hep-th0306138/ch8.tex
1: \section{Resummation of the heat kernel expansion}\label{sec8}
2: The heat kernel coefficients define the one-loop counterterms
3: in the background field formalism. In many cases the heat kernel
4: can also give a useful information on the finite part of the
5: effective action. Just one of the examples is the large mass
6: expansion (\ref{largemass}) which is valid when all background
7: fields and their derivatives are small compared to the mass of
8: the quantum field. In order to get the effective action in other
9: limiting cases one has to re-arrange the heat kernel 
10: expansion\footnote{A diagrammatic technique which can be used in resummations
11: of the heat kernel expansion is described in \cite{Moss:1999wq}.}.
12: %%%%%%
13: \subsection{Modified large mass expansion}\label{s8finite}
14: In many physical applications there is a quantity $\mathcal{M}^2$
15: which is large compared to the rest of the background fields
16: and their derivatives. Therefore, it is a well motivated problem
17: to construct an expansion of the effective action in a
18: power series of $\mathcal{M}^{-1}$. To do so, one has to re-express
19: the heat kernel as
20: \begin{equation}
21: K(t;D)=e^{-t\mathcal{M}^2} \sum_k t^{(k-n)/2} \tilde a_k \,.
22: \label{modifHK}
23: \end{equation}
24: To obtain the effective action one has to integrate the heat kernel
25: over $t$ (cf eq.\ (\ref{WKtD})). To simplify the argumentation
26: we assume the cut-off regularization (\ref{WLamreg}), although in other
27: regularization schemes the result will be essentially the same.
28: The divergent and finite parts of the effective action are given
29: by (\ref{divWLam}) and (\ref{largemass}) respectively up to the obvious
30: replacements: $m\to \mathcal{M}$, $(4\pi )^{-n/2}b_k \to \tilde a_k$.
31: After a suitable renormalization, this procedure indeed gives a large
32: $\mathcal{M}$ expansion of the effective action.
33: 
34: Therefore, the problem is reduced to calculation of 
35: $\tilde a_k$. If $\mathcal{M}$ commutes with $D$ the coefficient
36: $\tilde a_k$ are the heat kernel coefficients for
37: the operator $D-\mathcal{M}^2$. This case returns us to the
38: standard large mass expansion. If $\mathcal{M}^2$ does not commute
39: with $D$ then
40: \begin{equation}
41: e^{-tD}\ne e^{-t(D-\mathcal{M}^2)}e^{-t\mathcal{M}^2} \,.\label{noteqexps}
42: \end{equation}
43: To achieve an equality the right hand side of (\ref{noteqexps}) must
44: be corrected by commutator terms. In this case calculation of
45: $\tilde a_k$ requires some amounts of extra work.
46: 
47: Heavy particles of non-equal masses described by the mass matrix
48: $\mathcal{M}^2$ is probably the most immediate example of a physical
49: system to which the modified large mass expansion should be applied.
50: Corresponding technical tools were developed recently 
51: \cite{Osipov:2000rg,Osipov:2001bj,Osipov:2001nx}.
52: 
53: The next example is a scalar field in curved space (cf. sec\ 
54: \ref{s3scalar}). Parker and Toms suggested 
55: \cite{Parker:1985dj,Parker:1985mv} to use the modified large
56: mass expansion with $\mathcal{M}^2=m^2+\xi R$ which partially
57: sums up contributions of the scalar curvature $R$ to the effective
58: action. This formalism was developed further in 
59: \cite{Jack:1985mw,Sushkov:2002rf}.
60: 
61: In some cases the term-by-term integration of the heat kernel
62: expansion gives good estimates of the vacuum energy even if no
63: large parameter is explicitly present in the model (see, e.g.
64: calculations of quantum corrections to the mass of two dimensional
65: solitons \cite{Dunne:1999du,AlonsoIzquierdo:2002eb}). 
66: %%%%%%
67: \subsection{Covariant perturbation theory}\label{s8covpt}
68: Suppose that the matrix potential $E$, the Riemann curvature,
69: and the field strength $\Omega$ are small but rapidly varying.
