy momentum tensor, are related to F by the usual thermodynamic relations, as is the entropy. 3 Functional Determinants 2 The major problem is apparently the calculation of det . For consistency, andin order to establish notation, Iwill give a short introduction to the topic here. Descriptions can be found in e.g. Hawking [2] or Ramond [1]. 3 A priori, the determinant of an operator A must be given by the product of its eigenvalues λ detA = λ. (4) Now, obviously this is not an easy thingYto calculate. This is where the zeta function, ζ (s), and the heat-kernel, G (x,x′;σ), comes into play. Define A A ζ (s) = λ−s, (5) A X then detA = e−ζA′ (0). (6) The heat-kernel is defined through the differential equation ∂ A G (x,x′;σ) = G (x,x′;σ), (7) x A A −∂σ subject to the boundary condition limG (x,x′;σ) = δ(x x′). (8) A σ→0 − The reason for the terminology is transparent: when A is d2 , the Laplacean dθ2 on S1, then ζ (s) ζ(s), where ζ(s) = ∞ n−s, is the Riemann zeta- A ∝ n=1 function, and when A = 2 and σ is the temperature, then the heat kernel ∇ P satisfies the usual heat equation. Denoting the eigenfunctions of A by ψ we have λ G (x,x′;σ) = ψ (x)ψ∗(x′)e−λσ, (9) A λ λ λ X and one easily proves the important relationship 1 ∞ ζ (s) = dσσs−1 d4xG (x,x;σ). (10) A A Γ(s) Z0 Z Notice that the integral is only along the diagonal x = x′. We don’t have to know the eigenfunctions in order to solve the heat equation, in a d- dimensional Euclidean space, for instance, one can show [1] GA(x,x′;σ) = (4πσ)−d/2e−(x−4σx′)2, 4 in Cartesian coordinates. We will later use a mixture of techniques to arrive at a workable expression for the heat kernel. The angular and thermal part will be handled by a mode sum, whereas we will have to make do with an asymptotic expransion for the radial part in order to find the Casimir energy density. We can arrive at an important interpretation of the value of the heat kernel along the diagonal by noting 1 ∂ 1 ∞ f(x) dσσs−1G (x,x;σ), (11) A ≡ −2β ∂s Γ(s) (cid:12)s=0 Z0 (cid:12) (cid:12) is the density of Helmholtz’ free energy (cid:12) F = f(x) g d4x, (12) | | Z p and is thus exactly the quantity we are interested in calculating. 4 Finding the Heat-Kernel We now return to our particular problem, the minimally coupled scalar field in a Schwarzschild back-ground, i.e. the calculation of the determinant of the d’Alembertian. For this particular operator we know that the eigenfunctions can be written ψ (τ,r,Ω) = e−iωnτY (Ω)g (r), (13) λ lm λ where Ω denotes the angles, and where ω is the Matsubara frequency n 2πn ω = n = 0, 1, 2,.... (14) n β ± ± Wehavenointentionoffindingtheeigenvaluesλ, noroffindingg (r),instead λ we will rewrite the d’Alembertian as ω2 1 ∂ ∂ l(l+1) 2 = n r2h . (15) − h − r2∂r ∂r − r2 And the heat-kernel will be written as G(x,x′;σ) = g (r,r′;σ)Y (Ω)Y∗ (Ω′)e−λnl(r)σ−iωn(τ−τ′), (16) nl lm lm nlm X 5 where we have defined l(l+1) ω2 λ (r) + n. (17) nl ≡ r2 h The reasoning behind this formula is as follows. The eigenfunctions can be written according to (13) as a product of a spherical harmonic, a plane wave invloving the Matsubara frequency and the complexified time, and finally some unknown radial function. Clearly, we would expect the heat kernel, being a sum of products of eigenfunctions, to have a similar form, but since we cannot find the radial eigenvalues λ we cannot simply use (13) to find G. On the other