Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

In this chapter we introduce the Selberg trace formula, which is a natural generalization of the Poisson summation formula to non-abelian groups. Applications of the trace formula will be given in the next two chapters.

9.1 Cocompact Groups and Lattices

We will study the following situation. G will be a locally compact group, and H will be a closed subgroup, which is unimodular. We say that a subgroup H of a topological group G is a cocompact subgroup if the quotient \(G/H\) is a compact space.

Examples 9.1.1

  • A classical example is \(G={\mathbb R}\) and \(H={\mathbb Z}\).

  • For a unimodular group H and a compact group K together with a group homomorphism \(\eta:K\to{\rm Aut}(H)\) such that the ensuing map \(K\times H\to H\) given by \((k,h)\mapsto \eta(k)(h)\) is continuous, one can form the semi-direct product \(G=H\rtimes K\). As a topological space, one has \(G=H\times K\). The multiplication is \((h,k)(h_1,k_1)=(h\,\eta(k)(h_1),kk_1)\). Then H is a cocompact subgroup of G.

  • For \(G={\rm SL}_2({\mathbb R})\), one uses hyperbolic geometry [Bea95] to construct discrete subgroups \({\Upgamma}=H\subset G\), which provide examples of the situation considered here.

Proposition 9.1.2

Let G be a locally compact group. If G admits a unimodular closed cocompact subgroup, then G is unimodular itself.

Proof

Let \(H\subset G\) be a unimodular closed cocompact subgroup. The inversion on G induces a homeomorphism \(G/H\to H{\backslash} G\) by \(gH\mapsto Hg^{-1}\). Therefore \(H{\backslash} G\) is compact as well. We install a Radon measure μ on the compact Hausdorff space \(H{\backslash} G\) using Riesz’s representation theorem as follows. For \(f\in C(H{\backslash} G)\) choose \(g\in C_c(G)\) with \(^H\!g=f\), where \(^H\!g(x){\ \stackrel {{\rm def}}{=}} \int_Hg(hx)\, dh.\) In this case define \(\int_{H{\backslash} G} f(x)\,d\mu(x)\) by \(\int_Gg(x)\,dx\). For this to be well-defined, we have to show that \(^H\!g=0\) implies \(\int_Gg (x)\,dx=0\). Note that, as H is unimodular, we have \(\int_Hg(h^{-1}x)\,dh=\int_Hg(hx) \,dh=0\). Let \(\phi\in C_c(G)\) with \(^H\phi\equiv 1\). One gets

$$\begin{aligned} \int_G g(x)\,dx &= \int_G\int_H\phi(hx)g(x)\,dh\,dx\\ &= \int_H\int_G\phi(hx)g(x)\,dx\,dh\\ &= \int_H\int_G\phi(x)g(h^{-1}x)\,dx\,dh\\ &= \int_G\phi(x)\int_Hg(h^{-1}x)\,dh\,dx= 0.\end{aligned}$$

This measure μ on \(H{\backslash} G\) satisfies

$$\begin{aligned} \int_{H{\backslash} G}f(xy)\,d\mu(x)&= \int_Gg(xy)\,dx=\Updelta(y^{-1})\int_Gg(x) \,dx\\ &= \Updelta(y^{-1})\int_{H{\backslash} G}f(x)\,d\mu(x).\end{aligned}$$

So in particular, for \(f\equiv 1\in C(H{\backslash} G)\) one gets

$$\begin{aligned} 0 <\ \int_{H{\backslash} G}f(x)\,d\mu(x)&=\int_{H{\backslash} G}f(xy)\,d\mu(x) =\Updelta(y^{-1})\int_{H{\backslash} G}f(x)\,d\mu(x),\end{aligned}$$

which implies \(\Updelta(y^{-1})=1.\)

Definition

A subgroup \({\Upgamma}\subset G\) of a topological group is called a discrete subgroup if the subspace-topology on \({\Upgamma}\) is the discrete topology.

Lemma 9.1.3

  1. (a)

    A subgroup \({\Upgamma}\subset G\) is discrete if and only if there is a unit-neighborhood \(U\subset G\) with \({\Upgamma}\cap U=\{1\}\).

  2. (b)

    A discrete subgroup is closed in G.

Proof

As \({\Upgamma}\) is a topological group, it is discrete if and only if \(\{1\}\) is open, which is equivalent to the existence of an open set U with \({\Upgamma} \cap U=\{1\}\). This implies part (a). For part (b) let \({\Upgamma}\) be a discrete subgroup, and let \({\gamma}_j\to x\) be a net in \({\Upgamma}\), convergent in G. We have to show that \(x\in{\Upgamma}\). Let U be a unit-neighborhood in G such that \({\Upgamma}\cap U=\{1\}\) and let V be a symmetric unit-neighborhood such that \(V^2\subset U\). As xV is a neighborhood of x, there exists an index j 0 such that for \(i,j\ge j_0\) one has \({\gamma}_i,{\gamma}_j\in xV\), therefore \({\gamma}_i^{-1}{\gamma}_j\in V^2\subset U\), hence \({\gamma}_i={\gamma}_j\), so the net is eventually constant, hence \(x={\gamma}_{j_0}\in{\Upgamma}\). □

Example 9.1.4

  • The subgroup \({\mathbb Z}\) of \({\mathbb R}\) is a discrete subgroup.

  • Let G be Hausdorff and compact. Then a subgroup \({\Upgamma}\subset G\) is discrete if and only if it is finite.

  • Let \(G={\rm SL}_2({\mathbb R})\) be the group of real 2 × 2 matrices of determinant one. Then \({\Upgamma}={\rm SL}_2({\mathbb Z})\) is a discrete subgroup of G.

Definition

Let G be a locally compact group. A discrete subgroup \({\Upgamma}\) such that \(G/{\Upgamma}\) carries an invariant Radon measure μ with \(\mu(G/{\Upgamma})\ <\ \infty\) is called a lattice in G. A cocompact lattice is also called a uniform lattice .

Proposition 9.1.5

Let G be a locally compact group. A discrete, cocompact subgroup \({\Upgamma}\) is a uniform lattice.

Proof

Let \({\Upgamma}\subset G\) as in the proposition. Then G is unimodular by Proposition 9.1.2. As \({\Upgamma}\) is unimodular, by Theorem 1.5.3 there exists an invariant Radon measure on \(G/{\Upgamma}\), and, as the latter space is compact, its volume is finite, so \({\Upgamma}\) is a lattice. □

Remark

There are lattices, which are not uniform, like \({\Upgamma}={\rm SL}_2({\mathbb Z})\) in \(G={\rm SL}_2({\mathbb R})\) (See [Ser73]).

Theorem 9.1.6

Let G be a locally compact group. If G admits a lattice, then G is unimodular.