70: Can we get any information on the heat kernel coefficients
71: containing a fixed power of the quantities listed above and
72: {\it arbitrary} number of derivatives? For the linear and
73: quadratic orders on a manifold without boundaries the answer
74: is positive and the results can be obtained either by the
75: functorial methods of sec.\ \ref{s4gen} 
76: \cite{Gilkey:1988,Branson:1990b}
77: or by solving the DeWitt equation 
78: \cite{Avramidi:1989er,Avramidi:1990ug,Avramidi:1991je}
79: (cf. sec.\ \ref{s4iter}). Retaining only the leading terms
80: we have for $k\ge 1$:
81: \begin{eqnarray}
82: &&a_{2k}(x,D) =(4\pi )^{-n/2} \left[ \alpha_1(k) 
83: \Delta^{k-1} R +\alpha_2 (k) \Delta^{k-1} E \right.\label{leadhk}\\
84: &&\qquad\qquad\qquad\qquad \left. 
85: +\mbox{higher order terms} \right]\,, \nonumber
86: \end{eqnarray}
87: where \cite{Gilkey:1979,Gilkey:1988}
88: \begin{equation}
89: \alpha_1(k)=\frac{k!k}{(2k+1)!} \,,\qquad
90: \alpha_2(k)=\frac{2k!}{(2k)!}\,.\label{al12k}
91: \end{equation}
92: 
93: One can sum up the expansion (\ref{leadhk}) and corresponding higher
94: order terms to obtain an information on the behaviour of the full
95: heat kernel in the limit described above. There is, however, a more
96: straightforward method to control non-localities which is called
97: the covariant perturbation theory 
98: \cite{Barvinsky:1987uw,Barvinsky:1990up,Barvinsky:1990uq}. To make the
99: idea of this method most transparent we consider a simple example
100: of flat $M$ and zero connection $\omega$. The exponent
101: $\exp (-tD)=\exp (t(\Delta + E))$ with $\Delta =\partial_\mu^2$
102: can be expanded in a power series
103: in $E$:
104: \begin{equation}
105: e^{-tD}=e^{t\Delta} + \int\limits_0^t ds e^{(t-s)\Delta} E e^{s\Delta}
106: +\int\limits_0^t ds_2 \int\limits_0^{s_2} ds_1
107: e^{(t-s_2)\Delta}E e^{(s_2-s_1)\Delta} E e^{s_1\Delta} +\dots
108: \label{pertex}
109: \end{equation}
110: The heat trace can be also expanded,
111: \begin{equation}
112: K(t,D)={\rm Tr}\left( e^{-tD}\right)=
113: \sum_{j=0}^\infty K_j(t),\label{pexhtr}
114: \end{equation}
115: where $K_j$ contains the $j$th power of $E$.
116: 
117: Covariant perturbation theory approach prescribes to take the
118: $0$th order heat kernel in the free space form (cf. eq.\ 
119: (\ref{simplestHK}) for $m=0$):
120: \begin{equation}
121: K_0(x,y;t)=(4\pi t)^{-n/2} \exp
122: \left( -\frac{(x-y)^2}{4t} \right) \,,\label{Kzero}
123: \end{equation}
124: which is an exact kernel on $M=\mathbb{R}^n$ only\footnote{The
125: original paper \cite{Barvinsky:1987uw} contained a curved space
126: generalisation of (\ref{Kzero}). This, however, does not improve
127: the global issues discussed below.}. This formula neglects all
128: global contributions and, therefore, is valid only for sufficiently
129: close $x$ and $y$ and for small $t$. We have:
130: \begin{equation}
131: K_0(t)=\iM {\rm tr}_V K_0(x,x;t)=(4\pi t)^{-n/2} \iM {\rm tr}_V (I_V)
132: \,,\label{trKzero}
133: \end{equation}
134: where $\ptr I_V$ simply counts discrete (spin and internal) 
135: indices of $D$. This formula reproduces the $a_0$ contribution
136: to the heat kernel.
