ncg

bachelorthesis in physics
git clone git://popovic.xyz/ncg.git
Log | Files | Refs

4_heatkernel.tex (13630B)


      1 \subsection{Heat Kernel Expansion\label{sec:4}}
      2 \subsubsection{The Heat Kernel}
      3 The heat kernel $K(t; x, y; D)$ is the fundamental solution to the heat
      4 equation
      5 \begin{align}
      6     (\partial _t + D_x)K(t;x, y;D) =0,
      7 \end{align}
      8 which depends on the operator $D$ of Laplacian type.
      9 
     10 For a flat manifold $M = \mathbb{R}^n$ and $D = D_0 := -\Delta_\mu\Delta^\mu +m^2$ the
     11 Laplacian with a mass term and the initial condition
     12 \begin{align}
     13     K(0;x,y;D) = \delta(x,y),
     14 \end{align}
     15 takes the form of the standard fundamental solution
     16 \begin{align}\label{eq:standard}
     17     K(t;x,y;D_0) = (4\pi t)^{-n/2}\exp\left(-\frac{(x-y)^2}{4t}-tm^2\right).
     18 \end{align}
     19 
     20 Let us consider now a more general operator $D$ with a potential term or a
     21 gauge field, the heat kernel then reads
     22 \begin{align}
     23     K(t;x,y;D) = \langle x|e^{-tD}|y\rangle.
     24 \end{align}
     25 We can expand the heat kernel in $t$, still having a
     26 singularity from the equation \eqref{eq:standard} as $t \rightarrow 0$, on
     27 obtains
     28 \begin{align}
     29     K(t;x,y;D) = K(t;x,y;D_0)\left(1 + tb_2(x,y) + t^2b_4(x,y) + \dots
     30     \right),
     31 \end{align}
     32 where $b_k(x,y)$ become regular as $y \rightarrow x$. These coefficients are called the heat
     33 kernel coefficients.
     34 %%----------------------- KANN WEGGELASSEN WERDEN
     35 %\newline
     36 %\textbf{KANN WEGELASSEN WERDEN BIS ZUM NÄCHSTEN KAPITEL}
     37 %Let's turn our attention to a propagator $D^{-1}(x,y)$ defined through the
     38 %heat kernel, with an integral representation
     39 %\begin{align}
     40 %    D^{-1} (x,y) = \int_0^\infty dt K(t;x,y;D).
     41 %\end{align}
     42 %If we assume the heat kernel vanishes for $t\rightarrow \infty$, we can
     43 %integrate formally to get
     44 %\begin{align}
     45 %    D^{-1}(x,y) \simeq
     46 %    2(4\pi)^{-n/2}\sum_{j=0}\left(\frac{|x-y|}{2m}\right)^{-\frac{n}{2}+j+1}
     47 %    K_{-\frac{n}{2}+j+1}(|x-y|m)b_{2j}(x,y),
     48 %\end{align}
     49 %where $b_0 = 1$ and $K_\nu (z)$ is the Bessel function
     50 %\begin{align}
     51 %    K_\nu(z) = \frac{1}{\pi} \int_0^\pi \cos(\nu\tau-z\sin(\tau))d\tau.
     52 %\end{align}
     53 %The Bessel function solves the following differential equation
     54 %\begin{align}
     55 %    z^2 \frac{d^2K}{dz^2} + z \frac{dK}{dz} + (z^2 - \nu^2)=0.
     56 %\end{align}
     57 %By looking at an integral approximation for the propagator we conclude that
     58 %the singularities of $D^{-1}$ coincide with the singularities of the heat
     59 %kernel coefficients. Thus we can say, that a generating functional in terms of
     60 %$\det(D)$ is called the one-loop effective action (quantum field theory)
     61 %\begin{align}
     62 %    W = \frac{1}{2}\ln(\det D).
     63 %\end{align}
     64 %We have a direct relation with one-loop effective action $W$ and the
     65 %heat kernel. Furthermore notice that for each eigenvalue $\lambda >0$ of $D$
     66 %we can write the identity.
