<<

arXiv:0711.1178v1 [hep-th] 7 Nov 2007 prtr n o ih n opt it? compute one might question: how the and operator, considers QFT? talk in important This are why : Introduction 1. ‡ Theory Quantum in Determinants Functional lnr aka T5 unu hoisadSmere,Vall Symmetries, and Theories Quantum QTS5, at talk Plenary • • • ucin n h rc soe nifiiesto unu stat quantum of set infinite an over is trace the and function, rfco stedtriato h utainoeao des operator fluctuation the whil solution: of classical solution, the about classical fluctuations the a is on evaluated prefactor e action leading the the nuclea wherein involves approach, and semiclassical tunneling a in analyzed : semiclassical and Tunneling ie ya xrsino h omΩ=t ln tr = Ω form the of expression an (QFT by theory field given quantum and appea theory These many-body relativistic potentials: canonical grand actions, Effective h hsclmtvto o hsqeto oe rmsevera from comes question this for motivation physical The ayn h rn aoia oeta reetv cinwi action form effective the of or is problem Schwing potential the Symbolically, or canonical field. Hartree-Fock condensate grand on self-consistent the QFT, relativistic solving varying and by theory many-body states in equations: Gap pcrmo ievle.Hr eotrcn rgesi ex in of progress determinants dramati recent functional report Yaglom (i.e., w dimensions I computed and higher Here be can eigenvalues. Gel’fand determinant of functional of spectrum cases. the non-trivial that result so in classical problem compute a to particular difficult problems, in however, and are, physics, They mathematical and theoretical in Abstract. Storrs, Connecticut, of University Physics, of Department Universit¨at Heidelbe Physik, f¨ur Theoretische Institut Dunne V. Gerald plctosi unu edtheory. field quantum in applications σ Z ( x D = ) e φ ucinldtriat fdffrniloeaospa pr a play operators differential of determinants Functional − δσ S [ δ ( φ x ] ≈ ) ndt[ det ln q e(utainoperator) det(fluctuation hti h eemnn faprildifferential partial a of determinant the is what D + e σ − ( S x [ ](2) )] φ cl ] G ndet ln = dld uut2007. August adolid, partial g 92 edleg Germany Heidelberg, 69120 rg, G nrltvsi n non- and relativistic in r pnnilcontribution xponential ieeta prtr) with operators), differential T029 USA 06269, CT where , inpoesscnbe can processes tion ra fphysics: of areas l h next-to-leading the e edn hsapoc to approach this tending nqatmfil theory. field quantum in es. ,adaegenerically are and ), rDsnequations, er-Dyson totcmuigthe computing ithout a n ground find can e o n dimensional one For rbn quantum cribing hrsett a to respect th al ipie the simplifies cally G saGreen’s a is mnn role ominent (1) ‡ Functional Determinants in 2

where is some differential operator. Clearly, to be able to vary the determinant D efficiently, one needs a clever way to evaluate the determinant. Lattice gauge theory: integrating out quarks leads to fermion determinant factors • 4 ¯ ψ ψ¯ e− d x ψ[D/−m]ψ = det (D/ m) (3) D D − Z R which are needed for dynamical fermion computations in QCD. There are still fundamental challenges in evaluating such determinants within the Monte Carlo approach to lattice gauge theory, especially at finite density and temperature. Faddeev-Popov determinants: the determinant of the Faddeev-Popov operator • plays an important role in gauge fixing and confinement. Mathematical physics: the determinant of a partial differential operator, such as a • Dirac or Klein-Gordon operator, encodes interesting spectral information.

2. What we know

Exact results for functional determinants exist only for certain special cases. For example, the gauge theory effective action [the logarithm of the determinant of a Dirac or Klein-Gordon operator] for background fields with (covariantly) constant field strength can be expressed as simple proper-time integrals, building on the seminal work of Heisenberg and Euler [1, 2, 3]. Mathematically, these expressions are closely related to the multiple zeta and gamma functions [4]. This constant curvature case is soluble because we can choose a Fock-Schwinger gauge with A = 1 F x , in which case µ − 2 µν ν the spectral problem decomposes into a set of harmonic oscillator problems, for which the spectral information is very simple. Similarly, in gravitational problems, for special spaces such as harmonic spaces or symmetric spaces, the spectral problem is sufficiently separable and simple that the determinant can be evaluated [5]. For example, the determinant of Dirac or Klein-Gordon operators on spheres, cones and tori are known in quite explicit form [5, 6]. Remarkably, this type of analysis can also be extended to general Riemann surfaces [7, 8], with important applications to worldsheet problems in string theory and with beautiful applications in number theory and differential geometry. Things become more interesting when including general non-constant curvatures [either gauge or gravitational]. Then there are a few especially symmetric cases that are still soluble, but they are all one-dimensional. For example, in gauge theory, if the field strength points in a particular fixed direction, and has a magnitude that is a function of 2 only one coordinate [say xµ], and is of the form F sech (k xµ), then the corresponding [Dirac or Klein-Gordon] spectral problem is hypergeometric, and the determinant can be computed explicitly. Also, for massless quarks in an instanton background, the fermion determinant is known. Such a computation was pioneered by ’t Hooft for the single su(2) instanton [9], but has since been generalized to instantons of full ADHM form [10], to finite temperature [11], and to calorons with nontrivial holonomy [12]. These computations all rely on the fact that the Green’s functions for massless particles in Functional Determinants in Quantum Field Theory 3

an instanton background are known explicitly. The generalization to nonzero masses is nontrivial, and is discussed below in Section 8. Some general bounds are known, based on lattice regularizations and generalizations of Kato’s inequality for Dirac and Klein-Gordon operators. For example [13, 14],