hand, it must be possible to expand it on the functions Y (Ω)Y∗ (Ω′) and e−iωn(τ−τ′), hence the expression (16). lm lm A mass, µ, of the scalar field can be reinserted simply by adding µ2 to λ (r). nl The heat equation determines the unknown functions g . We will Taylor nl expand these in σ, writing (asymptotic expansion)1 ∞ gnl(r,r′;σ) = 1 e−(r−4σr′)2 ak(r,r′)σk, (18) √4πσ k=0 X where the first term is the form g would have in a flat space-time. The nl boundary condition implies a 1. This asymptotic expansion is similar 0 ≡ to the standard Schwinger-DeWitt expansion [5, 6]. The difference here lies in the fact that we only make an asymptotic expansion of the radial part of the heat kernel, whereas we use a mode sum for the remaining part. Our expansion is thus a hybrid, combining the asymptotic Schwinger-DeWitt expansion with the exact mode sum. Inserting this expansion into the heat equation yields a recursion relation for the a ’s. As we only need to know the value of the heat-kernel along k the diagonal x = x′, we will put r = r′ to arrive at the following recursion relation by straightforward computation M 1+k + a = L a L a , (19) k+1 1 k 2 k−1 r − − (cid:18) (cid:19) where we have defined the operators 2M d2 2 8M d L = 1 + 1 − r dr2 r − r2 dr (cid:18) (cid:19) (cid:18) (cid:19) 1For simplicity we will follow the common abuse of terminology and refer to this ex- pression as “the heat kernel” although it strictly speaking is only an asymptotic formula valid for σ not too large. 6 d2 2 d = h + (2h 1) , (20) dr2 r − dr d L = hλ′ +2hλ′ r−1 +h′λ′ +hλ′′ . (21) 2 nldr nl nl nl Putting H = L a = 2hλ′ r−1 +h′λ′ +hλ′′ , (22) 2 0 nl nl nl the first few functions a becomes k a = 1, (23) 0 a = 0, (24) 1 2H a = − , (25) 2 5 h − ( 1)222 H a = − L , (26) 3 1 7 h 5 h − − ( 1)323 1 H ( 1)222 H a = − L L + − L , (27) 4 1 1 2 9 h 7 h 5 h 9 h 5 h − − − − − ( 1)424 1 1 H ( 1)323 1 H a = − L L L + − L L + 5 1 1 1 1 2 11 h 9 h 7 h 5 h 11 h 9 h 5 h − − − − − − − ( 1)323 1 H − L L . (28) 2 1 11 h 7 h 5 h − − − There are singularities in these at r = 2M, as one would expect. Noting that both H and L are linear in ω2, we see that a can contain all even powers 2 n k 2[k/2] of the Matsubara frequencies up to and including ω . n The method of summation of modes when calculating zeta functions have been applied in a number of other spacetimes by various authors, [10], and a number of explicit results exist for some rather simple cases. 5 The Free Energy Density We have now found an asymptotic expansion for the heat-kernel and are ready to integrate it. The σ integration is trivial (it usually is), and will just give a sum of Γ-functions. Explicitly 1 ∞ 1 ∞ f˜(x;s) dσ Y (Ω) 2 a e−λnl(r)σσs+k−3/2 lm k ≡ Γ(s) | | √4π Z0 nlm k=0 X X 7 = Y 2 1 ∞ a (r)Γ(s+k − 12)λ21−s−k. (29) | lm| √4π k Γ(s) nl nlm k=0 X X One should note that even though the asymptotic series is only valid for σ small, this integral still makes sense since the factor λ acts as a mass term nl regularisation. If we hadn’t used our hybrid method we would have had to assume a mass for the scalar field in order to obtain a meaningfull integral. We can still do this, if we wish, by simply adding µ2 to λ . The addition of nl ˜ such a finite mass, however, will not modify the convergence properties of f, hence the succesive regularisations to be performed below are still needed. This is so because only the