     67 %\begin{align}
     68 %    \ln \lambda  = -\int_0^\infty \frac{e^{-t\lambda}}{t}dt
     69 %\end{align}
     70 %This expression is correct up to an infinite constant which does not depend
     71 %on the eigenvalue $\lambda$, thus we can ignore it. By substituting
     72 %$\ln(\det D) = \text{Tr}(\ln D)$ we can rewrite the one-loop effective action
     73 %$W$ into
     74 %\begin{align}
     75 %    W = -\frac{1}{2} \int_0^\infty dt \frac{K(t, D)}{t},
     76 %\end{align}
     77 %where
     78 %\begin{align}
     79 %    K(t, D) = \text{Tr}(e^{-tD}) = \int d^n x \sqrt{g}K(t;x,x;D).
     80 %\end{align}
     81 %The problem now is that the integral of $W$ is divergent at both limits. Yet
     82 %the divergences at $t\rightarrow \infty$ are caused by $\lambda \leq 0$ of $D$
     83 %(infrared divergences) and can be ignored. The divergences at $t\rightarrow 0$
     84 %are cutoff at $t=\Lambda^{-2}$, simply written as
     85 %\begin{align}
     86 %    W_\Lambda = -\frac{1}{2} \int_{\Lambda^{-2}}^\infty dt \frac{K(t, D)}{t}.
     87 %\end{align}
     88 %We can calculate $W_\Lambda$ up to an order of $\lambda ^0$
     89 %\begin{align}
     90 %    W_\Lambda &= -(4\pi)^{-n/2} \int d^n x\sqrt{g}\bigg(
     91 %    \sum_{2(j+l)<n}\Lambda^{n-2j-2l}b_{2j}(x,x) \frac{(-m^2)^l l!}{n-2j-2l} +\\
     92 %    &+ \sum_{2(j+l) =n }\ln(\Lambda) (-m^2)^l l! b_{2j}(x,x)
     93 %    \mathcal{O}(\lambda^0) \bigg)
     94 %\end{align}
     95 %There is an divergence at $b_2(x,x)$ for $k\leq n$. Computing the limit
     96 %$\Lambda \rightarrow \infty$ we get
     97 %\begin{align}
     98 %    -\frac{1}{2}(4\pi)^{n/2}m^n\int d^n x\sqrt{g} \sum_{2j>n}
     99 %    \frac{b_{2j}(x,x)}{m^{2j}}\Gamma(2j-n),
    100 %\end{align}
    101 %where $\Gamma$ stands for the gamma function.
    102 %%----------------------- KANN WEGGELASSEN WERDEN
    103 
    104 
    105 \subsubsection{Spectral Functions}
    106 Manifolds $M$ with a disappearing boundary condition for the operator
    107 $e^{-tD}$ for $t>0$, i.e.\ a trace class operator on $L^2(V)$. Meaning for any
    108 smooth function $f$ on $M$ the Heat kernel can be defined as
    109 \begin{align}
    110     K(t,f,D) := \text{Tr}_{L^2}(fe^{-tD}).
    111 \end{align}
    112 Alternately an integral representation is
    113 \begin{align}
    114     K(t, f, D) = \int_M d^n x \sqrt{g} \text{Tr}_V(K(t;x,x;D)f(x)),
    115 \end{align}
    116 in the regular limit $y \rightarrow y$. The Heat Kernel can be written in terms
    117 of the spectrum of $D$. Hence for an orthonormal basis $\{\phi_\lambda\}$ of
    118 eigenfunctions for $D$, which correspond to the eigenvalue $\lambda$ one
    119 can rewrite the heat kernel into
    120 \begin{align}
    121     K(t;x,y;D) = \sum_\lambda \phi^\dagger_\lambda(x)
    122     \phi_\lambda(y)e^{-t\lambda}.
    123 \end{align}
    124 An asymptotic expansion as $t \rightarrow 0$ for the trace is then
    125 \begin{align}
    126     \text{Tr}_{L^2}(fe^{-tD}) \simeq \sum_{k\geq 0}t^{(k-n)/2}a_k(f,D),
    127 \end{align}
    128 where
    129 \begin{align}
    130     a_k(f,D) = (4\pi)^{-n/2} \int_M d^4x \sqrt{g} b_k(x,x) f(x).
    131 \end{align}
    132 \subsubsection{General Formulae}
    133 Let us summarize what we have obtained in the last chapter. We considered a
    134 compact Riemannian manifold $M$ without boundary condition, a vector bundle
    135 $V$ over $M$ to define functions which carry discrete (spin or gauge)
    136 indices, an operator $D$ of Laplace type over $V$ and smooth function $f$ on
    137 $M$.