2 2 2 2 det [m D/ ] det m Dµ − 1 ; − 1 . (4) 2 2 h i det [m ∂/ ] ≤ det m2 ∂2 ≥ − µ − h i While very general, these bounds are still quite weak, and do not contain information about the phase of the determinant, which can be physically significant. Tighter bounds have been found recently in QED [15], but general results are still relatively limited. For Schr¨odinger operators with general potentials, or Dirac/Klein-Gordon operators with arbitrary gauge and/or gravitational backgrounds, we rely heavily on approximate methods. One such approximation is the heavy mass expansion. Define ∞ ds 2 ln det m2 + = tr ln m2 + = e−m s tr e−s D (5) D D − 0 s h i h i Z n o Then an inverse mass expansion follows (after renormalization) from the small s asymptotic expansion of the heat kernel operator [16] 1 ∞ tr e−s D sk a [ ] . (6) ∼ (4πs)d/2 k D n o kX=0 Here the a [ ] are known functionals of the potentials appearing in . Similarly, a k D D derivative expansion can be derived by expanding the heat kernel trace tr e−s D about the soluble constant background case, in powers of derivatives of the background.n o This corresponds to resumming all non-derivative terms in the inverse mass expansion. While very general, these expansions are asymptotic, with higher terms becoming rapidly unwieldy, and so have a somewhat limited range of application. Another approach is to define the functional determinant via a zeta function 1 ζ(s) , (7) ≡ λs Xλ where the sum is over the spectrum of the relevant differential operator. Then, by the ′ formal manipulations ζ (0) = λ ln λ = ln ( λ λ), one defines the determinant as ′ − − det = e−ζ (0) . P Q (8) The problem becomes one of analytically continuing ζ(s) from the region in which it converges [typically (s) > d/2] to the neighbourhood of s = 0. For example, in a ℜ radially separable problem, one might estimate the spectrum using WKB phase shifts 1 dδ [and the corresponding spectral function ρ(k) = π dk ], thereby obtaining information about the zeta function. The zeta function approach is complementary to the so-called replica method, where one defines the logarithm of an operator by considering positive integer powers n of the operator, and then analytically continues to n = 0. There are also non-trivial interesting one dimensional quantum mechanical problems for which the determinant can be computed from the zeta function [17]. Functional Determinants in Quantum Field Theory 4

There are several special features of one-dimensional problems that deserve mention. First, the issue of renormalization does not really enter in 1d. Second, the theory of inverse scattering allows us to characterize a potential in terms of the associated scattering data. This is especially useful for solving variational and gap equations, as the variational problem can then be reformulated in terms of the scattering data, which is more direct. Beautiful applications of this idea occur in interacting 1+1 dimensional QFT’s such as the Gross-Neveu model, Sine-Gordon model and their generalizations [18, 19]. [Inverse scattering has been generalized in higher dimensions for radial problems [20]]. A third special feature of 1d problems is the Gel’fand Yaglom theorem [21, 22], which provides a way to compute the determinant of an ordinary differential operator without computing its eigenvalues. The basic idea is that in 1d one can define the Green’s function without using an expansion, simply from the product of two independent solutions.

3. Gel’fand-Yaglom theorem: ordinary differential operators

The simplest form of this result can be stated as follows [21, 22, 23, 24, 25, 26]. Consider a Schr¨odinger operator on the interval x [0, L], with Dirichlet boundary conditions: ∈ d2 + V (x) ψ(x)= λ ψ(x) ; ψ(0) = ψ(L)=0 (9) "−dx2 # Then to compute the determinant it is in fact not necessary to compute the infinite discrete set of eigenvalues λ ,λ ,... . Instead, solve the related initial value problem { 1 2 } d2 + V (x) φ(x)=0 ; φ(0)=0 ; φ′(0)=1 (10) "−dx2 # Then d2 det + V (x) = φ(L) . (11) "−dx2 # It is worth pausing to appreciate the simplicity of this result, as well as its practical utility. It is straightforward to implement numerically. The determinant follows, without having to compute any eigenvalues, let alone multiply them all together! Actually, (11) is not quite right, as we can only define the ratio of two determinants [27], and (11) should be understood in this sense. Often in physical applications one is computing the determinant relative to the corresponding determinant for the free operator, for which the initial value problem is usually known in closed form [in this case it is numerically better to solve directly the initial value problem for the ratio]. As an illustration, consider the following simple example of the massive Helmholtz operator 2 d2 + m2 . The Dirichlet spectrum is λ = m2 + nπ , so from the eigenvalues: − dx2 n L h i 2   2 2 nπ d 2 ∞ m + ∞ 2 det 2 + m L mL sinh(m L) − dx   2 =  2  = 1+ = (12) h d i nπ n π m L det 2 n=1 n=1 "   # − dx Y L Y h i    Functional Determinants in Quantum Field Theory 5

On the other hand, if we use the Gel’fand-Yaglom result, we solve the initial value problems: d2 + m2 φ = 0, to find φ(x)= sinh(mx) ; and d2 φ = 0, to find φ (x)= x − dx2 m − dx2 0 0 for the masslessh operator.i Then (11) implies d2 2 det 2 + m φ(L) sinh(m L) − dx = = (13) h d2 i det φ0(L) m L − dx2 These clearly agree,h buti the point is that if m2 were replaced by a nontrivial potential V (x), the first approach, from the eigenvalues, would be extremely difficult, while the Gel’fand-Yaglom approach is still easy. As another example, consider the P¨oschl- 2 Teller system with H = [ d + m2 j(j + 1) sech2 x], with integer j. This is a − dx2 − reflectionless potential with j bound states at E = m2 l2, (l = 1,...j), and phase j − shift δ(k)=2 l=1 arctan(l/k). The determinant follows directly from the spectrum as d2 2 2 j ∞ det[ P2 + m j(j + 1)sech x] dk dδ ln − dx − = ln(m2 l2)+ ln(m2 + k2)  d2 2  det[ 2 + m ] − 0 π dk − dx Xl=1 Z   Γ(m)Γ(m + 1) = ln (14) Γ(m j)Γ(m + j + 1) − On the other hand, the solutions to the Gel’fand-Yaglom initial value problems (10) are j−m+2 j−m+1 1 Γ Γ φ(x) = 2 2 P m( th L) Qm(th x) Qm( th L)P m(th x) 4m  j+m+2   j+m+1  j − j − j − j Γ 2 Γ 2   1     φ (x)= (cosh(m L) sinh(m x) + sinh(m L) cosh(m x)) (15) (j=0) m on the interval x [ L, L]. The P m and Qm are Legendre functions. Then ∈ − j j d2 2 2 det[ 2 + m j(j + 1)sech x] φ(L) Γ(m)Γ(m + 1) − dx − = lim = (16) d2 2 →∞ det[ + m ] L φ(j=0)(L) Γ(m j)Γ(m + j + 1) − dx2 − Similarly, one can define a general reflectionless potential with N bound states at 2 2 2 d2 E = m κi , (i =1,...N), by V (x)= m 2 dx2 ln det A(x) where A(x) is the N N − c c − − × A = δ + i j e (κi+κj )x. The c are (constant) moduli which affect the shape ij ij κi+κj i N of the potential, but not the spectrum. From the phase shift δ(k)=2 i=1 arctan(κi/k), one finds the determinant P d2 2 N det[ 2 + m + V (x)] m − dx = exp 2 arctanh (17) d2 2 det[ + m ] "− κi # − dx2 Xi=1   It is instructive to verify this numerically for various potentials using (10) and (11). The basic result (11) generalizes in several ways. First, it generalizes to other boundary conditions [24, 26, 28], characterized by two 2 2 matrices M and N as × ψ(0) ψ(L) M ′ + N ′ =0 (18) ψ (0) ! ψ (L) ! 2 Construct two independent solutions u of d + V (x) φ(x) = 0 such that (1),(2) − dx2 ′ h i u(1)(0)=1 ; u(1)(0) = 0 ′ u(2)(0)=0 ; u(2)(0)=1 (19) Functional Determinants in Quantum Field Theory 6