Proof

Let Δ be the modular function of G, and let \(H={\rm ker}(\Updelta)\). Then H is unimodular by Corollary 1.5.5. Let \({\Upgamma}\subset G\) be a lattice. As there is an invariant measure on \(G/{\Upgamma}\), Theorem 1.5.3 implies that Δ is trivial on \({\Upgamma}\), so \({\Upgamma}\subset H\). On \(G/H\) and on \(H/{\Upgamma}\) there are invariant measures by Theorem 1.5.3, so by the uniqueness of these respective measures we conclude that

$$\infty>\ {\rm vol}(G/{\Upgamma})=\int_{G/{\Upgamma}} 1\,dx=\int_{G/H}\int_{H/{\Upgamma}} 1\,dx \,dy.$$

Therefore \(G/H\) has finite volume. Being a locally compact group, it follows that \(G/H\) is compact. By Proposition 9.1.2. the group G is unimodular. □

9.2 Discreteness of the Spectrum

Let G be a locally compact group, and let \({\Upgamma}\subset G\) be a cocompact lattice. It is a convention that one uses the right coset space \({\Upgamma}{\backslash} G\) instead of the left coset space. This quotient carries a Radon measure μ that is invariant under right translations by elements of G and has finite volume. On the ensuing Hilbert space \(L^2({\Upgamma}{\backslash} G)\) the right translations give a unitary representation R of G by

$$R(x)\phi(y){\ \stackrel{{\rm def}}{=}\,} \phi(yx).$$

Example 9.2.1

Let \(G={\mathbb R}\), and \({\Upgamma}={\mathbb Z}\) then \({\Upgamma}\) is a closed cocompact subgroup and the theory of Fourier series, [Dei05] Chap. 1, implies that the representation R decomposes into the direct sum \(\widehat{\bigoplus}_{n\in{\mathbb Z}}\chi_n\), where \(\chi_n:{\mathbb R}\to{\mathbb T}\) is the character given by \(\chi_n(x)=e^{2\pi i n x}\). Thus R decomposes into a discrete sum of irreducible representations. This is a general phenomenon, as the next theorem shows.

For a representation π and a natural number N we write \(N\pi\) for the N-fold direct sum \(\pi\oplus\pi\dots\oplus\pi\).

Theorem 9.2.2

Let G be a locally compact group, and let \({\Upgamma}\subset G\) be a unimodular closed cocompact subgroup. The representation R on \(L^2({\Upgamma}{\backslash} G)\) decomposes as a direct sum of irreducible representations with finite multiplicities, i.e.,

$$L^2({\Upgamma}{\backslash} G)\ \cong\ \bigoplus_{\pi\in \widehat G} N_{\Upgamma}(\pi)\pi,$$

where the sum runs over the unitary dual \(\widehat G\) of G, and \(N_{\Upgamma}(\pi)\in{\mathbb N}_0\) is a finite multiplicity for \(\pi\in\widehat G\).

Proof

We need a lemma that tells us that for \(f\in C_c(G)\) the operator R(f) is given by a continuous integral kernel. For later use we will extend this to a greater class of functions f. Let U be a compact unit-neighborhood in G. For a continuous function f on G let \(f_U: G\to [0,\infty)\) be defined by

$$f_U(y){\ \stackrel{{\rm def}}{=}} \sup_{x,z\in U}|f(xyz)|.$$

Lemma 9.2.3

The function f U is continuous.

Proof

It suffices to show that for \(a\ge 0\) the sets \(f_U^{-1}((a,\infty))\) and \(f_U^{-1}([0,a))\) are open. For the former assume \(f_U(x)>a\), then there exist \(u_1,u_2\in U\) with \(|f(u_1xu_2)|>a\). As the function \(y\mapsto f(u_1yu_2)\) is continuous, there exists an open neighborhood V of x such that \(|f(u_1vu_2)|>a\) for every \(v\in V\). This implies \(f_U(v)>a\) for every \(v\in V\).

The second assertion is equivalent to saying that \(f_U^{-1}([a,\infty))\) is closed. So let (x j ) be a net in this set, convergent to some \(x\in G\). This means that \(f_U(x_j)\ge a\) and we have to show that \(f_U(x)\ge a\). For each j fix some \(u_j,v_j\in U\) such that \(f_U(x_j)=|f(u_jx_jv_j)|\). By switching to a subnet, we may assume that the nets \((u_j),(v_j)\) are both convergent in the compact set U with limits u and v respectively. It follows that

$$\begin{aligned} f_U(x)\ge |f(uxv)|=\lim_j|f(u_jx_jv_j|\ge a. \end{aligned}$$

Definition

We say that a continuous function f is a uniformly integrable function if there exists a compact unit-neighborhood U such that f U is in \(L^1(G)\). If f is uniformly integrable, then \(f\in L^1(G)\) as \(|f|\le f_U\). Let \(C_{\rm unif}(G)\) be the set of all continuous functions f on G that are uniformly integrable.

Example 9.2.4

  • If \(f\in C_c(G)\), then f is uniformly integrable, as \(f_U\in C_c(G)\) again.

  • Every Schwartz-function on \({\mathbb R}\) is uniformly integrable (Exercise 9.1).

Lemma 9.2.5

Let G be unimodular. Every uniformly integrable function vanishes at infinity, and the space \(C_{\rm unif}(G)\) is an algebra under convolution.

Proof

Assume that f is uniformly integrable and let U be a compact symmetric unit-neighborhood in G such that f U is integrable. If f does not vanish at ∞, there exists an \({\varepsilon}>0\) such that for every compact set \(K\subset G\) there is \(x\in G{\smallsetminus} K\) with \(|f(x)|\ge{\varepsilon}\). Let \(x_1\in G\) be any element with \(|f(x_1)|\ge{\varepsilon}\). Then there exists \(x_2\notin x_1U^2\) such that \(|f(x_2)|\ge{\varepsilon}\). As U is symmetric, it follows that \(x_1U\cap x_2U=\emptyset\). Next we pick an element x 3 outside of \(x_1U^2\cup x_2U^2\) with \(|f(x_3)|\ge{\varepsilon}\). Repeating this argument, we find a sequence \(\{x_n: n\in{\mathbb N}\}\) in G such that \(x_nU\cap x_mU=\emptyset\) for all \(n\neq m\) and \(|f(x_n)|\geq{\varepsilon}\) for every n. But then \(f_U\geq{\varepsilon}\) on x nU for every n, which contradicts the integrability of f U .

Since integrable functions in \(C_0(G)\) are square integrable, we get

$$C_{\rm unif}(G)\ \subset\ L^2(G).$$

Let \(f,g\in C_{\rm unif}(G)\). We can write \(f*g(x)={\left\langle{f,L_xg^*}\right\rangle}\). The map \(x\mapsto L_xg^*\) is continuous as a map from G to \(L^2(G)\). The inner product is continuous, hence so is \(f*g\). Finally, choose a unit-neighborhood U such that f U and g U are both integrable. Then

$$\begin{aligned} (f*g)_U(y) = & \sup_{x,z\in U}\left|\int_G f(\xi)g(\xi^{-1} xyz)\,d\xi \right| \\ = & \sup_{x,z\in U}\left|\int_G f(x\xi)g(\xi^{-1} yz)\,d\xi\right|\\ &\le \sup_{x,z\in U}\int_G |f(x\xi)g(\xi^{-1} yz)|\,d\xi\\ &\le \int_G f_U(\xi)g_U(\xi^{-1} y)\,d\xi = f_U *g_U(y).\end{aligned}$$

This implies that \((f*g)_U\) is integrable over G. □

Lemma 9.2.6

For \(f\in C_{\rm unif}(G)\) and \(\phi\in L^2({\Upgamma}{\backslash} G)\) one has

$$R(f)\phi(x)=\int_{{\Upgamma}{\backslash} G}k(x,y)\phi(y)\,dy,$$

where \(k(x,y)=\sum_{{\gamma}\in{\Upgamma}}f(x^{-1}{\gamma} y)\,dh\). The kernel k is continuous on \({\Upgamma}{\backslash} G\times{\Upgamma}{\backslash} G\).