    138 
    139 There is an asymptotic expansion where the heat kernel coefficients with an
    140 odd index $k=2j+1$ vanish $a_{2j+1}(f,D) = 0$. On the other hand coefficients
    141 with an even index are locally computable in terms of geometric invariants
    142 \begin{align}
    143     a_k(f,D) &= \text{Tr}_V\left(\int_M d^n x\sqrt{g}(f(x)a_k(x;D)\right)
    144     =\nonumber\\
    145     &=\sum_I \text{Tr}_V\left(\int_M d^nx \sqrt{g}(fu^I
    146     \mathcal{A}^I_k(D))\right).
    147 \end{align}
    148 The notation $\mathcal{A}^I_k$ corresponds to all possible independent invariants of
    149 dimension $k$ and $u^I$ are constants. The invariants are constructed from
    150 $E, \Omega, R_{\mu\nu\varrho\sigma}$ and their derivatives. If $E$ has
    151 dimension two, then the derivative has dimension one. In this way if $k=2$ there are
    152 only two independent invariants, $E$ and $R$. This corresponds to the
    153 statement $a_{2j+1}=0$.
    154 
    155 If we consider $M = M_1 \times M_2$ with coordinates $x_1$ and $x_2$ and a
    156 decomposed Laplace style operator $D = D_1 \otimes 1 + 1 \otimes D_2$ the
    157 functions acting on operators and on coordinates are separable linearly by the
    158 following
    159 \begin{align}
    160     e^{-tD} &= e^{-tD_1} \otimes e^{-tD_2},\\
    161     f(x_1, x_2) &= f_1(x_1)f_2(x_2),
    162 \end{align}
    163 thus the heat kernel coefficients are separated by
    164 \begin{align}
    165     a_k(x;D) &= \sum_{p+q=k} a_p(x_1; D_1)a_q(x_2;D_2).
    166 \end{align}
    167 Lets say the eigenvalues of $D_1$ are $l^2, l\in \mathbb{Z}$, we
    168 can obtain the heat kernel asymmetries with the Poisson summation formula
    169 giving us an approximation in the order of $e^{-1/t}$
    170 \begin{align}
    171     K(t, D_1) &= \sum_{l\in\mathbb{Z}} e^{-tl^2} = \sqrt{\frac{\pi}{t}}
    172     \sum_{l\in\mathbb{Z}} e^{-\frac{\pi^2l^2}{t}} = \nonumber \\
    173     &\simeq \sqrt{\frac{\pi}{t}} + \mathcal{O}(e^{-1/t}).
    174 \end{align}
    175 The exponentially small terms have no effect on the heat kernel
    176 coefficients and the only nonzero coefficient is $a_0(1, D_1) =
    177 \sqrt{\pi}$, therefore the heat coefficients can be written as
    178 \begin{align}
    179     a_k(f(x^2), D) = \sqrt{\pi}\int_{M_2}
    180     d^{n-1}x\sqrt{g}\sum_I\text{Tr}_V\left(f(x^2)u^I_{(n-1)}
    181     \mathcal{A}^I_n(D_2)\right).
    182 \end{align}
    183 
    184 Since all of the geometric invariants  associated with $D$ are in the $D_2$
    185 part, they are independent of $x_1$. This allows us to to choose for
    186 $M_1$.
    187 
    188 For $M_1 = S^1$ with $x\in (0, 2\pi)$ and $D_1=-\partial_{x_1}^2$
    189 we can rewrite the heat kernel coefficients into
    190 \begin{align}
    191     a_k(f(x_2), D) &= \int_{S^1\times M_2}d^nx \sqrt{g} \sum_I
    192     \text{Tr}_V(f(x_2) u_{(n)}^I \mathcal{A}^I_k(D_2))= \nonumber\\
    193     &= 2\pi \int_{M_2} d^nx\sqrt{g} \sum_I\text{Tr}_V(f(x_2) u_{(n)}^I
    194     \mathcal{A}^I_k(D_2)).
    195 \end{align}
    196 Computing the two equations above we see that
    197 \begin{align}
    198     u_{(n)}^I = \sqrt{4\pi} u^I_{(n+1)}.