Then the infinite dimensional determinant reduces to a simple 2 2 determinant × 2 d u(1)(L) u(2)(L) × det 2 + V (x) = det2 2 M + N ′ ′ (20) "−dx # " u(1)(L) u(2)(L) !# For example, 1 0 0 0 Dirichlet : M = , N = det = u(2)(L) 0 0 ! 1 0 ! ⇒

0 0 0 1 ′ Neumann : M = , N = det = u(1)(L) 0 1 ! 0 0 ! ⇒ Periodic (P) : M = 1, N = 1 det = 2 u (L)+ u′ (L) − ⇒ − (1) (2)  ′  Antiperiodic (AP) : M = 1, N = 1 det = 2 + u (L)+ u (L) ⇒ (1) (2) ′   In the P/AP cases we recognize (u(1)(L)+ u(2)(L)) as the Floquet-Bloch discriminant, which takes values 2 for P/AP . Second, these results generalize ± straightforwardly to coupled systems of ODE’s [24, 28], and also to linear ODE’s of any order [29]. Third, they generalize to Sturm-Liouville problems [28]. For example, consider the radial operator in d dim., with Dirichlet b.c.’s − − d2 l + d 3 l + d 1 + 2 2 + V (r) . (21) M(l) ≡ −dr2  r2 

Define the function φ(l)(r) as the solution to the radial initial value problem + m2 φ (r)=0 ; φ (r) rl+(d−1)/2 , as r 0 . (22) M(l) (l) (l) ∼ → Then (this shouldh also bei understood as applying to ratios of determinants) det + V (r) = φ (R) . (23) M(l) (l) As an explicit example,h consideri the 2d radial Helmholtz problem, with eigenvalues in 2 2 j(l),n terms of the zeros j(l),n of the Jl Bessel function: λ(l),n = m + R . Then   j 2 2 ∞ m2 + (l),n ∞ 2 det (l) + m R m R M =     = 1+ (24) h i j 2   det (l) n=1 (l),n n=1 j(l),n ! M Y R Y h i      On the other hand, if we use the Gel’fand-Yaglom result (23), we readily solve the relevant initial value problems to find φ(l) in terms of the modified Bessel function Il:

φ(l)(R) l! Il(m R) free = l . (25) φ(l) (R) m R 2   The agreement between (24) and (25) expresses the product formula for Il(x).

4. Gel’fand-Yaglom theorem: partial differential operators?

In generalizing from ordinary to partial differential operators, it is natural to consider first the case of separable operators. But the na¨ıve extension fails. Formally, we simply take a product over the radial determinants [each of which is evaluated as above] for all Functional Determinants in Quantum Field Theory 7

possible angular momenta, with the appropriate degeneracies. However, this product diverges [24]. For example, for the 2d radial Helmholtz system (25)

2 det (l) + m l!2l I (m R) 1 ln M = ln l O , l (26)  deth i (m R)l ! ∼ l →∞ M(l)     The degeneracy factor ish 1 foril = 0, and 2 for all l 1, so the angular momentum ≥ sum over l diverges. This should not be so surprising from a physical point of view, as we expect to require renormalization in dimension higher than 1. To formulate the problem more precisely, consider the radially separable operators = ∆+ V (r), and M − free = ∆, where ∆ is the Laplace operator in IRd, and V (r) is a radial potential M − vanishing at infinity as r−2−ǫ for d = 2 and d = 3, and as r−4−ǫ for d = 4. Since V = V (r), we can separate variables and consider the Schr¨odinger-like radial operator in (21). For dimension d 2, the radial eigenfunctions ψ have degeneracy M(l) ≥ (l) (2l + d 2)(l + d 3)! deg(l; d) − − . (27) ≡ l!(d 2)! − Formally, we write ∞ 2 det [ + m2] det (l) + m ln M = deg(l; d) ln M . (28) det [ free + m2]! det h free + m2i  M Xl=0 M(l)  h i Each term in the sum is computed straightforwardly using (23), but the l sum is divergent. However, this divergence can be understood physically, leading to a finite and renormalized determinant ratio [30]: ∞ det[ + m2] φ ( ) ∞ φ ( ) drrV (r) ln M = ln (0) ∞ + 2 ln (l) ∞ 0 det[ free + m2]! φfree( )  φfree( ) − 2l  d=2 (0) l=1  (l) R  M ∞ ∞ X ∞   µr   + drrV ln + γ  Z0   2   2 ∞ ∞ det[ + m ] φ(l)( ) 0 drrV (r) ln M = (2l + 1) ln ∞ 1 det[ free + m2]!  φfree( ) − 2(l + )  d=3 l=0  (l) R 2  M X ∞ det[ + m2]   ln M =   det[ free + m2]! d=4 M ∞ ∞ ∞ 3 2 2 φ(l)( ) 0 drrV (r) 0 drr V (V +2m ) (l + 1) ln free ∞ + 3  φ ( ) − R 2(l + 1) R 8(l + 1)  Xl=0  (l) ∞  1 ∞   µr  drr3 V (V +2m2) ln + γ +1 . (29) − 8 Z0   2   Here γ is Euler’s constant, and µ is a renormalization scale, which is essential for physical applications. A conventional renormalization choice is to take µ = m in (29). In each of (29), the sum over l is convergent once the indicated subtractions are made. The