sum over n need to be regularised, and for fixed l, λ acts like a finite mass irrespective of whether µ2 = 0 or not. nl 6 For k = 0, differentiating this with respect to s at s = 0 simply amounts 6 to removing the Γ-function in the denominator and putting s = 0 in the remaining terms. Thek = 0termissingular andhastobetreatedseparately. We have f˜ (x;s) = Y 2 1 Γ(s− 21)λ21−s, k=0 | lm| √4π Γ(s) nl nlm X sowehaveadivergentsum(fors = 0)inthenumerator,namely √z2 +4π2n2, n where z2 = l(l+1)h(r)β2. We can carry out the sum for an arbitrary s, see e.g. r2 P Ramond [1]: ∞ 1 z2 +4π2n2 21−s = 1 z1−2s + 2ζ(2s+1)+ Γ(s) Γ(s) " (4π2)s−21 n=−∞ X (cid:0) (cid:1) 2 ∞ x2ν ( 1)ν 1 − Γ(s+ν )ζ(2s+2ν 1) . Γ(s+ 21) ν=1 ν! (4π2)s+ν−21 − 2 − # X Differentiating and putting s = 0 we obtain 1 z2 2 ∞ z2ν ( 1)ν 1 z + (γ 2ln2π)(4π )+ − Γ(ν )ζ(2ν 1), 2π3/2 − − π √π ν! (4π2)ν−1 − 2 − 2 ν=2 X which is finite. This expression is to be multiplied by β−1h−1/2. The free energy density f(x) is then ˜ 1 ∂f f(x) = −2β ∂s (cid:12) (cid:12)s=0 (cid:12) (cid:12) (cid:12) 8 1 l(l+1) = Y 2(4π)−1/2 2√π −2β | lm| − r2 − " r lm X 1 4π l(l+1)√hβ 2√π (γ 2ln2π) + 2π3/2 − √hβ − πr2 ! ∞ ( 1)ν l(l+1)hβ2 ν 1 4√πβ−1h−1/2 − Γ(ν )ζ(2ν 1) ν! 4π2r2 − 2 − − # ν=2 (cid:18) (cid:19) X ∞ 1 1 Ylm 2(4π)−1/2 ak(r)λnl(r)21−kΓ(k ). (30) −2β | | − 2 nlm k=2 X X On the face of it, this function depends on all the coordinates, but looking closer we see that there is no explicit time-dependence, and using the sum rule 2l+1 Y (Ω)Y∗ (Ω′) = P (u u′), lm lm 4π l · m X where u,u′ are unit vectors given by the solid angles Ω,Ω′, we see that [4] 2l+1 Y (Ω) 2 = P (1) lm l | | 4π m X [1l] 2l+1 2 ( 1)ν(2l 2ν)! = − − . 2l+2π ν!(l ν)!(l 2ν)! ν=0 − − X So the angular dependency also disappears. In accordance with what we would expect, the energy density depends only on the radial coordinate. We will write the density as a mode sum f(r) = f (r), (31) l l X where 1 l(l+1)β2 f (r) π−1β−2 (4π)−3/2(2l+1)2−l + l ≡ 2β r2 ( r 1 4π l(l+1)√hβ2 (γ 2ln2π) + 2π3/2 − √hβ2 − πr2 ! 9 2 ∞ ( 1)ν l(l+1)hβ2 ν 1 − Γ(ν )ζ(2ν 1) β√hπ ν! 4π2r2 − 2 − !− ν=2 (cid:18) (cid:19) X ∞ ∞ 1 1 (4π)−3/2(2l+1)2−l Γ(k )a (r)λ (r)1/2−k k nl 2β − 2 × ) n=−∞k=2 X X [l/2] ( 1)ν(2l 2ν)! − − . (32) ν!(l ν)!(l 2ν)! ν=0 − − X The major problem here is that the coefficients a (r) depend on n through k the Matsubara frequencies. This expression needs regularisation. For instance, when l = 0 we encounter a singularity. Remembering that a can include all even powers of n up to k and including n2[k/2] we write [k] 2 a (r) = b(t)(r)n2t, k k t=0 X where b(s) is independent of n. Defining λ = 4π2/hβ2 we then have k ∞ ∞ ∞ [k/2] ∞ ak|l=0Γ(k − 21)λn120−k = 2 Γ(k − 21)λ21−k b(kt)(r) n2t+1−2k n=−∞k=2 k=2 t=0 n=1 X X X X X ∞ [k/2] 1 = 2 Γ(k )λ21−k b(t)(r)ζ(2k 2t 1), − 2 k − − k=2 t=0 X X which is divergent whenever 2k 2s 1 = 1, as s 1k this can only happen − − ≤ 2 when k = 2,s = 1. 6 The Minkowski Space Contribution and the Casimir Energy Density To get a physical quantity we must subtract the Minkowski space contri- bution [5, 6], because the physical energy must be such that it vanishes in flat spacetime. Doing this we thereby get the “Casimir energy density”. The relationship between this Casimir energy density, ζ-functions and cut-off reg- ularisations has been studied in [9]. This subtracting off the flat spacetime 10

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.