Proof

Let \(f\in L^1(G)\). For \(\phi\in L^2({\Upgamma}{\backslash} G)\), one computes with the quotient integral formula

$$\begin{aligned} R(f)\phi(x) &= \int_G f(y) R(y)\phi(x)\,dy=\int_Gf(y)\phi(xy)\,dy\\ &= \int_Gf(x^{-1}y)\phi(y)\,dy\\ &= \int_{{\Upgamma}{\backslash} G}\sum_{{\gamma}\in{\Upgamma}}f(x^{-1}{\gamma} y)\phi({\gamma} y)\,dy = \int_{{\Upgamma}{\backslash} G}\sum_{{\gamma}\in{\Upgamma}}f(x^{-1}{\gamma} y)\phi(y)\,dy,\end{aligned}$$

as claimed, and the integral converges almost everywhere in x. In particular, for \(f\in C_{\rm unif}(G)\) this argument works with f replaced by f U for a suitable symmetric unit-neighborhood U to get a kernel k U . We choose \(g\in C_c(G)\) with \(g\ge 0\) and \(\sum_{{\gamma}\in{\Upgamma}}g({\gamma} x)= 1\) for all \(x\in G\) and use the quotient integral formula to compute

$$\begin{aligned} \int_{{\Upgamma}{\backslash} G\times{\Upgamma}{\backslash} G}k_U(x,y)\,dx\,dy &= \int_{{\Upgamma}{\backslash} G}\int_{{\Upgamma}{\backslash} G}\sum_{{\gamma}\in{\Upgamma}}g({\gamma} x)\sum_{\tau\in{\Upgamma}}f_U(x^{-1} \tau y)\,dx\,dy\\ &= \int_{{\Upgamma}{\backslash} G}\int_{G}g(x)\sum_{\tau\in{\Upgamma}}f_U(x^{-1} \tau y)\,dx\,dy\\ &= \int_{G}\int_{G}g(x)f_U(x^{-1}y)\,dx\,dy=\Vert{g*f_U}\Vert_1<\infty\end{aligned}$$

By the quotient integral theorem, the sum \(\sum_{{\gamma}\in{\Upgamma}} f_U(x^{-1}{\gamma} y)\) converges almost everywhere in \((x,y)\), so it converges on a dense set of \((x,y)\). Let \((x_0,y_0)\) be such a point of convergence. We show that \(k(x,y)\) is continuous in the set \(x_0U\times y_0U\). For given \({\varepsilon}>0\) there exists a finite set \(S\subset{\Upgamma}\) such that \(\sum_{{\gamma}\notin S}f_U(x_0^{-1}{\gamma} y_0)<{\varepsilon}/2\). This means that for \((x,y)\in x_0U\times y_0U\) one has \(\sum_{{\gamma}\notin S}|f(x^{-1}{\gamma} y)|<{\varepsilon}/2\). This implies for \((x',y')\in x_0U\times y_0U\),

$$|k(x,y)-k(x',y')|\le\sum_{{\gamma}\in S}|f({x}^{-1}{\gamma}y)-f({x'}^{-1}{\gamma} y')|+{\varepsilon},$$

from which the continuity of k follows □

As \(k\in C({\Upgamma}{\backslash} G\times{\Upgamma}{\backslash} G)\subset L^2({\Upgamma}{\backslash} G\times{\Upgamma}{\backslash} G)\), the operator R(f) is a Hilbert-Schmidt operator, hence compact. The theorem follows from the next lemma.

Lemma 9.2.7

Let A be a *-closed subspace of \(C_c(G).\) Let \((\eta, V_\eta)\) be a unitary representation of G such that for every \(f\in A,\) the operator \(\eta(f)\) is compact and such that for every non-zero \(v\in V_\eta\) the space \(\eta(A)v\) is non-zero. Then η is a direct sum of irreducible representations with finite multiplicities.

Note that the condition \(\eta(A)v\ne 0\) is always satisfied if A contains a Dirac-net \((\phi_U)_U\).

Proof

Zorn’s Lemma provides us with a subspace E, maximal with the property that it decomposes as a sum of irreducibles. The assumption of the Lemma also holds for the orthocomplement \(E^\perp\) of E in \(V=V_\eta\). This orthocomplement cannot contain any irreducible subspace. We have to show that it is zero. In other words, we have to show that a representation η as in the assumption, always contains an irreducible subspace.

The space A is generated by its self-adjoint elements. Let \(f\in A\) be self-adjoint. Then \(\eta(f)\) is self-adjoint and compact. By the spectral theorem for compact operators 5.2.2 the space \(V=V_\eta\) decomposes,

$$V= V_{f,0}\ \oplus\ \bigoplus_{j=1}^\infty V_{f,j},$$

where \(V_{f,0}\) is the kernel of \(\eta(f)\) and \(V_{f,j}\) is the eigenspace of \(\eta(f)\) for an eigenvalue \({\lambda}_j\ne 0\). The sequence \({\lambda}_j\) tends to zero and each \(V_{f,j}\) is finite dimensional for \(j>0\). For every closed invariant subspace \(E\subset V\) one has a similar decomposition

$$E= E_{f,0}'\ \oplus\ \bigoplus_{j=1}^\infty E_{f,j}',$$

and \(E'_{f,j}\subset V_{f,j}\) for every \(j\ge 0\). It follows that every non-zero closed invariant subspace E has a non-zero intersection with one of the \(V_{f,i}\) for some \(i>0\) and some \(f\in A\), since otherwise \(E\subset{\rm ker}\,\eta(f)\) for every \(f\in A\). Fix f and j and consider the set of all non-zero intersections \(V_{f,j} \cap E\), where E runs over all closed invariant subspaces. Among these intersections choose one \(W=V_{f,j}\cap E\) of minimal dimension \(\ne 0\), which is possible as \(V_{f,j}\) is finite-dimensional. Let \(E^1=\bigcap_{E:E\cap V_{f,j}=W}E,\) where the intersection runs over all closed invariant subspaces E with \(E \cap V_{f,j}=W\). Then E 1 is a closed invariant subspace. We claim that it is irreducible. For this assume that \(E^1=F\oplus F'\) with closed invariant subspaces \(F,F'\). Then by minimality, \(W\subset F\) or \(W\subset F'\), which implies that one of the spaces F or \(F'\) is zero, so E 1 is indeed irreducible. We have shown that V indeed has an irreducible subspace E 1. As shown above this implies that V decomposes as a sum of irreducibles.

It remains to show that the multiplicities are finite. For this note first that if τ and σ are unitarily equivalent representations, then \({\lambda}\) is an eigenvalue of \(\tau(f)\) if and only if it is one for \(\sigma(f)\). Thus finiteness of the multiplicities follows from the fact that any collection of orthogonal subspaces, which give rise to unitarily equivalent representations, must all have non-trivial intersection with the same eigenspaces \(V_{f,j}\). Since the \(V_{f,j}\) are finite dimensional, there can only exist finitely many of them. □

9.3 The Trace Formula

Let X be a locally compact Hausdorff space, and let μ be a Radon measure on X. A continuous L 2-kernel k on X is called admissible kernel if there exists a function \(g\in C(X)\cap L^2(X)\) such that \(|k(x,y)|\le g(x)g(y)\). Note that if X is compact, then every continuous kernel is admissible.

An operator \(S: L^2(X)\to L^2(X)\) is called an admissible operator if there exists an admissible kernel k such that

$$S\phi(x)=\int_X k(x,y)\phi(y)\,dy,$$

where we have written dy for \(d\mu(y)\).