function φ(l)(r) is defined in (22). Notice that the results (29) state once again that the determinant is determined by the boundary values of solutions of [ + m2] φ = 0, with the only additional information M Functional Determinants in Quantum Field Theory 8

being a finite number of integrals involving the potential V (r). We also stress the computational simplicity of (29); no phase shifts or eigenvalues need be computed to evaluate the renormalized determinant. These results have been derived [30] using the zeta function formalism. Scattering theory [31] and standard contour manipulations [28, 32] yield an integral representation

[valid for (s) > d/2] of the zeta function in terms of the Jost function fl(ik): ℜ ∞ sin πs ∞ ∂ ζ(s)= deg(l; d) dk [k2 m2]−s ln f (ik) . (30) π − ∂k l Xl=0 mZ The technical problem is the analytic continuation of (30) to a neighborhood about s = 0. This analytic continuation relies on the uniform asymptotic behavior of the Jost

function fl(ik), which follows from standard results in scattering theory [31]: ∞ ∞ r ln f (ik)= drrV (r)K (kr)I (kr) drrV (r)K2(kr) dr′ r′ V (r′)I2(kr′)+ (V 3) l ν ν − ν ν O Z0 Z0 Z0 where ν l + d 1. Then the calculation of the asymptotics of the Jost function ≡ 2 − reduces to the known uniform asymptotics of the modified Bessel functions Kν and Iν. asym 2 Using these asymptotics, we can define ln fl (ik) as the O(V ) and O(V ) parts of this uniform asymptotic expansion, and then by construction, ∞ ζ′ (0) deg(l; d) [ln f (im) ln f asym(im)] . (31) f ≡ − l − l Xl=0 is now convergent. Adding back the subtracted terms in a dimensional scheme, leads [30] to the results in (29).

5. Comparison With Approach

These determinants can also be derived in a Feynman diagrammatic approach, where the perturbative expansion in powers of the potential V is det [ + m2] ∞ ( 1)k+1 ln M − A(k) det [ free + m2]! ≡ k M kX=1 1 1 = + + ... , (32) − 2 3 Here the solid dots denote insertions of the potential V . If the potential is radial, then with dimensional regularization one can express the renormalized 4d determinant as [33] 2 ∞ ∞ det [ + m ] 2 ln M = (l + 1) ln fl(im) drrV (r)Kl+1(mr) Il+1(mr) " det [ free + m2]!# − 0 M d=4 Xl=0  Z ∞ r 2 ′ ′ ′ 2 ′ (1) 1 (2) + drrV (r)Kl+1(mr) dr r V (r )Il+1(mr ) + Afin Afin (33) Z0 Z0  − 2 where the finite contributions from the first and second order Feynman diagrams in the MS scheme are [33] (note the small typo in equation (4.32) of [33]): 2 ∞ (1) m 3 Afin = drr V (r) − 8 Z0 Functional Determinants in Quantum Field Theory 9

∞ 2 √ 2 2 √ 2 2 (2) 1 3 ˜ 4m + q 4m + q + q Afin = dq q V (q) 2 ln (34) 128π4 0 " − q √4m2 + q2 q !# Z − Here V˜ (q) is the 4d Fourier transform of the radial potential V (r). These expressions (33) and (34) look quite different from (29), but in fact they can be shown to be completely equivalent, using some Bessel identities [30]. However, it is clear that (29) is much easier to evaluate, especially if (as often happens) V (r) is known only numerically.

6. Removing Zero Modes

In certain quantum field theory applications the determinant may have zero modes, and correspondingly one is actually interested in computing the determinant with these zero modes removed, together with a collective coordinate factor [23]. For example, consider the case of a self-interacting scalar field theory in d dim Euclidean d space with Euclidean action S[Φ] = d x [∂µΦ∂µΦ+ U(Φ)], where there is a [radially symmetric] classical solution Φ (r) solving ∆Φ = dU , with Φ′(0) = 0, and Φ 0 as cl R dΦ → r . The fluctuation operator about this classical solution is a radial operator, with →∞ d2U 2 V (r) = dΦ2 m . In the l = 1 sector the corresponding radial operator Φ=Φcl(r) − hash a di-fold degenerate zero mode associated with translational invariance: M(l=1) Φ′ (r) ψ (r)= cl . (35) zero Φ′′(0) Including the collective coordinate contribution [23] due to translational invariance, the required determinant factor is [the prime denotes exclusion of zero mode(s)] −1/2 d/2 ′ 2 S[Φ ] det (l=1) + m cl M . (36) 2π !  det h free + m2 i M(l=1)   Since the determinant is expressedh in terms ofi a solution to an ODE, we can use simple ODE theory to show that (generalizing the d = 1 [34] and d = 4 [35] analyses) ′ det + m2 ∞ d−1 2 d (l=1) 0 drr ψzero(r) M = − ′ , (37) h free 2 i d 1 free det + m " limR→∞ R( 2R ψzero(R) ψzero(R))# M(l=1) − free h d/2 d i d/2−1 Here, ψzero(r) = 2 Γ( 2 + 1) Id/2(r)/r , and the classical solution behaves as Kd/2−1(R) ∞ ∞ Φcl(R) Φ Rd/2−1 , as R , for some constant Φ . Thus the R limit exists ∼ →∞ →∞ d ∞ d−1 2 Γ( 2 +1)S[Φcl] and is simple. Furthermore, from the virial theorem, drr ψ (r)= 2 , 0 zero d/2 ′′ π (Φcl(0)) so we arrive at the simple expression for the net contributioR n of this l = 1 mode: −1/2 d/2 ′ 2 S[Φcl] det (l=1) + m d/2−1 ′′ d/2 M = (2π) Φ∞ Φcl(0) (38). 2π !  det h free + m2 i | | M(l=1) h i   This expresses the zero mode factorh solely in termsi of the asymptotic properties of the

classical solution Φcl, which are already known numerically as part of the determination

of Φcl and the fluctuation potential V (r). No further computation is needed. Functional Determinants in Quantum Field Theory 10