137: 
138: In the next order, we have:
139: \begin{eqnarray}
140: K_1(t)=&&{\rm Tr} \left( \int_0^t ds e^{(t-s)\Delta}Ee^{s\Delta}
141: \right)=
142: {\rm Tr}\left( \int_0^t ds e^{t \Delta} E \right) \nonumber \\
143: =&& t\, \ptr \iM K_0 (x,x;t) E(x) \nonumber\\
144: =&&\frac t{(4\pi t)^{n/2}}
145: \iM \ptr E(x) \,,\label{Kone}
146: \end{eqnarray}
147: where we used the cyclic property of the trace and the expression
148: (\ref{Kzero}) for the $0$th order heat kernel. This expression is
149: consistent with the $k=1$ term of (\ref{leadhk}). The terms with
150: $k>1$ are total derivatives, and, therefore, they do not contribute
151: to the integrated heat kernel. 
152: 
153: The quadratic order of the heat kernel reads
154: \begin{eqnarray}
155: K_2(t)=&&{\rm Tr} \left( \int\limits_0^t ds_2 \int\limits_0^{s_2} ds_1
156: e^{(t-s_2)\Delta}E e^{(s_2-s_1)\Delta} E e^{s_1\Delta} \right) \nonumber \\
157: =&&\ptr \int_M dy \int_M dz 
158: \int\limits_0^t ds_2 \int\limits_0^{s_2} ds_1 
159: K_0 (z,y;t-s_2+s_1)E(y)\nonumber\\
160: &&\qquad\qquad\qquad\qquad\qquad \times K_0(y,z;s_2-s_1)E(z) \,.\label{K2a}
161: \end{eqnarray}
162: Next we introduce the rescaled variables $\xi =s/t$ and get rid of redundant
163: integrations to obtain:
164: \begin{eqnarray}
165: K_2(t)=&&\frac {t^2}2 \ptr \int_M dy \int_M dz \int\limits_0^1 d\xi
166: K_0(z,y;t(1-\xi )) E(z) \nonumber\\
167: &&\qquad\qquad\qquad\qquad\qquad 
168: \times K_0(y,z;t\xi)E(z) \,.\label{K2b}
169: \end{eqnarray}
170: Now we use the identity
171: \begin{equation}
172: K_0(z,y;t(1-\xi )) K_0(y,z;t\xi)=(4\pi t)^{-n/2}
173: K_0(z,y;t\xi(1-\xi)) \label{KKKide}
174: \end{equation}
175: and relate the heat kernel on the right hand side of (\ref{KKKide})
176: to a matrix element of $\exp (t\xi(1-\xi )\Delta )$ to obtain
177: the final result:
178: \begin{equation}
179: K_2(t)=\frac{t^2}{(4\pi t)^{n/2}}
180: \iM E f(-t\Delta ) E \,,\label{Ktwo}
181: \end{equation}
182: where the non-local form-factor $f$ reads:
183: \begin{equation}
184: f(q)=\frac 12 \int\limits_0^1 d\xi e^{-q\xi (1-\xi)}=
185: \frac 12 e^{-q/4} \sqrt{\pi /q} {\rm Erfi} \left[ \sqrt{q}/2 \right]\,.
186: \end{equation}
187: As we have already discussed above, applicability of this formula is
188: limited by our choice of $K_0(x,y;t)$. Namely, the potential $E(x)$
189: should have a compact support and $t$ should be reasonably small.
190: Note, that the small $t$ approximation in the expansion (\ref{pertex})
191: is self-consistent: if $t$ is small, the integration variables $s_i$
192: are even smaller.
193: 
194: The main difficulty in constructing an expansion in powers
195: the Riemann curvature and of the field strength is to organise
196: the procedure in a {\it covariant} way. The details of the
197: construction and higher order form-factors can be found in  
198: \cite{Barvinsky:1987uw,Barvinsky:1990up,Barvinsky:1990uq,Barvinsky:1994hw,
199: Barvinsky:1994ic}. From further developments of this method
200: we mention the work of Gusev and Zelnikov \cite{Gusev:1999cv}
201: who demonstrated that in two dimension one can achieve considerable
202: simplifications in the perturbation expansion by using the dilaton
203: parametrisation of the potential (cf. eq.\ (\ref{ehnEh})).