    199 \end{align}
    200 
    201 \subsubsection{Heat Kernel Coefficients}
    202 To calculate the heat kernel coefficients consider the following variational
    203 equations
    204 \begin{align}
    205     &\frac{d}{d\varepsilon}\bigg|_{\varepsilon=0}a_k(1, e^{-2\varepsilon f}D) =
    206     (n-k) a_k(f, D),\label{eq:var1}\\
    207     &\frac{d}{d\varepsilon}\bigg|_{\varepsilon=0}a_k(1, D-\varepsilon F) =
    208     a_{k-2}(F,D),\label{eq:var2}\\
    209     &\frac{d}{d\varepsilon}\bigg|_{\varepsilon=0}a_k(e^{-2\varepsilon f}F,
    210     e^{-2\varepsilon f}D) =
    211     0\label{eq:var3}.
    212 \end{align}
    213 Let us explain the equations above. To get the first equation \eqref{eq:var1}
    214 we simply differentiate
    215 \begin{align}
    216     \frac{d}{d\varepsilon}\bigg|_{\varepsilon=0} \text{Tr}(\exp(-e^{-2\varepsilon
    217     f}tD) = \text{Tr}(2ftDe^{-tD}) = -2t\frac{d}{dt}\text{Tr}(fe^{-tD})),
    218 \end{align}
    219 additionally expand both sides in $t$ and get equation \eqref{eq:var1}. Equation
    220 \eqref{eq:var2} is derived similarly.
    221 
    222 For equation \eqref{eq:var3} look at the following operator
    223 \begin{align}
    224     D(\varepsilon,\delta) = e^{-2\varepsilon f}(D-\delta F),
    225 \end{align}
    226 for $k=n$ we take equation \eqref{eq:var1} and we obtain
    227 \begin{align}
    228     \frac{d}{d\varepsilon}\bigg|_{\varepsilon=0}a_n(1,D(\varepsilon,\delta))
    229     =0.
    230 \end{align}
    231 Then we take the variation in terms of $\delta$, evaluated at $\delta =0$ and
    232 swap the differentiation, allowed by theorem of Schwarz
    233 \begin{align}
    234     0 &=
    235     \frac{d}{d\delta}\bigg|_{\delta=0}\frac{d}{d\varepsilon}\bigg|_{\varepsilon=0}a_n(1,
    236     D(\varepsilon,\delta)) =\nonumber\\
    237       &=\frac{d}{d\varepsilon}\bigg|_{\varepsilon=0}\frac{d}{d\delta}\bigg|_{\delta=0}a_n(1,
    238     D(\varepsilon,\delta)) =\nonumber\\
    239       &=a_{n-2} ( e^{-2\varepsilon f}F, e^{-2\varepsilon f}D),
    240 \end{align}
    241 which gives us equation \eqref{eq:var3}.
    242 
    243 Now that the ground basis is established, we can calculate the constants
    244 $u^I$, and by that the first three heat kernel coefficients read
    245 \begin{align}
    246     a_0(f, D) &= (4\pi)^{-n/2}\int_Md^n x\sqrt{g} \text{Tr}_V(a_0 f),\\
    247     a_2(f, D) &= (4\pi)^{-n/2}\frac{1}{6}\int_Md^n
    248     x\sqrt{g}\text{Tr}_V)(f\alpha _1 E+\alpha _2 R),\\
    249     a_4(f, D) &= (4\pi)^{-n/2}\frac{1}{360}\int_Md^n
    250     x\sqrt{g}\text{Tr}_V(f(\alpha_3 E_{,kk} + \alpha_4\ R\ E + \alpha_5 E^2
    251     \alpha_6 R_{,kk} + \nonumber\\
    252     &+\alpha_7 R^2 + \alpha_8 R_{ij}R_{ij} + \alpha_9
    253     R_{ijkl}R_{ijkl} +\alpha_{10} \Omega_{ij}\Omega{ij})),
    254 \end{align}
    255 where the comma subscript $,$ denotes the derivative and constants $\alpha_I$
    256 do not depend on the dimension of the Manifold and we can compute them with
    257 our variational identities.