7. Application 1: fluctuation determinant for false vacuum decay

The phenomenon of nucleation drives first order phase transitions in many applications in physics. The semiclassical analysis of the rate of such a nucleation process was pioneered by Langer [36], who identified a semiclassical saddle point solution that gives the dominant exponential contribution to the rate, with a prefactor to the exponential given by the quantum fluctuations about this classical solution. The nucleation rate is given by the quantum mechanical rate of decay of a metastable ”false” vacuum, Φ−, into the ”true” vacuum, Φ+. Decay proceeds by the nucleation of expanding bubbles of true vacuum within the metastable false vacuum [36, 37, 38, 39, 40, 41]. The semiclassical prefactor requires computing the determinant of the differential operator associated with quantum fluctuations about the classical solution. This is a technically difficult problem, but it is ideally suited to the above results for determinants of radially separable operators, because the fluctuation operator is radial. The expression (42) below can be applied to any metastable field potential U(Φ) [35], without relying on the ”thin-wall” limit (the limit of degenerate vacua) . For definiteness, consider the specific (rescaled) quartic potential: U(Φ) = 1 Φ2 1 Φ3 + α Φ4, where α characterizes the shape of the 2 − 2 8 potential [the thin-wall limit is α 1]. The decay rate per unit volume and unit time → is − S [Φ ] 2 det′ ( 2 + U ′′(Φ )) 1/2 cl cl cl −Scl[Φcl]−δctS[Φcl] γ = − ′′ e , (39) 2π ! det ( 2 + U (Φ−))

− where the prime on the determinant means that the zero modes (corresponding to

translational invariance) are removed. Here Φcl is a classical solution known as the “bounce” solution [39], defined below, and the prefactor terms in (39) correspond to quantum fluctuations about this bounce solution. The second term in the exponent,

δctS[Φcl], denotes the counterterms needed for renormalizing the classical action Scl. The first step in computing the false vacuum decay rate γ is to find the classical bounce

solution, Φcl(r), which is a radially symmetric stationary point of the classical Euclidean

action, interpolating between the false and true vacuum. The bounce Φcl(r) solves the nonlinear ordinary differential equation 3 3 α Φ′′ Φ′ + Φ Φ2 + Φ3 =0 (40) − cl − r cl cl − 2 cl 2 cl ′ with boundary conditions Φ (0) = 0, Φ (r) Φ− 0, as r . Φ (r) must be cl cl → ≡ → ∞ cl computed numerically, for example by the method of shooting, adjusting the initial value Φ (0) until the boundary conditions at r = are satisfied. Given the bounce cl ∞ solution, Φcl(r), the corresponding radial fluctuation potential is 3α U ′′(Φ (r))=1 3Φ (r)+ Φ2 (r) . (41) cl − cl 2 cl There are three different types of eigenvalue of the fluctuation operator, each having a different role physically and mathematically. (i) Negative Mode: The l = 0 sector has a negative eigenvalue mode of the fluctuation operator, and is responsible for the instability leading to decay. This mode contributes a factor to the decay rate γ related Functional Determinants in Quantum Field Theory 11

to the absolute value of the determinant of the l = 0 fluctuation operator [36, 39]. (ii) Zero Modes: In the l = 1 sector, there is a four-fold degenerate zero eigenvalue of the fluctuation operator. Integrating over the corresponding collective coordinates Scl produces the factors of 2π in (39). In computing the rate γ, we need the determinant of the fluctuation operator with the zero mode removed [36, 39, 40]. This can be evaluated simply using the formula (38), in terms of the asymptotic behavior of the classical bounce solution Φ (r). (iii) Positive Modes: For l 2, the fluctuation operator has positive cl ≥ eigenvalues, each of degeneracy (l +1)2. For each l, the associated radial determinant is computed numerically using (23). But the sum over l is divergent and the answer must

H1-ΑL3 G 0.3

0.28

0.26

0.24

0.22

Α 0.9 0.92 0.94 0.96 0.98

0.18

Figure 1. Comparison as α 1 of the exact numerical result [dots and solid blue → line] from (42) with the analytic thin-wall limit result 9 1 2π 0.16788.. of [42]. 32 − 9√3 ≈ h i be renormalized. Using the MS scheme, the renormalized effective action is [35]:

1 π 3 2 α 3 Γ = ln T ( ) 2 ln Φ∞ Φ Φ + Φ MS 2 (0) ∞ − 2 0 − 2 0 2 0   ∞  ∞ ∞ 1 1 3 1 2 2 0 drrV (r) 8 0 drr V (V + 2) + (l + 1) ln T(l)( ) + 2 ( ∞ − R (l + 1) R (l + 1)3 ) l=2   X∞ ∞ 3 1 3 1 r drrV (r)+ drr V (V + 2) γE ln (42) − 4 Z0 16 Z0 2 − − 2 The first term is from the negative mode, the second from the zero mode(s) [using (38)], the third term from the numerical computation of the higher modes, and the last term is the renormalization counterterm contribution. The expression (42) can be applied to any field potential U(Φ), without relying on the thin-wall limit. A comparison with the analytic thin-wall computation of Konoplich and Rubin [42] is plotted in Fig. 1, and shows excellent agreement. This approach can also be extended to false vacuum decay in curved space, where the structure of the bounce solutions [43], and the corresponding fluctuation equations [44], is much richer. Functional Determinants in Quantum Field Theory 12