Proposition 9.3.1

Let X be a locally compact Hausdorff space equipped with a Radon measure. Assume that X is first countable or compact. Let \(T\) be an integral operator with continuous L 2-kernel on X. Assume that there exists admissible operators \(S_1,S_2\) with \(T=S_1S_2.\)

Then T is of trace class and

$$\hspace{1pt}{\rm tr}\hspace{2pt}(T)=\int_X k(x,x)\,dx.$$

Proof

Replacing S 2 with \(S_2^*\) we can assume \(T=S_1S_2^*\) in the proposition. The map \(\sigma:(S_1,S_2)\mapsto S_1S_2^*\) is sesquilinear, so it obeys the polarization rule,

$$\sigma(S,R)=\frac14\left(\sigma(S+R)-\sigma(S-R)+i\left(\sigma(S+iR)-\sigma(S-iR)\right)\right),$$

where we have written \(\sigma(S+R)\) for \(\sigma(S+R,S+R)\). This implies that, in order to prove the proposition, it suffices to assume that \(T=SS^*\) for some admissible operator S. As S is a Hilbert-Schmidt operator, \(T=SS^*\) is trace class and also \({\hspace{1pt}{\rm tr}\hspace{2pt}}(T)=\Vert S\Vert_{\rm HS}^2\). Let \(l(x,y)\) be the admissible kernel of S. For \(\phi\in C_c(X)\) we compute

$$\begin{aligned} SS^*\phi(x) &= \int_Xl(x,u)S^*\phi(u)\,du\\ &=\int_X\int_X l(x,u){\overline{l(y,u)}}\phi(y)\,dy\,du\\ &=\int_X\left(\int_X l(x,u){\overline{l(y,u)}}\,du\right)\phi(y)\,dy.\end{aligned}$$

So the operator \(SS^*\) has the kernel \(l*l^*(x,y)=\int_X l(x,z){\overline{l(y,z)}}\,dz.\) If l is admissible and X is first countable, it suffices to check convergence with sequences, and therefore the Theorem of Dominated Convergence implies that the kernel \(l*l^*\) is continuous. If X is compact, the Radon measure is finite and the function l is uniformly continuous in the following sense: For every \(x_0\in X\) and every \({\varepsilon}>0\) there exists a neighborhood U of x 0 such that for all \(x\in U\) and all \(z\in X\) we have \(|l(x,z)-l(x_0,z)|<{\varepsilon}.\) This implies the continuity of \(l*l^*\). For \(\phi,\psi\in C_c(X)\) we have

$$\begin{aligned} \int_X\int_X l*l^*(x,y)\phi(y)\psi(x)\,dy\,dx &={\left\langle{SS^*}\phi,{\overline{\psi}}\right\rangle}\\ &=\int_X\int_X k(x,y)\phi(y)\psi(x)\,dy\,dx.\end{aligned}$$

Varying φ and ψ we conclude that the continuous kernels k and \(l*l^*\) coincide. Therefore,

$$\int_{X} k(x,x)\,dx {=}\int_{X}\int_{X} |l(x,u)|^{2}\, du \, dx{=}\Vert{S}\Vert_{\rm HS}={\hspace{1pt}{\rm tr}\hspace{2pt}}(T).$$

The proposition follows. □

Recall that for \(\pi\in{\widehat G}\) the number \(N_{\Upgamma}(\pi)\ge 0\) is the multiplicity of π as a subrepresentation of \((R,L^2({\Upgamma}{\backslash} G))\). Let \(\widehat{G}_{\Upgamma}\) denote the set of all \(\pi\in\widehat{G}\) with \(N_{\Upgamma}(\pi)>~0\).

Definition

We write \(C_{\rm unif}(G)^2\) for the space of all linear combinations of functions of the form \(g*h\) with \(g,h\in C_{\rm unif}(G)\). Moreover, if ν is a lattice in G and \({\gamma}\in{\Upgamma}\), then we denote by \([{\gamma}]\) the conjugacy class of \({\gamma}\) in \({\Upgamma}\), \(G_{\gamma}{\ \stackrel{ {\rm def}}{=}\,}\{x\in G: x{\gamma} x^{-1}={\gamma}\}\) denotes the centralizer of \({\gamma}\) in G, and \({\Upgamma}_{\gamma}{\ \stackrel{{\rm def}}{=}\,} G_{\gamma}\cap{\Upgamma}\) denotes the centralizer of \({\gamma}\) in \({\Upgamma}\). Note that the map from \({\Upgamma}\) to \([{\gamma}]\), which sends ν to \(\nu^{-1}{\gamma}\nu\), factors through a bijection \({\Upgamma}_{\gamma}{\backslash}{\Upgamma}\cong [{\gamma}]\).

Theorem 9.3.2

(Trace Formula) Let G be a locally compact group and \({\Upgamma}\subset G\) a uniform lattice. Let \({\widehat{G}}_{\Upgamma}\) denote the set of all \(\pi\in \widehat{G}\) which appear as subrepresentations of the representation R on \(L^2({\Upgamma}{\backslash} G)\) and let \(f\in C_{\rm unif}(G)^2.\) For every \(\pi\in\widehat G_{\Upgamma}\) the operator \(\pi(f)\) is of trace class and

$$\sum_{\pi\in\widehat G_{\Upgamma}}N_{\Upgamma}(\pi)\,{\hspace{1pt}{\rm tr}\hspace{2pt}}\pi(f)=\sum_{[{\gamma}]}{\rm vol}({\Upgamma}_{\gamma}{\backslash} G_{\gamma})\,{\cal O}_{\gamma}(f),$$

where the summation on the right runs over all conjugacy classes \([{\gamma}]\) in the group \({\Upgamma},\) and \({\cal O}_{\gamma}(f)\) denotes the orbital integral ,

$$\cal O_{\gamma}(f)=\int_{G_{\gamma}{\backslash} G} f(x^{-1}{\gamma} x)\,dx.$$

We shall see in Lemma 9.3.3 below that the centralizer \(G_{\gamma}\) is unimodular and that \({\Upgamma}_{\gamma}{\backslash} G_{\gamma}\) has finite measure for every \({\gamma}\in{\Upgamma}\). The expression \({\rm vol}({\Upgamma}_{\gamma}{\backslash} G_{\gamma}){\cal O}_{\gamma}(f)\) is therefore well-defined. It does not depend on the choice of Haar measure on \(G_{\gamma}\).