    258 
    259 The first coefficient $\alpha_0$ can be read from the heat kernel expansion of
    260 the Laplacian on $S^1$ (above), $\alpha_0 = 1$. For $\alpha_1$ we use
    261 \eqref{eq:var2}, the coefficient $k = 2$ is
    262 \begin{align}
    263     \frac{1}{6} \int_M d^n x\sqrt{g} \text{Tr}_V(\alpha_1F) = \int_M d^n
    264     x\sqrt{g} \text{Tr}_V(F),
    265 \end{align}
    266 which means $\alpha_1 = 6$. Looking at the coefficient $k=4$ we have
    267 \begin{align}
    268     \frac{1}{360}\int_Md^n x \sqrt{g}\text{Tr}_V(\alpha_4\ F\ R + 2\alpha_5\ F\ E)
    269     = \frac{1}{6} \int_Md^n x\sqrt{g}\text{Tr}_V(\alpha_1\ F\ E + \alpha_2\ F\ R),
    270 \end{align}
    271 thus $\alpha_4 = 60\alpha_2$ and $\alpha_5 = 180$.
    272 
    273 By applying  \eqref{eq:var3} to $n=4$ we get
    274 \begin{align}
    275     \frac{d}{d\varepsilon}|_{\varepsilon=0} a_2(e^{-2\varepsilon f}F,
    276     e^{-2\varepsilon f}D) = 0.
    277 \end{align}
    278 Collecting the terms with $\text{Tr}_V(\int_Md^nx\sqrt{g}(Ff_{,jj}))$ we
    279 obtain $\alpha_1 = 6\alpha_2$, that is $\alpha_2 = 1$, so $\alpha_4 = 60$.
    280 
    281 Now we let $M=M_1\times M_2$ and split $D = -\Delta_1 -\Delta_2$, where
    282 $\Delta_{1/2}$ are Laplacians for $M_1, M_2$. This allows us to decompose the heat
    283 kernel coefficient for $k=4$ into
    284 \begin{align}
    285     a_4(1,-\Delta_1-\Delta_2) =& a_4(1, -\Delta_1) a_0(1,
    286     -\Delta_2)\nonumber+ \\
    287                                &+a_2(1,-\Delta_1) a_2(1,-\Delta_2)\nonumber \\
    288                                &+ a_0(1,-\Delta_1)a_4(1,-\Delta_2),
    289 \end{align}
    290 with $E=0$ and $\Omega =0$ and by calculating the terms with $R_1R_2$  (scalar
    291 curvature of $M_{1/2}$) we obtain $\frac{2}{360}\alpha_7 =
    292 (\frac{\alpha_2}{6})^2$, thus $\alpha_7 = 5$.
    293 
    294 For $n=6$ we get
    295 \begin{align}
    296     0 &= \text{Tr}_V(\int_Md^nx\sqrt{g}
    297     (F(-2\alpha_3-10\alpha_4+4\alpha_5)f_{,kk}E +\nonumber\\
    298     &+(2\alpha_3 + 10\alpha_6)f_{,iijj}+\nonumber\\
    299     &+(2\alpha_4 -2\alpha_6 - 20\alpha_7 -2\alpha_8)f_{,ii}R\nonumber\\
    300     &+(-8\alpha_8 -8\alpha_6)f_{,ij}R_{ij}))
    301 \end{align}
    302 we obtain $\alpha_3 = 60$, $\alpha_6=12$, $\alpha_8 = -2$ and $\alpha_9 = 2$
    303 
    304 To get $\alpha_{10}$ we use the Gauss-Bonnet theorem, ultimately giving us
    305 $\alpha_{10}=30$. We leave out this lengthy calculation and refer to
    306 \cite{heatkernel} for further reading.
    307 
    308 Let us summarize our calculations which ultimately lead us to the following heat kernel
    309 coefficients
    310 \begin{align}
    311     \alpha_0(f, D) &= (4\pi)^{-n/2}\int_M d^n x \sqrt{g} \text{Tr}_V(f),\\
    312     \alpha_2(f, D) &= (4\pi)^{-n/2}\frac{1}{6}\int_M d^n x \sqrt{g}
    313     \text{Tr}_V(f(6E+R)),\\
    314     \alpha_4(f, D) &= (4\pi)^{-n/2}\frac{1}{360}\int_M d^n x \sqrt{g}
    315     \text{Tr}_V(f(60E_{,kk}+60RE+ 180E^2 +\\
    316     &+12R_{,kk} + 5R^2 - 2 R_{ij}R_{ij}
    317     2R_{ijkl}R_{ijkl} +30\Omega_{ij}\Omega_{ij})).\label{eq: a_4}
    318 \end{align}
    319