8. Application 2: determinant for massive quarks in instanton background

The fermion determinant in an instanton background is known in the massless (chiral) limit [9], and in the heavy quark limit [45]. Here we compute it for any m [46]. The ˜S relevant part of the renormalized effective action, Γren(m), behaves as 1 α(1/2)+ (ln m + γ ln2) m2 + ... , m 0 ˜S 2 − → Γren(m)=  (43)  ln m 1 17 232 7916  + + , m − 6 − 75m2 − 735m4 2835m6 − 148225m8 ··· →∞  where α(1/2) = 0.145873. The small mass expansion [9] is based on the known ≃ massless propagators in an instanton background, while the large mass expansion [45] follows most easily from the Schwinger-DeWitt or heat kernel expansion. Since the instanton is self-dual, we can simplify things slightly and work with the scalar Klein-Gordon operator. For a single su(2) instanton, the spectral problem 1 separates into partial waves [9], with radial Schr¨odinger-like operators (take isospin 2 ) ∂2 3 ∂ 4l(l + 1) 4(j l)(j + l + 1) 3 (l,j) + + − , (44) H ≡ "−∂r2 − r ∂r r2 r2 +1 − (r2 + 1)2 # Here l =0, 1 , 1, 3 , , and j = l 1 , and there is a degeneracy factor of (2l+1)(2j +1). 2 2 ··· | ± 2 | Since the ”potential” involves l and j, we cannot directly use the result (29). In order to compare with ’t Hooft’s massless quark analysis [9], we use a Pauli-Villars regulator Λ, and define the renomalized effective action by charge renormalization. To extract the renormalized effective action we need to consider the Λ limit in conjunction → ∞ with the infinite sum over l. Split the partial wave sum into two parts: L ∞ S S S ΓΛ(m)= ΓΛ,(l)(m)+ ΓΛ,(l)(m) (45) l=0, 1 ,... l=L+ 1 X2 X 2 where L is a large but finite integer. The first sum involves low partial wave modes, and the second sum involves the high partial wave modes. For the low partial wave modes, we can remove the regulator and evaluate the (finite) sum using the (numerical) Gel’fand-Yaglom result (23). But this sum diverges quadratically with L. In the second sum in (45) we cannot take the large L and large Λ limits blindly, as each leads to a divergence. Radial WKB is a good approximation for the high l modes. With WKB we can compute analytically the large Λ and large L divergences of the second sum in (45), using the WKB approximation for the corresponding determinants [46]: ∞ 1 1 m2 ΓS (A; m) lnΛ+2L2 +4L + ln L Λ,(l) ∼ 6 − 6 2 l=L+ 1 ! X 2 127 1 m2 m2 1 + ln2+ m2 ln2+ ln m + O (46) " 72 − 3 2 − 2 # L It is important to identify the physical role of the various terms in (46). The first term is the expected logarithmic charge renormalization counterterm. The next three terms Functional Determinants in Quantum Field Theory 13

˜S Γren(m) 0.15 0.1 0.05 m 0.5 1 1.5 2 2.5 3 -0.05 -0.1 -0.15 -0.2

S Figure 2. Plot of our numerical results for Γ˜ren(m) from (47), compared with the analytic extreme small and large mass limits [dashed curves] from (43). The dots denote numerical data points from (47), and the solid line is a fit through these points.

give quadratic, linear and logarithmic divergences in L. These divergences cancel exactly against corresponding divergences in the first sum in (45), which are found numerically. Combining the numerical results for low partial wave modes with the radial WKB ˜S results for the high partial wave modes the renormalized effective action Γren(m) is

L 2 ˜S 2 1 m Γren(m) = lim  (2l + 1)(2l + 2)P (l)+2L +4L + ln L L→∞ − 6 2 l=0, 1 ,... !  X2 2 2 127 1 m m +  ln2+ m2 ln2+ ln m . (47) " 72 − 3 2 − 2 #) ˜S This is finite for any mass m. Figure 2 shows these results for Γren(m), compared with the analytic small and large mass expansions in (43). The agreement is remarkable [46]. It is a highly nontrivial check on the WKB computation that the correct mass dependence rises to interpolate between the large mass and small mass regimes. As an interesting analytic check, the formula (47) also provides a simple computation [46] of ’t Hooft’s leading small mass result α(1/2) = 17 1 ln2+ 1 2ζ′( 1)=0.145873.... − 72 − 6 6 − −

9. Concluding remarks

In this talk I have reviewed some of the reasons for considering determinants of partial differential operators in quantum field theory. Mathematically, much is known for ordinary differential operators, especially through the work of Gel’fand-Yaglom [21, 22] and its extensions [24, 26, 28], but considerably less is known for partial differential operators with non-trivial potentials. The new mathematical results [30] reported here are the formulas in (29), which provide simple expressions for the determinant of a radially separable partial differential operator of the form ∆ + m2 + V (r). This − Functional Determinants in Quantum Field Theory 14 generalizes the (Dirichlet) Gel’fand-Yaglom result (11) to higher dimensions. These results lead to many direct applications in quantum field theory, where they extend the class of solvable fluctuation determinant problems away from the restrictive class of constant background fields, or one dimensional background fields, to the more general class of separable higher dimensional background fields. This includes, for example, applications in quantum field theory to the study of quantum fluctuations in the presence of vortices, monopoles, sphalerons, instantons, domain walls (branes), etc ... A number of further generalizations could be made. First, it is practically important to include directly the matrix structure that arises from Dirac-like differential operators and from non-abelian gauge degrees of freedom. This has been done in various ways for certain applications [47, 48, 49, 50, 51], but a simpler unified formalism along the lines of the explicit formulas (29) is probably possible, as in one dimension [52]. Second, a practical extension to finite temperature and density would be useful in various physical applications. Third, the biggest technical challenge is to ask if the restriction of separability can be relaxed. A promising current approach is the numerical worldline loop method [53]. Without symmetry, computing functional determinants analytically remains a difficult problem, and any progress will be interesting.