Proof of the trace formula

The algebra \(C_{\rm unif}(G)^2\) consists of all finite linear combinations of convolution products of the form \(g*h^*\) with \(g,h\in C_{\rm unif}(G)\). So it suffices to show the trace formula for \(f=g*h^*\). According to Lemma 9.2.6, the operators R(g) and R(h) are integral operators with continuous kernels

$$k_g(x,y)=\sum_{{\gamma}\in{\Upgamma}}g(x^{-1}{\gamma} y)\qquad{\rm and}\qquad k_h(x,y)=\sum_{{\gamma}\in{\Upgamma}}h(x^{-1}{\gamma} y).$$

By Proposition 9.3.1 the operator R(f) is of trace class and \({\hspace{1pt}{\rm tr}\hspace{2pt}}(R(f))=\int_{{\Upgamma}{\backslash} G} k_f(x,x)\,dx.\) With R(f), all its restrictions to subrepresentations are of trace class. It follows that

$$\begin{aligned} \sum_{\pi\in\widehat G_{\Upgamma}}N_{\Upgamma}(\pi){\hspace{1pt}{\rm tr}\hspace{2pt}}\pi(f)&={\hspace{1pt}{\rm tr}\hspace{2pt}}(R(f)) =\int_{{\Upgamma}{\backslash} G}k_f(x,x)\,dx\\ &=\int_{{\Upgamma}{\backslash} G}\sum_{{\gamma}\in{\Upgamma}}f(x^{-1}{\gamma} x)\,dx.\end{aligned}$$

We order the sum according to conjugacy classes \([{\gamma}]\) in \({\Upgamma}\), interchange integration and summation and use the quotient integral formula to get

$$\begin{aligned} {\hspace{1pt}{\rm tr}\hspace{2pt}}(R(f))&=\int_{{\Upgamma}{\backslash} G}\sum_{[{\gamma}]}\sum_{\sigma\in \Upgamma_\gamma\backslash \Upgamma}f(x^{-1}\sigma^{-1} {\gamma} \sigma x)\,dx\\ &=\sum_{[{\gamma}]}\int_{{\Upgamma}{\backslash} G}\sum_{\sigma\in \Upgamma_\gamma\backslash \Upgamma} f((\sigma x)^{-1} {\gamma} \sigma x)\,dx\\ &=\sum_{[{\gamma}]}\int_{{\Upgamma}_{\gamma}{\backslash} G}f(x^{-1} {\gamma} x)\,dx.\end{aligned}$$

Lemma 9.3.3

For every \({\gamma}\in{\Upgamma},\) the centralizer \(G_{\gamma}\) is unimodular and \({\Upgamma}_{\gamma}{\backslash} G_{\gamma}\) has finite invariant measure \({\rm vol}({\Upgamma}_{\gamma}{\backslash} G_{\gamma})\).

Proof

The above calculation shows that for every f as in the theorem, which is positive, i.e., for \(f\ge 0\), one has \(\int_{{\Upgamma}_{\gamma}{\backslash} G} f(x^{-1}{\gamma} x)\,dx < \infty\) for every \({\gamma}\in{\Upgamma}\). Consider the projection \(p:{\Upgamma}_{\gamma}{\backslash} G\to G_{\gamma}{\backslash} G\). Since G is unimodular by Theorem 9.1.6, the space \({\Upgamma}_{\gamma}{\backslash} G\) carries an invariant Radon measure ν. Let μ be the image under p of ν, i.e., we define \(\int_{G_{\gamma}{\backslash} G} f(x)\,d\mu(x)= \int_{{\Upgamma}_{\gamma}{\backslash} G} f(p(y))\,d\nu(y)\) for every \(f\in C_c(G_{\gamma}{\backslash} G)\). We need to show that this is finite. For this let \(0\le f\in C_c(G_{\gamma}{\backslash} G)\), and let \(\Upphi: G_{\gamma}{\backslash} G\to G\) be given by \(\Upphi(G_{\gamma} x)=x^{-1}{\gamma} x\). Then \(K=\Upphi({\operatorname{supp}} f)\subseteq G\) is compact in G and by Tietze’s extension theorem we can find a function \(0\le \tilde f\in C_c(G)\) such that \(\tilde f(y^{-1}{\gamma} y)=f(G_{\gamma} y)\) for every \(y\in G\) with \(f(G_{\gamma} y)>0\). Choose \(0\le F\in C_c(G)^2\) such that \(\tilde f\le F\). To show the existence of such a function, let \(g\ge 0\) in C c (G) such that \(g>0\) in a neighborhood of \({\operatorname{supp}}(f)\). There exists a unit-neighborhood U, such that \(\phi_U*g\) is \(>0\) on the set \({\operatorname{supp}}(f)\), where φ U is a Dirac function of support in U. Set \(F=c \phi_U*g\) for some sufficiently large \(c>0\), then \(f\in C_c(G)^2\) and \(F\ge f\). Further F satisfies the conditions of the theorem, so we get

$$\begin{aligned} \int_{G_{\gamma}{\backslash} G} f(x)\,dx = & \int_{{\Upgamma}_{\gamma}{\backslash} G} f(p(y))\,d\nu(y) =\int_{{\Upgamma}_{\gamma}{\backslash} G} \tilde f(y^{-1}{\gamma} y)\,d\nu(y)\\ &\le \int_{{\Upgamma}_{\gamma}{\backslash} G} F(y^{-1}{\gamma} y)\, d\nu(y),\end{aligned}$$

which is finite by computations preceding the lemma. Thus μ is a well defined Radon measure. One easily checks that it is invariant. By Theorem 1.5.3 it follows that the modular function of \(G_{\gamma}\) agrees with the one of G. As G is unimodular, so is \(G_{\gamma}\). Finally, since for every function in \(C_c(G_{\gamma}{\backslash} G)\) the function \(f\circ p\) is invariant under \(G_{\gamma}\), it follows from the finiteness of \(\int_{{\Upgamma}_{\gamma}{\backslash} G} f(p(y))\,d\nu(y)\) that \({\Upgamma}_{\gamma}{\backslash} G_{\gamma}\) has finite invariant volume. □

By the lemma, we can continue the calculation above to arrive at

$$\begin{aligned} {\hspace{1pt}{\rm tr}\hspace{2pt}} R(f)&= \sum_{[{\gamma}]}\int_{G_{\gamma}{\backslash} G}\int_{{\Upgamma}_{\gamma}{\backslash} G_{\gamma}} f((\sigma x)^{-1}{\gamma} \sigma x)\,d\sigma\,dx\\ &=\sum_{[{\gamma}]}{\rm vol}({\Upgamma}_{\gamma}{\backslash} G_{\gamma})\,{\cal O}_{\gamma}(f).\end{aligned}$$

This proves the theorem. □

Example 9.3.4

Consider the case \(G={\mathbb R}\) and \({\Upgamma}={\mathbb Z}\). Every \(t\in{\mathbb R}\) gives a character \(x\mapsto e^{2\pi itx}\), and in this way we identify \(\widehat G\) with \({\mathbb R}\). The subset \(\widehat G_{\Upgamma}\) is mapped to \({\mathbb Z}\) and the multiplicities \(N_{\Upgamma} (\pi)\) are each equal to one. Therefore the spectral side of the trace formula equals \(\sum_{k\in{\mathbb Z}} \hat f(k),\) and the geometric side equals \(\sum_{k\in{\mathbb Z}} f(k).\) In other words, the trace formula is the same as the Poisson summation formula.

For applications of the trace formula, the following lemma will be important.

Lemma 9.3.5

Let G be a locally compact group, and let A be a *-subalgebra of C c (G) which contains a Dirac net \((\phi_U)_U.\) Assume further that A is stable under left translations L y , \(y\in G.\) Let \((\pi, V)\) and \((\sigma,W)\) be unitary representations of G such that for every \(f\in A\) the operators \(\pi(f)\) and \(\sigma(f)\) are trace class and that \({\hspace{1pt}{\rm tr}\hspace{2pt}}\pi(f)={\hspace{1pt}{\rm tr}\hspace{2pt}}\sigma(f).\) Then π and σ are both direct sums of irreducible representations with finite multiplicities, and they are equivalent.

Proof

The fact that π and σ are both direct sums of irreducible representations with finite multiplicities is Lemma 9.2.7. There is a subrepresentation of π, maximal with the property of being isomorphic to a subrepresentation of σ. Restricting to the orthogonal space we can assume that π and σ do not have isomorphic subrepresentations. We have to show that they both are zero.