204: Recently Barvinsky and Mukhanov \cite{Barvinsky:2002uf} suggested
205: a new method for calculation of the non-local part of the effective
206: action based on the resummation of the perturbation series for the
207: heat kernel. This method was extended in \cite{Barvinsky:2003rx} to
208: include late time asymptotics of the heat kernel in curved space.
209: 
210: Let us stress that the expansion (\ref{pertex}) can be used also
211: for singular potentials. For example, it is very effective for 
212: the calculation of the heat kernel for $\delta$-potentials
213: concentrated on a co-dimension one subsurface 
214: \cite{Gaveau:1986,Bordag:1999ed,Moss:2000gv} (cf.\ sec.\ \ref{s6dom}).
215: If the $\delta$-potential has its support on a submanifold
216: of co-dimension greater than one, the expansion diverges 
217: \cite{Bordag:1999ed}. 
218: With a suitable choice of the zeroth order heat kernel $K_0(x,y;t)$
219: and of an operator to replace $E$ in (\ref{pertex}) one can treat
220: manifolds with boundaries \cite{Bordag:2001fj} where the perturbative
221: expansion takes the form of the multiple reflection expansion
222: \cite{Balian:1970,Balian:1977za,Hansson:1983xt}.
223: %%%%%%
224: \subsection{``Low energy'' expansion}\label{s8low}
225: Let us now turn to the opposite case when the derivatives are less
226: important when the potential and the curvatures. In this case
227: one has to collect the terms which are of a fixed order in the
228: derivatives, but contain all powers of $E$, $R$ and $\Omega$.
229: Since the derivatives are sometimes identified with the energy,
230: this approximation is being called the low energy expansion.
231: This scheme goes back to Schwinger's calculations
232: \cite{Schwinger:1951} in constant electromagnetic 
233: fields\footnote{We like to mention also the paper
234: \cite{Heisenberg:1936} which treated the effective action
235: in external electromagnetic field from a different point view.}.
236: 
237: Let us consider a simple example \cite{Blau:1991iz}. Let
238: $M=\mathbb{R}^2$. Consider a scalar particle in constant
239: electromagnetic field with the field strength $F_{12}=-F_{21}=B$.
240: As a potential we choose $A_1=0$, $A_2=Bx^1$. Then the operator
241: acting on quantum fluctuations is
242: \begin{equation}
243: D=-\partial_1^2-(\partial_2 -iBx^1)^2 +m^2 \,.\label{cmagD}
244: \end{equation}
245: This operator commutes with $\partial_2$. Therefore, we can look
246: for the eigenfunctions of $D$ in the form 
247: $\phi_k (x)=\exp (ikx^2) \tilde\phi (x^1)$. 
248: \begin{equation}
249: D\phi_k (x) = \left( -\partial_1^2 +B^2 (x^1 - k/B)^2 +m^2 \right)
250: \phi_k (x) \,.\label{cmDk}
251: \end{equation}
252: In the $x^1$ direction we have the one-dimensional harmonic
253: oscillator potential (cf. (\ref{harmos})). Therefore, the 
254: eigenvalues are $\lambda_{k,p}= (2p+1)|B|+m^2$. These eigenvalues
255: do not depend on $k$. For this reason the heat kernel
256: \begin{equation}
257: K(t,D)=\sum_{k,p} \exp (-t\lambda_{k,p}) \label{cmagKt}
258: \end{equation}
259: is ill defined. To overcome this difficulty it was suggested 
260: \cite{Blau:1991iz} to put the system in a box (without specifying
261: any boundary conditions, however) and replace the sum over $k$
262: by the degeneracy factor $(|B| {\rm Vol})/(2\pi )$, where ${\rm Vol}$
263: is volume of the box. The degeneracy factor is chosen in such a way
264: that the resulting heat kernel
265: \begin{equation}
266: K(t,D)= \frac{B{\rm Vol}}{4\pi } e^{-m^2t} [{\rm sinh} (Bt)]^{-1}
267: \label{cmagKB}
268: \end{equation}
269: behaves as ${\rm Vol}/(4\pi t)$ in the limit $B\to 0$.