Acknowledgments

I thank the organizers, especially Mariano Del Olmo, for a stimulating conference, in a beautiful city. I acknowledge support from the DFG through the Mercator Guest Professor Program, and from the DOE through grant DE-FG02-92ER40716. [1] W. Heisenberg and H. Euler, “Consequences of Dirac’s theory of positrons,” Z. Phys. 98, 714 (1936); English translation at [arXiv:physics/0605038]. [2] J. S. Schwinger, “On gauge invariance and vacuum polarization,” Phys. Rev. 82, 664 (1951). [3] G. V. Dunne, “Heisenberg-Euler effective Lagrangians: Basics and extensions,” in Ian Kogan Memorial Collection, ’From Fields to Strings: Circumnavigating Theoretical Physics’, M. Shifman et al (ed.) vol. 1 pp 445-522 (World Scientific, 2005). [arXiv:hep-th/0406216]. [4] S. N. M. Ruijsenaars, “On Barnes’ Multiple Zeta and Gamma Functions”, Adv. Math. 156, 107 (2000). [5] R. Camporesi, “Harmonic analysis and propagators on homogeneous spaces,” Phys. Rept. 196, 1 (1990); I. G. Avramidi, “Heat kernel and quantum ,” Lect. Notes Phys. M64, 1 (2000). [6] J. S. Dowker, “Functional determinants on spheres and sectors,” J. Math. Phys. 35, 4989 (1994) [Erratum-ibid. 36, 988 (1995)] [arXiv:hep-th/9312080]; J. S. Dowker and K. Kirsten, “The Barnes zeta-function, sphere determinants and Glaisher- Kinkelin-Bendersky constants,” Anal. Appl. 3, 45 (2005) [arXiv:hep-th/0301143]. [7] E. D’Hoker and D. H. Phong, “On Determinants Of Laplacians On Riemann Surfaces,” Commun. Math. Phys. 104, 537 (1986). [8] P. Sarnak, “Determinants of Laplacians”, Commun. Math. Phys. 110, 113 (1987). [9] G. ’t Hooft, “Computation of the quantum effects due to a four-dimensional pseudoparticle,” Phys. Rev. D 14, 3432 (1976) [Erratum-ibid. D 18, 2199 (1978)]; R. D. Carlitz and D. B. Creamer, “Light Quarks And Instantons,” Annals Phys. 118, 429 (1979). [10] H. Osborn, “Semiclassical Functional Integrals For Selfdual Gauge Fields,” Annals Phys. 135, 373 (1981). Functional Determinants in Quantum Field Theory 15

[11] D. J. Gross, R. D. Pisarski and L. G. Yaffe, “QCD And Instantons At Finite Temperature,” Rev. Mod. Phys. 53, 43 (1981). [12] D. Diakonov, N. Gromov, V. Petrov and S. Slizovskiy, “Quantum weights of dyons and of instantons with non-trivial holonomy,” Phys. Rev. D 70, 036003 (2004) [arXiv:hep-th/0404042]. [13] H. Hogreve, R. Schrader and R. Seiler, “A Conjecture On The Spinor Functional Determinant,” Nucl. Phys. B 142, 525 (1978); R. Schrader and R. Seiler, “A Uniform Lower Bound On The Renormalized Scalar Euclidean Functional Determinant,” Commun. Math. Phys. 61, 169 (1978). D. Brydges, J. Frohlich and E. Seiler, “On The Construction Of Quantized Gauge Fields. I. General Results,” Annals Phys. 121, 227 (1979). [14] C. Vafa and E. Witten, “Eigenvalue Inequalities For Fermions In Gauge Theories,” Commun. Math. Phys. 95, 257 (1984). [15] M. P. Fry, “QED in inhomogeneous magnetic fields,” Phys. Rev. D 54, 6444 (1996) [arXiv:hep-th/9606037]; “Fermion determinant for general background gauge fields,” Phys. Rev. D 67, 065017 (2003) [arXiv:hep-th/0301097]. [16] D. V. Vassilevich, “Heat kernel expansion: User’s manual,” Phys. Rept. 388, 279 (2003) [arXiv:hep-th/0306138]. [17] A. Voros, “ Spectral Functions, Special Functions And Selberg Zeta Function”, Commun. Math. Phys. 110, 439 (1987). [18] R. F. Dashen, B. Hasslacher and A. Neveu, “Particle Spectrum In Model Field Theories From Semiclassical Functional Integral Techniques,” Phys. Rev. D 11, 3424 (1975); “Semiclassical Bound States In An Asymptotically Free Theory,” Phys. Rev. D 12, 2443 (1975). [19] M. Thies, “From relativistic quantum fields to condensed matter and back again: Updating the Gross-Neveu phase diagram,” J. Phys. A 39, 12707 (2006) [arXiv:hep-th/0601049]. [20] L. F. Abbott and H. J. Schnitzer, “Semiclassical Bound State Methods In Four-Dimensional Field Theory: Trace Identities, Mode Sums, And Renormalization For Scalar Theories,” Phys. Rev. D 14, 1977 (1976); J. Avan and H. J. de Vega, “Classical Solutions By Inverse Scattering Transformation In Any Number Of Dimensions. 1. The Gap Equation And The Effective Action,” Phys. Rev. D 29, 2891 (1984). [21] I. M. Gelfand and A. M. Yaglom, “Integration In Functional Spaces And It Applications In Quantum Physics,” J. Math. Phys. 1, 48 (1960). [22] S. Levit and U. Smilansky, “A theorem on infinite products of eigenvalues of Sturm-Liouville type operators”, Proc. Am. Math. Soc. 65, 299 (1977). [23] S. R. Coleman, “The Uses Of Instantons,” in Aspects of Symmetry, (Cambridge University Press, 1985). [24] R. Forman, “ Functional determinants and geometry ”, Invent. Math. 88, 447 (1987); Erratum, ibid 108, 453 (1992). [25] D. Burghelea, L. Friedlander and T. Kappeler, “On the determinant of elliptic differential and finite difference operators in vector bundles over S1”, Commun. Math. Phys. 138, 1 (1991). [26] H. Kleinert, “Path Integrals in , Statistics, Polymer Physics, and Financial Markets,” (World Scientific, Singapore, 2004). [27] B. Simon, “Notes on infinite determinants of Hilbert space operators”, Adv. Math. 24, 244 (1977). [28] K. Kirsten and A. J. McKane, “Functional determinants by contour integration methods,” Annals Phys. 308, 502 (2003) [arXiv:math-ph/0305010]; “Functional determinants for general Sturm-Liouville problems,” J. Phys. A 37, 4649 (2004) [arXiv:math-ph/0403050]. [29] T. Dreyfus and H. Dym, “Product formulas for the eigenvalues of a class of boundary value problems”, Duke Math. J. 45, 15 (1978). [30] G. V. Dunne and K. Kirsten, “Functional determinants for radial operators,” J. Phys. A 39, 11915 (2006) [arXiv:hep-th/0607066]. [31] J. R. Taylor, Scattering Theory, (Wiley, New York, 1972). Functional Determinants in Quantum Field Theory 16