Let \(V=\bigoplus_{{\alpha}\in I}V_{\alpha}\) be a decomposition in pairwise orthogonal subrepresentations and let \(v_{{\alpha},\mu}\in V_{\alpha}\) be vectors, such that

$$\sum_{\alpha}\sum_\mu\Vert{\pi(f) v_{{\alpha},\mu}}\Vert^2\ <\ \infty$$

for every \(f\in A\). Choose a vector \(0\ne w\in W\). We claim that for every \({\varepsilon}>0\) there exists some \(f\in A\), such that

$$\sum_{\alpha}\sum_\mu\Vert{\pi (f)v_{{\alpha},\mu}}\Vert^2<{\varepsilon}\Vert{\sigma(f)w}\Vert^2.$$

If this were not the case, the map

$$\sum_{\alpha}\sum_\mu\pi(f)v_{{\alpha},\mu}\mapsto\sigma(f)w, \qquad f\in A,$$

would be well-defined and could be extended to a non-trivial intertwining map from the closure L of the space

$$\left\{\sum_{\alpha}\sum_\mu\pi(f)v_{{\alpha},\mu}: f\in A\right\}\subset \bigoplus_{\alpha}\bigoplus_\mu V_{{\alpha},\mu}$$

to W. Here every \(V_{{\alpha},\mu}\) is a new copy of the space \(V_{\alpha}\). The space L decomposes as a direct sum of irreducibles and it only contains isotypes, which don’t occur in W, so such an intertwiner cannot exist.

Now assume \(\sigma\ne 0\). As A contains Dirac functions of arbitrary small supports, there is a function h in A, such that \(\sigma(h)\ne 0\). Let \(f=h*h^*\). Then \(\sigma(f)\) is of trace class and positive. Therefore \(\sigma(f)\) possesses a largest eigenvalue and we can scale h in such a way that this eigenvalue is equal to 1. Let \(w\in W\) of norm one with \(\sigma(f)w=w\). Let \({\lambda}>0\) be the largest eigenvalue of \(\pi(f)\). For every \({\alpha}\), let \((v_{{\alpha},\mu})\) be an orthonormal basis of \(V_{\alpha}\) consisting of eigenvectors of \(\pi(f)\), say \(\pi(f)v_{{\alpha},\mu} ={\lambda}_{{\alpha},\mu} v_{{\alpha},\mu}\). For every \(g\in A\), the sum \(\sum_{\alpha}\sum_\mu\Vert{\pi(g)v_{{\alpha},\mu}}\Vert^2\) equals the square of the Hilbert-Schmidt norm of \(\pi(g)\) and therefore is finite. By the above remark there is a \(g\in A\) with

$$\sum_{\alpha} \sum_\mu\Vert{\pi(g)v_{{\alpha},\mu}}\Vert^2\ <\ \frac 1{{\lambda}^2} \Vert{\sigma(g) w}\Vert^2.$$

The trace of \(\pi(g*f)^*\pi(g*f)\) equals

$$\begin{aligned} \sum_{\alpha}\sum_\mu\Vert{\pi(g*f)v_{{\alpha},\mu}}\Vert^2 = & \sum_{\alpha}\sum_\mu{\lambda}_{{\alpha},\mu}^2\Vert{\pi(g)v_{{\alpha}\mu}}\Vert^2\\ &\le{\lambda}^2\sum_{\alpha}\sum_\mu\Vert{\pi(g)v_{{\alpha},\mu}}\Vert^2.\end{aligned}$$

The right hand side is strictly smaller than \(\Vert{\sigma(g)w}\Vert^2=\Vert{\sigma(g*f)w}\Vert^2,\) which again is smaller than

$$\hspace{1pt}{\rm tr}\hspace{2pt}\left(\sigma(g*f)^*\sigma(g*f)\right)={\hspace{1pt}{\rm tr}\hspace{2pt}}\left(\pi(g*f)^*\pi(g*f)\right),$$

a contradiction!

This implies \(\sigma=0\) and, by symmetry, also \(\pi=0\). □

9.4 Locally Constant Functions

Let G be a locally compact group with a cocompact lattice \({\Upgamma}\subset G\). A function f on G is called locally constant , if for every \(x\in G\) there exists a neighborhood U x of x such that f is constant on U x . There are not many locally constant functions if G is connected, but there are many if G is totally disconnected.

The function f is called uniformly locally constant , if there exists a unit-neighborhood U such that f is constant on every set of the form UxU, \(x\in G\). For example, if f is locally constant and of compact support, then f is uniformly locally constant.

Proposition 9.4.1

Let G be totally disconnected and f a uniformly locally constant and integrable function on G. Then \(f\in C_{\rm unif}(G)^2.\) Hence the trace formula is valid for f.

Proof

Let U denote a compact open subgroup such that f is constant on every set of the form UxU. Then \(f_U=|f|\) and therefore \(f\in C_{\rm unif}(G)\). The same holds for the function \(g=\frac{1}{{\rm vol}(U)}1_U\) and one sees that \(f=g*f\), so \(f\in C_ {\rm unif}(G)^2\) as claimed.

9.5 Lie Groups

In this section, we shall present a simple class of functions which satisfy the trace formula in the case of a Lie-group, i.e., a group which is a smooth manifold such that the group operations are smooth maps. We shall freely make use of the notion of a smooth manifold as in [War83].

Theorem 9.5.1

Let G be a Lie group of dimension n and let \({\Upgamma}\subset G\) be a cocompact lattice. Let f be a continuous integrable function on G, such that the sum

$$k(x,y)=\sum_{{\gamma}\in{\Upgamma}}f(x^{-1}{\gamma} y)$$

converges uniformly and the kernel k is 2r-times continuously differentiable, where r is the smallest integer with \(r>n/2.\) Then the trace formula is valid for f.

In particular, the trace formula holds for every \(f\in C_c^r(G)\).

Proof

To show the theorem, we need a partition of unity with a smooth square-root. The easiest way to achieve that is to adapt the classical construction of a partition of unity as is done in the following proposition.

Lemma 9.5.2

Let M be a smooth manifold and let \((U_i)_{i\in I}\) be an open covering of M. Then there are smooth functions \(u_i: M\to [0,1],\) such that the support of u i in contained in U i and that

$$\sum_{i\in I} u_i\ \equiv 1,$$

where the sum is locally finite, i.e., for every \(p\in M\) there is a neighborhood U, such that the set

$$\{i\in I: u_i|_U\ne 0\}$$

is finite. The family (u i ) is called a smooth partition of unity subordinate to the covering \((U_i).\) One can choose the u i in a way that for each \(i\in I\) the function \(\sqrt{u_i}\) is smooth as well.

Proof

Except for the smoothness of the square-root, this is Theorem 1.11 of [War83], the proof of which is adapted to give the lemma. In loc.cit., one constructs a sequence \((\psi_j)_{j\ge 1}\) of smooth functions, which are \(\ge 0\) and such that the sets \(\{p:\psi_j(p)>0\}\) form a locally-finite covering of M, subordinate to the covering (U i ). So the function

$$\psi=\sum_{j\ge 1}\psi_j^2$$

is a smooth function with \(\psi(p)>0\) for every \(p\in M\). The functions

$$u_j=\frac{\psi_j^2}{\psi}$$

are smooth with smooth square root and satisfy \(\sum_ju_j=1\). □

A Borel measure ν on \({\mathbb R}^n\) is called a smooth measure , if it has a smooth, strictly positive density with respect to the Lebesgue measure \({\lambda}\), i.e., if there is a smooth function

$$d:{\mathbb R}^n\to (0,\infty)$$

such that

$$\nu(A)=\int_A d(x)\,d{\lambda}(x)$$

holds for every Borel set \(A\subset{\mathbb R}^n\).