270: 
271: Calculations of the effective action in covariantly constant
272: background gauge fields by the spectral theory methods have 
273: been performed by many authors 
274: \cite{Batalin:1977uv,Dittrich:1983ej,Elizalde:1984yb,
275: Elizalde:1985zv,Blau:1991iz}. Such calculations were motivated, at least
276: partially, by various models of quark confinement (see 
277: \cite{Pagels:1978dd,Adler:1982rk} and references therein).
278: 
279: There exists an algebraic method \cite{Avramidi:1993yf,Avramidi:1995ik} 
280: which allows to evaluate the low energy heat kernel by using 
281: exclusively the commutator algebra. Let us briefly formulate the
282: results of \cite{Avramidi:1993yf,Avramidi:1995ik}. Consider a flat
283: manifold ($R_{\mu\nu\rho\sigma}=0$) with the background fields
284: satisfying the ``low-energy conditions'':
285: \begin{equation}
286: \nabla_\mu \Omega_{\mu\nu}=0,\qquad
287: \nabla_\mu\nabla_\nu\nabla_\rho E=0 
288: \label{lowecon}
289: \end{equation}
290: with usual definitions of $E$ and $\Omega$ (see (\ref{D1}) - 
291: (\ref{Omega})). Moreover, let us suppose that the background
292: is ``approximately abelian'', i.e. that $\Omega_{\mu\nu}$,
293: $E$ and all their covariant derivatives are mutually commuting matrices. 
294: 
295: With all these assumptions a closed expression for the heat kernel
296: may be obtained:
297: \begin{equation}
298: K(t;x,x;D)=(4\pi t)^{-n/2} \exp \left(
299: tE +\Phi (t) +\frac {t^3}4 E_{;\mu} \Psi^{\mu\nu} E_{;\nu} \right)\,,
300: \label{avrahk}
301: \end{equation}
302: where $\Phi$ and $\Psi$ are complicated functions of $t$, $E$ and $\Omega$
303: \cite{Avramidi:1995ik}. If $\nabla_\mu\nabla_\nu E=0$,
304: \begin{eqnarray}
305: &&\Phi (t)=-\frac 12 \ln \det \left( \frac{\sinh t \Omega}{t\Omega}
306: \right) \,,\nonumber\\
307: &&\Psi (t) = (t\Omega )^{-2} \left( t\Omega \coth (t\Omega )-1
308: \right) \,,\label{PhiPsi1}
309: \end{eqnarray}
310: where $\Omega$ has to be understood as a space-time matrix
311: $\Omega_{\mu\nu}$, so that multiplication in (\ref{PhiPsi1})
312: is the matrix multiplication, and $\det$ is the determinant
313: of an $n\times n$ matrix. These formulae generalise the equation
314: (\ref{cmagKB}) and justify the choice of the degeneracy factor made
315: to derive it.
316: 
317: If $\Omega =0$,
318: \begin{eqnarray}
319: &&\Phi (t) = -\frac 12 \ln \det \left( 
320: \frac{\sinh (2t\sqrt{P}}{2t\sqrt{P}} \right) \,,\nonumber\\
321: &&\Psi (t) = -\left( t\sqrt{P} \right)^{-3}
322: \left( \tanh (t\sqrt{P}) -t\sqrt{P}) \right) \,,\label{PhiPsi2}
323: \end{eqnarray}
324: where $P_{\mu\nu}=-(1/2)\nabla_\mu\nabla_\nu E$. In the particular
325: case of one-dimensional harmonic oscillator these formulae
326: reproduce (\ref{harmheat}). 
327: 
328: In curved space the best one can do in the framework of the
329: low-energy expansion is to consider
330: locally symmetric manifolds, i.e. vanishing $\Omega$,
331: covariantly constant Riemann curvature and constant $E$. 
332: In this case formulae similar to the ones presented above
333: are available \cite{Avramidi:1994zc,Avramidi:1996zp}.