[32] K. Kirsten, Spectral Functions in and Physics, (Chapman-Hall, 2001); K. Kirsten & P. Loya, “Computation of determinants using contour integrals,” arXiv:0707.3755 [hep-th]. [33] J. Baacke and G. Lavrelashvili, “One-loop corrections to the metastable vacuum decay,” Phys. Rev. D 69, 025009 (2004) [arXiv:hep-th/0307202]. [34] A. J. McKane and M. B. Tarlie, “Regularisation of functional determinants using boundary perturbations,” J. Phys. A 28, 6931 (1995) [arXiv:cond-mat/9509126]. [35] G. V. Dunne and H. Min, “Beyond the thin-wall approximation: Precise numerical computation of prefactors in false vacuum decay,” Phys. Rev. D 72, 125004 (2005) [arXiv:hep-th/0511156]. [36] J. S. Langer, “Theory Of The Condensation Point,” Annals Phys. 41, 108 (1967). [37] M. B. Voloshin, I. Y. Kobzarev, and L. B. Okun, “Bubbles In Metastable Vacuum,” Sov. J. Nucl. Phys. 20, 644 (1975) [Yad. Fiz. 20, 1229 (1974)]. [38] M. Stone, “The Lifetime And Decay Of ’Excited Vacuum’ States Of A Field Theory Associated With Nonabsolute Minima Of Its Effective Potential,” Phys. Rev. D 14, 3568 (1976); “Semiclassical Methods For Unstable States,” Phys. Lett. B 67, 186 (1977). [39] S. R. Coleman, “The Fate Of The False Vacuum. 1. Semiclassical Theory,” Phys. Rev. D 15, 2929 (1977) [Erratum-ibid. D 16, 1248 (1977)]; C. G. Callan and S. R. Coleman, “The Fate Of The False Vacuum. 2. First Quantum Corrections,” Phys. Rev. D 16, 1762 (1977). [40] A. W. Wipf, “Tunnel Determinants,” Nucl. Phys. B 269, 24 (1986). [41] D. A. Gorokhov and G. Blatter, “Metastability of (d+n)-dimensional elastic manifolds”, Phys. Rev. B 58, 5486 (1998). [42] R. V. Konoplich and S. G. Rubin, “Quantum Corrections To Nontrivial Classical Solutions In Phi**4 Theory,” Yad. Fiz. 37, 1330 (1983). [43] J. C. Hackworth and E. J. Weinberg, “Oscillating bounce solutions and vacuum tunneling in de Sitter spacetime,” Phys. Rev. D 71, 044014 (2005) [arXiv:hep-th/0410142]. [44] G. V. Dunne and Q. h. Wang, “Fluctuations about cosmological instantons,” Phys. Rev. D 74, 024018 (2006) [arXiv:hep-th/0605176]. [45] V. A. Novikov, M. A. Shifman, A. I. Vainshtein and V. I. Zakharov, “Calculations In External Fields In Quantum Chromodynamics. Technical Review,” Fortsch. Phys. 32, 585 (1984). [46] G. V. Dunne, J. Hur, C. Lee and H. Min, “Instanton determinant with arbitrary quark mass: WKB phase-shift method and derivative expansion,” Phys. Lett. B 600, 302 (2004) [arXiv:hep-th/0407222]; “Precise quark mass dependence of instanton determinant,” Phys. Rev. Lett. 94, 072001 (2005) [arXiv:hep-th/0410190]; “Calculation of QCD instanton determinant with arbitrary mass,” Phys. Rev. D 71, 085019 (2005) [arXiv:hep-th/0502087]. [47] J. Baacke, “Numerical Evaluation Of The One Loop Effective Action In Static Backgrounds With Spherical Symmetry,” Z. Phys. C 47, 263 (1990); “The effective action of a spin 1/2 field in the background of a chiral soliton”, Z. Phys. C 53, 407 (1992). [48] L. Carson, X. Li, L. D. McLerran and R. T. Wang, “Exact Computation Of The Small Fluctuation Determinant Around A Sphaleron,” Phys. Rev. D 42, 2127 (1990). [49] J. Baacke and S. Junker, “Quantum fluctuations around the electroweak sphaleron,” Phys. Rev. D 49, 2055 (1994) [arXiv:hep-ph/9308310]; “Quantum fluctuations of the electroweak sphaleron: Erratum and addendum,” Phys. Rev. D 50, 4227 (1994) [arXiv:hep-th/9402078]. [50] G. Isidori, G. Ridolfi and A. Strumia, “On the metastability of the standard model vacuum,” Nucl. Phys. B 609, 387 (2001) [arXiv:hep-ph/0104016]. [51] Y. Burnier and M. Shaposhnikov, “One-loop fermionic corrections to the instanton transition in two dimensional chiral Higgs model,” Phys. Rev. D 72, 065011 (2005) [arXiv:hep-ph/0507130]. [52] D. Waxman, “The of A Dirac Operator,” Annals Phys. 231, 256 (1994). [53] H. Gies and K. Langfeld, “Quantum diffusion of magnetic fields in a numerical worldline approach,” Nucl. Phys. B 613, 353 (2001) [arXiv:hep-ph/0102185]; “Loops and loop clouds: A numerical approach to the worldline formalism in QED,” Int. J. Mod. Phys. A 17, 966 (2002) [arXiv:hep-ph/0112198].