A measure μ on a smooth manifold M is called a smooth measure, if for every smooth chart \(\phi:U\to{\mathbb R}^n\) the induced measure \(\nu=\phi_*\mu\) on \({\mathbb R}^n\), given by \(\mu(A)=\nu(\phi^{-1}(A))\) is smooth.

Proposition 9.5.3

Let M be a compact smooth manifold of dimension n with a smooth measure μ. Let \(k:M\times M\to{\mathbb C}\) be continuous and 2r-times continuously differentiable in the first argument, where r is the smallest integer \(r>n/2.\) Then the induced integral operator \(T_k:L^2(M)\to L^2(M)\),

$$T_k(\phi)(x)=\int_Mk(x,y)\phi(y)\,dy,$$

is of trace class and

$$\hspace{1pt}{\rm tr}\hspace{2pt}(T_k)=\int_Mk(x,x)\,dx.$$

Proof

We first prove the assertion in the case \(M={\mathbb R}^n/{\mathbb Z}^n\) and μ being the Lebesgue measure \({\lambda}\). In this case we define

$$l(x,y)=\sum_{k\in{\mathbb Z}^n}\left(\frac1{1+4\pi^2\Vert k\Vert^2}\right)^{r}e_k(x){\overline{e_k(y)}},$$

where \(e_k(x)=e^{2\pi i(x_1k_1+x_2k_2+\dots+x_nk_n)}\) and \(\Vert k\Vert^2=k_1^2+\dots+k_n^2.\) As the corresponding integral converges, this sum converges absolutely uniformly and the kernel \(l(x,y)\) therefore is continuous. Let ∆ be the Laplace operator,

$$\Updelta=-\frac{\partial^2}{\partial x_1^2}-\dots-\frac{\partial^2}{\partial x_n^2}.$$

Lemma 9.5.4

For \(\phi\in C^{2r}({\mathbb R}^n/{\mathbb Z}^n)\) one has

$$T_l(1+\Updelta)^r\phi=\phi.$$

Proof

Both sides are continuous functions, so it suffices to show that they agree as L 2-functions. But for all e k , \(k\in{\mathbb Z}^n\), we have \((1+\Updelta)^re_k=(1+4\pi^2\Vert k\Vert^2)^re_k\) and \(T_le_k=(1+4\pi^2\Vert k\Vert^2)^{-r}e_k\), hence \(T_l(1+\Updelta)^re_k=e_k\). Since \(\{e_k: k\in{\mathbb Z}^n\}\) is an orthonormal basis of \(L^2({\mathbb R}^n/{\mathbb Z}^n)\) we see that \(T_l(1+\Updelta)^r={\rm Id}_{L^2({\mathbb R}^n/{\mathbb Z}^n)}\), and the result follows. □

We now show the proposition in the case \(M={\mathbb R}^n/{\mathbb Z}^n\) equipped with the Lebesgue measure. Let k be as in the theorem. Then for \(\phi\in L^2(M)\) the function \(T_k\phi\) is in \(C^{2r}(M)\), so by the lemma,

$$T_l(1+\Updelta)^rT_k\phi=T_k\phi.$$

One has \((1+\Updelta)^rT_k=T_{k'}\), where \(k'\) is the continuous kernel

$$k'(x,y)=(1+\Updelta_x)^rk(x,y).$$

So \(T_k=T_lT_{k'}\) is a product of two Hilbert-Schmidt operators, hence of trace class. As both, T l and \(T_{k'}\) are admissible, the claim follows from Lemma 9.3.1.

Next let M be an arbitrary smooth compact manifold of dimension n. Let \((U_i)_{i=1}^s\) be an open covering by chart sets, where we choose the chart maps ψ i not as maps to \({\mathbb R}^n\) but instead to \({\mathbb R}^n/{\mathbb Z}^n\), which is possible as well. So for every i the map ψ i is a homeomorphism of U i to some open set \(V_i\subset{\mathbb R}^n/{\mathbb Z}^n\). The sets \(V_1,\dots, V_s\subset{\mathbb R}^n/{\mathbb Z}^n\) can be chosen pairwise disjoint. Let (u i ) be a smooth partition of unity with smooth square root subordinate to the covering (U i ). For \(1\le i,j\le s\) let

$$k_{i,j}(x,y)=\sqrt{u_i(x)}k(x,y)\sqrt{u_j(y)}.$$

Then \(k_{i,j}\) is a continuous kernel on \(M\times M\). For given \(1\le i,j\le s\), define a continuous kernel \(\tilde k_{i,j}\) on \({\mathbb R}^n/{\mathbb Z}^n\times{\mathbb R}^n/{\mathbb Z}^n\) by

$$\tilde k_{i,j}(x,y)=\begin{cases}\sqrt{d_i(x)}k_{i,j}(\psi_i^{-1}(x),\psi_j^{-1}(y))\sqrt{d_j(y)}& (x,y)\in V_i\times V_j,\\ 0&\text{otherwise.}\end{cases}$$

Here d j denotes the density of the measure \((\psi_j)_*\mu\) with respect to the Lebesgue measure, i.e. \(\int_{V_i} f(x) d_i(x)\,d\lambda(x)=\int_{U_i} f(\psi_i(y))d\mu(y)\) for all \(f\in C_c(V_i)\). We define \(\tilde k=\sum_{i,j}\tilde k_{i,j}\) and for \(\phi\in L^2(M)\) we set

$$\phi_j(x)=\phi(x)\sqrt{u_j(x)}.$$

We define \(\tilde\phi_j\in L^2({\mathbb R}^n/{\mathbb Z}^n)\) by

$$\tilde\phi_j(x)=\begin{cases}\phi_j(\psi_j^{-1}(x))\sqrt{d_j(x)}& x\in V_j,\\ 0 &\text{otherwise.}\end{cases}$$

Finally set \(\tilde\phi=\sum_j\tilde\phi_j\).