334: Covariantly constant curvature means that locally 
335: the manifold $M$ is a symmetric space. Various approaches
336: to the heat kernel on such manifolds are described in detail
337: in monographs and survey articles 
338: \cite{Hurt:1983,Camporesi:1990wm,Bytsenko:2001xs}. In particular,
339: very detailed information may be obtained for group manifolds
340: (see, for example, \cite{Dowker:1971vu}) and for hyperbolic
341: spaces \cite{Bytsenko:1996bc}.
342: %%%%%%
343: \subsection{Heat kernel on homogeneous spaces}\label{s8homo}
344: In this section we briefly explain how one can find
345: the spectrum of some ``natural'' differential
346: operators on homogeneous spaces
347: by purely algebraic methods. We start with
348: some basic facts from differential geometry and harmonic analysis
349: \cite{Kobayashi:1969,Helgason:1984}. 
350: Consider a homogeneous space $G/H$ of
351: two compact finite-dimensional Lie groups $G$ and $H$.
352: The Lie algebra $\mathcal{G}$ of $G$ can be decomposed as
353: \begin{equation}
354: \mathcal{G}=\mathcal{H}\oplus \mathcal{M} \,,\label{Gdecom}
355: \end{equation}
356: where $\mathcal{H}$ is the Lie algebra of $H$ and $\mathcal{M}$
357: is the complement of $\mathcal{H}$ in $\mathcal{G}$ with
358: respect to some bi-invariant metric. We have:
359: \begin{equation}
360: [\mathcal{H},\mathcal{M}] \subset \mathcal{M} \,,\label{HMM}
361: \end{equation}
362: where $[\ ,\ ]$ is the Lie bracket on $\mathcal{G}$. If, moreover,
363: $[\mathcal{M},\mathcal{M}]\subset \mathcal{H}$, then $G/H$ is a
364: symmetric space. We do not impose this restriction.
365: 
366: $\mathcal{M}$ can be identified with the tangent space to $M=G/H$
367: at the origin (i.e. at the point which represents the unit element of $G$).
368: Eq.\ (\ref{HMM}) tells us that $\mathcal{H}$ acts on $\mathcal{M}$
369: by some (orthogonal) representation. This action defines the embedding
370: \begin{equation}
371: \mathcal{H}\subset so(n) \,.\label{Hinso}
372: \end{equation}
373: All physical fields are classified according to certain representations
374: of the Lie algebra $so(n)$. Restrictions of these representations to
375: $\mathcal{H}$ define transformation properties of the field with
376: respect to $\mathcal{H}$. From now on we work with each irreducible
377: representation of $\mathcal{H}$ separately.
378: 
379: The field $\Phi^A$ belonging to
380: an irreducible representation $\mathcal{T}(H)$ can be expanded as 
381: (see, e.g., \cite{Salam:1982xd}):
382: \begin{equation}
383:  \Phi_A (x)=({\rm Vol})^{-\frac 12} \sum_{j, \xi ,q}
384: \sqrt {\frac {d_j}{d_{\mathcal{T}}}} \mathcal{T}^{(j)}_{A\xi ,q}
385: (g_x^{-1}) \phi^{(j)}_{q,\xi }\ , \label{eq:(A.2)}
386: \end{equation}
387: where ${\rm Vol}$ is the volume of $G/H$, 
388: $d_{\mathcal{T}}={\rm dim}\mathcal{T}(H)$.
389: We sum over the representations $\mathcal{T}^{(j)}$ of $G$
390: which give $\mathcal{T}(H)$ after reduction to $H$. $\xi$ labels
391: multiple components  $\mathcal{T}(H)$ in the branching
392: $\mathcal{T}^{(j)}\downarrow H$, $d_j=\dim \mathcal{T}^{(j)}$. 