Lemma 9.5.5

The map \(\Uppsi:\phi\mapsto\tilde\phi\) is a linear isometry \(L^2(M)\hookrightarrow L^2({\mathbb R}^n/{\mathbb Z}^n)\) and one has

$$\Uppsi(T_k\phi)=T_{\tilde k}\Uppsi(\phi)$$

for every \(\phi\in L^2(M).\) The operator \(T_{\tilde k}\) equals \(PT_{\tilde k}P,\) where P is the orthogonal projection \(L^2({\mathbb R}^n/{\mathbb Z}^n)\to{\rm Im}(\Uppsi)\). Finally we have

$$\int_Mk(x,x)\,d\mu(x)=\int_{{\mathbb R}^n/{\mathbb Z}^n}\tilde k(x,x)\,dx.$$

Proof

The map Ψ is linear. For \(\phi\in L^2(M)\) we compute

$$\begin{aligned} \Vert{\Uppsi(\phi)}\Vert^2 &= \int_{{\mathbb R}^n/{\mathbb Z}^n}|\tilde\phi(x)|^2\,d{\lambda}(x)=\sum_j\int_{V_j}|\tilde\phi_j(x)|^2\,d{\lambda}(x)\\ &= \sum_j\int_{V_j}|\phi_j(\overbrace{\psi_j^{-1}(x)\!}^{=z})|^2\underbrace{d_j(x)\,d{\lambda}(x)}_{d(\psi_j)_*\mu(x)}\\ &=\sum_j\int_{U_j}|\phi_j(z)|^2\,d\mu(z)=\sum_j\int_M|\phi(z)|^2u_j(x)\,d\mu(z)\\ &=\int_M|\phi(z)|^2\,d\mu(x)=\Vert\phi\Vert^2,\end{aligned}$$

so Ψ is an isometry. In order to show \(\Uppsi T_k=T_{\tilde k}\Uppsi\), we compute

$$\begin{aligned} T_{\tilde k}\Uppsi(\phi)(x)&=T_{\tilde k}\tilde\phi(x)=\sum_j T_{\tilde k}\tilde\phi_j(x)= \sum_j \int_{V_j}\tilde k(x,y)\tilde\phi_j(y)\,d{\lambda}(y)\\ &=\sum_{i,j}\sqrt{d_i(x)}\int_{V_j}k_{i,j}(\psi_i^{-1}(x),\psi_j^{-1}(y))\phi_j(\psi_j^{-1}(y))\underbrace{d_j(y)\,d{\lambda}(y)}_{=d(\psi_j)_*\mu(y)}\\ &=\sum_{i,j}\sqrt{d_i(x)}\int_{U_j}k_{i,j}(\psi_i^{-1}(x),z)\phi_j(z)\,d\mu(z)\\ &=\sum_{i,j}\sqrt{d_i(x)}\int_{U_j}\sqrt{u_i(\psi_i^{-1}(x))}k(\psi_i^{-1}(x),z)u_j(z)\phi(z)\,d\mu(z)\\ &=\sum_{i}\sqrt{d_i(x)}\sqrt{u_i(\psi_i^{-1}(x))}\int_{M}k(\psi_i^{-1}(x),z)\phi(z)\,d\mu(z)\\ &=\sum_{i}\sqrt{d_i(x)}\sqrt{u_i(\psi_i^{-1}(x))}T_k\phi(\psi_i^{-1}(x)) =\Uppsi T_k\phi(x).\end{aligned}$$

To show \(T_{\tilde k}=PT_{\tilde k}P\) one has to show that for \(g\in{\rm Im}(\Uppsi)^\perp\),

$$T_{\tilde k}(g)=0 \quad\text{and}\quad{\left\langle{T_{\tilde k}h,g}\right\rangle}=0$$

holds for every \(h\in L^2({\mathbb R}^n/{\mathbb Z}^n)\). The first of these assertions follows from the fact that for fixed x, the map \(y\mapsto\tilde k(x,y)\) lies in the image of Ψ. The second assertion follows similarly from the observation, that for fixed y the map \(x\mapsto \tilde k(x,y)\) lies in the image of □.

We finish the proof of the lemma with

$$\begin{aligned} \int_{{\mathbb R}^n/{\mathbb Z}^n}\tilde k(x,x)\,d{\lambda}(x) &= \sum_{i,j}\int_{{\mathbb R}^n/{\mathbb Z}^n} \underbrace{\tilde k_{i,j}(x,x)}_{=0\,\text{if}\,i\ne i}\,d{\lambda}(x)\\ &=\sum_j\int_{V_j} k_{j,j}(\psi_j^{-1}(x),\psi_j^{-1}(x))d_j(x)\,d{\lambda}(x)\\ &=\sum_j\int_{U_j}k_{j,j}(x,x)\,d\mu(x)\\ &=\sum_j\int_{U_j}k(x,x)u_j(x)\,d\mu(x)\\ &=\int_{M}k(x,x)\,d\mu(x). \end{aligned}$$

The lemma implies

$$\begin{aligned} {\hspace{1pt}{\rm tr}\hspace{2pt}} T_k={\hspace{1pt}{\rm tr}\hspace{2pt}} T_{\tilde k}=\int_{{\mathbb R}^n/{\mathbb Z}^n}\tilde k(x,x)\,dx=\int_Mk(x,x)\,dx\end{aligned}$$

and the proposition follows. □

Now Theorem 9.5.1 follows from Proposition 9.5.3 in the same way as Theorem 9.3.2, i.e., the trace formula, follows from Proposition 9.3.1. □

9.6 Exercises

Exercise 9.1

Show that every Schwartz function on \({\mathbb R}\) is uniformly integrable.

Exercise 9.2

Let B be the group of upper triangular matrices in \({\rm SL}_2({\mathbb R})\). Show that B does not contain a lattice.

Exercise 9.3

Let \({\Upgamma}\) be a subgroup of the locally compact group G. Show that \({\Upgamma}\) is discrete if and only if for every compact subset \(C \subset G\) the intersection \({\Upgamma}\cap C\) is finite.

Exercise 9.4

A group \({\Upgamma}\) acts discontinuously on a topological space X if for every \(x\in X\) there is a neighborhood U of x, such that the set of all \({\gamma}\in{\Upgamma}\) with \(U\cap{\gamma} U\ne\emptyset\) is finite.

Let \(K\subset G\) be a compact subgroup of the locally compact group G. Show that a subgroup \({\Upgamma}\subset G\) is discrete if and only if it acts discontinuously on the space \(G/K\).

Exercise 9.5

Let G be a finite group, and let H be a subgroup. For each choose the counting measure as Haar measure. For \(x\in H\) let \(H_x, G_x\) be the centralizers of x in H and G, respectively. For \(\pi\in\widehat G\) let \(\chi_\pi: G\to{\mathbb C}\) be defined by \(\chi_\pi(x)={\hspace{1pt}{\rm tr}\hspace{2pt}}\pi (x)\). For \(x\in H\) let \([x]_H\) be the H-conjugacy class of x, i.e., \([x]_H=\{hxh^{-1}: h\in H\}.\) Show:

$$\sum_{\pi\in\widehat G}({\rm dim} V_\pi^H)\,\chi_\pi=\sum_{[h]_H}\frac{|G_h|}{| H_h|}\,{\bf 1}_{[h]_G},$$

where \(V_\pi^H\) denotes the space of H-invariant vectors in \(V_\pi\).

(Hint: Let g be one of the sides of the equation. For \(f\in L^2(G)\) consider the inner product \({\left\langle{f,g}\right\rangle}\) and use the trace formula.)

Exercise 9.6

Let \(f\in{\cal S}({\mathbb R})\) be a Schwartz function. Apply the trace formula to f with \(G={\mathbb R}\) and \({\Upgamma}={\mathbb Z}\) to give a proof of the classical Poisson summation formula:

$$\sum_{k\in{\mathbb Z}} f(k)=\sum_{k\in{\mathbb Z}} \hat f(k),$$

where \(\hat f(x)=\int_{\mathbb R} f(y)e^{-2\pi ixy}\, dy\).

Exercise 9.7

Let G be a first countable compact group and \(H\subset G\) a closed subgroup. Let \(f\in C(G)^2\) with \({\operatorname{supp}}(f)\cap G.H=\emptyset\), where G.H is the union of all conjugates gHg -1 of H. Show that

$$\sum_{\pi\in\widehat G} {\rm dim}(V_\pi^H)\,{\hspace{1pt}{\rm tr}\hspace{2pt}}\pi(f)=0.$$