393: $q$ runs from $1$ through $d_j$. The matrix
394: elements of $\mathcal{T}^{(j)}$ have the following orthogonality
395: property
396: \begin{equation} \iM
397: \mathcal{T}^{(j) \dag}_{A\zeta ,q} (g_x^{-1})
398: \mathcal{T}^{(j')}_{A\xi ,p} (g_x^{-1})
399: =({\rm Vol})d_j^{-1} d_{\mathcal{T}} \delta_{\zeta \xi} \delta_{p q}
400: \delta_{jj'} \label{eq:(A.3)}\end{equation}
401: Therefore, to construct the harmonic expansion on $G/H$ it is
402: necessary to have powerful methods for reduction of the representations 
403: from $G$ to $H$. There are several standard 
404: \cite{Barut:1977,Slansky:1981yr}
405: and less standard  \cite{Lyakhovsky:1989cr,Lyakhovsky:1991vv}
406: techniques which may be used depending on the particular
407: homogeneous space.
408: 
409: It is important that not only the harmonic expansion but also
410: the spectrum of the invariant operators can be analysed by the group
411: theoretical methods. The covariant derivative on $G/H$
412: reads \cite{Kobayashi:1969}:
413: \begin{equation}
414: \nabla_\mu = \nabla_\mu^{[c]}+\Gamma^{[R]}_\mu \,.\label{hscovdiv}
415: \end{equation}
416: Here $\nabla^{[c]}$ is the canonical covariant derivative on
417: $G/H$. At the origin $\nabla^{[c]}$ can be identified with 
418: the tangent space generators from $\mathcal{M}$ taken in
419: the representation $\mathcal{T}^{(j)}$. The part $\Gamma^{[R]}_\mu$
420: depends on the invariant metric on $G/H$ and on the
421: structure constants of $\mathcal{G}$ restricted to $\mathcal{M}$.
422: On symmetric spaces such structure constants are zero and,
423: therefore, the Laplace-Beltrami operator has a particularly
424: simple form:
425: \begin{equation}
426: D\simeq \nabla^{\mu[c]}\nabla_\mu^{[c]} \simeq C_2 (G)-C_2 (H)\,,
427: \label{LapCas}
428: \end{equation}
429: where $C_2$ are quadratic Casimir operators of $G$ and $H$ which
430: depend on the representations $\mathcal{T}^{(i)}$ and 
431: $\mathcal{T}(H)$ respectively.
432: On general homogeneous spaces the expressions are a bit more
433: complicated (see, e.g., 
434: \cite{Vassilevich:1988em,Lyakhovsky:1991vv} for explicit
435: examples). In any case, eigenvalues of $D$ are given by
436: a second order polynomial $Q(m_1,\dots,m_k)$ of several
437: natural numbers $m_l$. These eigenvalues are in general
438: degenerate with multiplicities defined essentially 
439: by dimensions of the representations of $G$ and $H$.
440: They are also polynomials\footnote{The spectrum of differential 
441: operators on coset spaces can be calculated exactly even if one adds
442: a homogeneous gauge field (see \cite{Dolan:2003bj} and references 
443: therein).}
444:  in $m_l$. The heat kernel is then
445: represented as an infinite sum
446: \begin{equation}
447: K(t,D)=\sum_{m_l} N(m_1,\dots,m_k) \exp (-t Q(m_1,\dots,m_k))\,.
448: \label{hsheatk}
449: \end{equation}
450: There exist several tricks which can be used to evaluate
451: the $t\to 0$ asymptotics of such sums. For example, one may use
452: the Poisson summation formula (\ref{eq:asPoi}). By taking derivatives
453: with respect to $t$ one obtains 
454: \begin{equation}
455: \sum_{l\in \mathbb{Z}} l^{2r} \exp \left( -tl^2\right)
456: \simeq \sqrt{\pi} 2^{-r} (2r-1)!! t^{-(2r+1)/2} +
457: \mathcal{O}\left( e^{-1/t} \right)\label{diffPoi}
458: \end{equation}
459: for $r\in\mathbb{N}$.
460: 
461: More general polynomials of $l$ may be treated by the 
462: Euler--Maclaurin formula which reduces sums to the integrals.
463: Let $F(\tau )$ be a function defined on $0 \le \tau < \infty$. If the
464: $2m$-th derivative
465: $F^{(2m)}(\tau)$ is absolutely integrable on $(0,\infty )$ 
466: \begin{eqnarray}
467: &&\sum_{i=0}^k F(i) -\int_0^\infty F(\tau )d\tau =\frac 12 [F(0)+F(k)]
468: \nonumber \\
469: &&\qquad\qquad\qquad +\sum_{s=1}^{m-1} \frac {B_{2s}}{(2s)!}
470: [F^{(2s-1)}(k)-F^{(2s-1)}(0)]+{\rm Rem}_m(n), \label{eq:asEM}
471: \end{eqnarray}
472: where the reminder ${\rm Rem}_m$ satisfies:
473: \begin{equation}
474: |{\rm Rem}_m(n)| \le (2-2^{1-2m}) \frac {|B_{2m}|}{(2m)!}
475: \int_0^n |F^{(2m)}(\tau )| d\tau .
476: \label{eq:asrem}
477: \end{equation}
478: $B_s$ are the Bernoulli numbers. If $F(i)$ is taken to be
479: $N_i \exp (-t \lambda_i )$, one can take limit $k\to \infty$
480: in (\ref{eq:asEM}) and restrict the summation to some finite
481: $m$ to obtain asymptotic series for the heat kernel.
482: For example, one can easily recover the expansion (\ref{diffPoi}).
483: 
484: In many cases calculation can be done by means of the Mellin transform.
485: This method will be clear from the following example.
486: Consider
487: \begin{equation}
488: K(t)=\sum_{k=1}^\infty k\exp (-tk^2) \label{(A.1)}
489: \end{equation}
490: The Mellin transform of $K(t)$ is
491: \begin{equation}
492: \tilde K(s)=\sum_k k \int_0^\infty dx\ x^{s-1}
493: \exp (-xk^2) =\sum_{k=1}^\infty k^{1-2s}\Gamma (s)
494: =\zeta_R(2s-1) \Gamma (s), \label{(A.2)}
495: \end{equation}
496: where $\zeta_R$ is the Riemann zeta-function. Inverse
497: Mellin transformation gives
498: \begin{equation}
499: K(t)=\frac 1{2\pi i} \int_C t^{-s} \zeta_R(2s-1)
500: \Gamma (s) ds. \label{(A.3)}
501: \end{equation}
502: The contour $C$ covers all poles of the integrand
503: at $s=1, 0, -1, -2,\dots$. By calculating the residues we
504: obtain the desired expansion:
505: \begin{equation}
506: K(t)=\frac 1{2t} -\frac 1{12} -\frac t{120}
507: -\frac {t^2}{504} -\frac {t^3}{1440} +O(t^4).
508: \end{equation}
509: 
510: Actually, the techniques introduced above should only be
511: used for complicated multi-parameter sums (see, e.g.,
512: \cite{Lyakhovsky:1991vv}).
513: For a simple one-parameter sum it is easier to
514: use combinations of known asymptotic series (cf. Appendix
515: of Ref. \cite{Birmingham:1987cy}).
516: 
517: In this method complexity of calculations of the heat kernel
518: coefficients $a_k$ is almost independent of $k$. Therefore,
519: the algebraic techniques were applied to higher dimensional theories
520: where one needs higher heat kernel coefficients to
521: perform renormalization or to calculate the anomalies.
522: Rather naturally, the most simple toroidal spaces were
523: considered first \cite{Appelquist:1983zs,Appelquist:1983vs,Rubin:1983zz,
524: Kogan:1983fp,Fradkin:1983kf,Buchbinder:1988np}, and then
525: computations on spheres were performed
526: \cite{Candelas:1984ae,Kikkawa:1985qc,Kantowski:1987ct,Birmingham:1987cy,
527: Birmingham:1988iv,Gleiser:1987mg,BuchOd,Buchbinder:1989cd,Elizalde:1996nb}.
528: Other homogeneous spaces were considered for example in 
529: \cite{Lyakhovsky:1991vv,Vassilevich:1991zi}. In the same approach
530: non-minimal operators on homogeneous spaces were treated
531: in \cite{Vassilevich:1992tb,Vassilevich:1993yt}. We refer to
532: \cite{Camporesi:1990wm} for a more extensive literature survey.
533: