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 prove the Stone-von Neumann Theorem, which gives a full characterization of the unitary dual of the Heisenberg group \({\cal H}\). We then apply the trace formula to describe the spectral decomposition of \({L^2}(\Lambda \backslash H)\), where π is the standard integer lattice in \({\cal H}\).

10.1 Definition

The Heisenberg group \({\cal H}\) is defined to be the group of real upper triangular 3 × 3 matrices with ones on the diagonal:

$$\cal H{\ \stackrel{{\rm def}}{=}} \left\{\left. \left(\begin{array}{ccc} 1 & x & z \\ & 1 & y \\ & & 1\end{array}\right) \right| x,y,z\in{\mathbb R}\right\}.$$

It can also be identified with \({\mathbb R}^3\), where the group law is given by

$$(a,b,c)(x,y,z){\ \stackrel{{\rm def}}{=}\,} (a+x,b+y,c+z+ay).$$

The inverse of \((a,b,c)\) is

$$(a,b,c)^{-1}= (\!\!-\!a,-b,ab-c).$$

The center of \({\cal H}\) is \(Z({\cal H})=\{(0,0,z)\mid z\in{\mathbb R}\}\) , and the projection to the first two coordinates induces an isomorphism

$$\cal H/Z({\cal H})\ \cong\ {\mathbb R}^2.$$

An easy calculation ([Dei05] Chap. 12) shows that \({\cal H}\) is unimodular and that a Haar integral on \({\cal H}\) is given by

$$\int_{\cal H} f(h)\, dh{\ \stackrel{{\rm def}}{=}} \int_{\mathbb R}\int_{\mathbb R}\int_{\mathbb R} f(a,b,c)\, da\, db\, dc,\ \ f\in C_c({\cal H}).$$

We will use this Haar measure on \({\cal H}\) for all computations in the sequel.

10.2 The Unitary Dual

In this section we are going to describe the unitary dual \(\widehat{\cal H}\) of the Heisenberg group \({\cal H}\).

Let \(\widehat{\cal H}_0\) denote the subset of \(\widehat{\cal H}\) consisting of all classes \(\pi\in\widehat{\cal H}\) such that \(\pi(h)=1\) whenever h lies in the center \(Z({\cal H})\) of \({\cal H}\). Since \({\cal H}/Z ({\cal H})\cong{\mathbb R}^2\), it follows that

$$\widehat{\cal H}_0=\widehat{{\cal H}/Z({\cal H})}\ \cong\ \widehat{{\mathbb R}^2}\ \cong\ \left(\widehat{\mathbb R}\right)^2,$$

and the latter can be identified with \({\mathbb R}^2\) in the following explicit way. Let \((a,b)\in{\mathbb R}^2\) and define a character

$$\begin{array}{cccl} \chi_{a,b}: & H & \to & {\mathbb T},\\ & (x,y,z) &\mapsto & e^{2\pi\! i(ax+by)}.\end{array}$$

The identification is given by \((a,b)\mapsto \chi_{a,b}\). In particular, it follows that all representations in \(\widehat{\cal H}_0\) are one-dimensional. This observation indicates the importance of the behavior of the center under a representation.

As a consequence of the Lemma of Schur 6.1.7, for each \(\pi\in\widehat{\cal H}\) there is a character \(\chi_\pi: Z({\cal H})\to{\mathbb T}\) with \(\pi(z)=\chi_\pi(z){\rm Id}\) for every \(z\in Z({\cal H})\). This character \(\chi_\pi\) is called the central character of the representation π.

For every character \(\chi\ne 1\) of \(Z({\cal H})\), we will now construct an irreducible unitary representation of the Heisenberg group that has χ for its central character. So let \(t\ne 0\) be a real number and consider the central character

$$\chi_t(0,0,c)= e^{2\pi\! i ct}.$$

For \((a,b,c)\in{\cal H}\) we define the operator \(\pi_t(a,b,c)\) on \(L^2({\mathbb R})\) by

$$\pi_t(a,b,c)\phi(x){\ \stackrel{{\rm def}}{=}\,} e^{2\pi\! i (bx+c)t} \phi(x+a).$$

It is straightforward that π t is a unitary representation.

Recall from Exercise 3.17 that the Schwartz space \({\cal S}({\mathbb R}^n)\) consists of all \(C^\infty\)-functions \(f:{\mathbb R}^n\to{\mathbb C}\) such that \(x^\alpha\partial^\beta f\) is bounded for all multi-indices \(\alpha,\beta\in{\mathbb N}_0^n\). We have \({\cal S}({\mathbb R}^n)\subseteq L^p({\mathbb R}^n)\) for every \(p\geq 1\) and \({\cal S}({\mathbb R}^n)\) is stable under Fourier transform.

Theorem 10.2.1

(Stone-von Neumann). For \(t\ne 0\) the unitary representation π t is irreducible. Every irreducible unitary representation of \({\cal H}\) with central character \(\chi_t\) is isomorphic to π t . It follows that

$$\widehat{\cal H}=\widehat{\mathbb R}^2\ \cup\ \{\pi_t: t\ne 0\}.$$

Proof

We show irreducibility first. Fix \(t\ne 0\), and let \(V\subset L^2({\mathbb R})\) be a closed non-zero subspace that is invariant under the set of operators \(\pi_t({\cal H})\). If \(\phi\in V\), then so is the function \(\pi_t(\!\!-\!a,0,0)\phi(x)=\phi(x-a)\). As V is closed, it therefore contains \(\psi *\phi(x)=\int_{\mathbb R}\psi(a)\phi(x- a)\,da\) for \(\psi\in{\cal S}={\cal S}({\mathbb R})\). These convolution products are smooth functions, so V contains a smooth function \(\phi\ne 0\). One has \(\pi_t(0,-b,0)\phi(x)= e^{-2\pi\! ibtx}\phi(x)\ \in\ V.\) By integration it follows that for \(\psi\in{\cal S}\) one has that \(\hat\psi(tx) \phi(x)\) lies in V. The set of possible functions \(\hat\psi(tx)\) contains all smooth functions of compact support, as the Fourier transform is a bijection on the space of Schwartz functions to itself. Choose an open interval I, in which φ has no zero. It follows that \(C_c^\infty(I)\subset V\). Translating, taking sums, and using a partition of unity argument gives \(C_c^\infty({\mathbb R})\subset V\). This space is dense in \(L^2({\mathbb R})\), so \(V=L^2({\mathbb R})\), which means that π t is irreducible.

For the second assertion of the theorem, let π be an irreducible unitary representation with central character χ t . We want to show that π is isomorphic to π t . For notational ease, let’s first reduce to the case t = 1. Consider the map \(\theta_t:{\cal H}\to{\cal H}\) given by \(\theta_t(a,b,c)=(a,bt,ct)\). A calculation shows that θ t is an automorphism of \({\cal H}\) with \(\chi_1\circ\theta_t=\chi_t\). So we conclude that π has central character χ t if and only if \(\pi\circ\theta_t^{-1}\) has central character χ1. It therefore suffices to show the uniqueness for t = 1.

Let \((\eta,V_\eta)\) be an irreducible unitary representation of \({\cal H}\) with central character χ1. Let \({\Gamma}\) be the subgroup of \({\cal H}\) consisting of all \((0,0,k)\) with \(k\in{\mathbb Z}\). Then \({\Gamma}\) lies in the center and \(\chi_1({\Gamma})=\{1\}\). So the representation η factors over the quotient group \(B{\ \stackrel{{\rm def}}{=}} {\cal H}/{\Gamma},\) which is homeomorphic to the space \({\mathbb R}^2\times{\mathbb R}/{\mathbb Z}\). For \(\phi\in{\cal S}({\mathbb R}^2)\), let \(\phi_B(a,b,c)=\phi(a,b)e^{-2\pi\! ic}\). Then \(\phi(a,b)=\phi_B(a,b,0)\). For \(\phi,\psi\in{\cal S}({\mathbb R}^2)\), one computes

$$\begin{aligned} \hspace{-20pt}\phi_B *\psi_B (a,b,c)&= \int_B\phi_B(h)\psi_B(h^{-1}(a,b,c)) \,dh\\ &= \int_{{\mathbb R}^2}\int_{{\mathbb R}/{\mathbb Z}}\phi(x,y)e^{-2\pi\! iz}\psi_B(a-x,b-y,c-z+xy-xb) \,dx\,dy\, dz\\ &= \int_{{\mathbb R}^2} \int_{{\mathbb R}/{\mathbb Z}}\phi(x,y)e^{-2\pi\! iz}\psi(a-x,b-y)e^{-2\pi\! i(c-z +xy-xb)}\,dx\,dy\,dz\\ &= e^{-2\pi\! ic} \int_{{\mathbb R}^2} \phi(x,y)\psi(a-x,b-y)e^{2\pi\! i x(b-y)}\,dx \,dy.\end{aligned}$$

Inspired by this, we define a new, non-commutative convolution product on \({\cal S}({\mathbb R}^2)\) by

$$\phi *_B\psi (a,b) {\ \stackrel{{\rm def}}{=}} \int_{{\mathbb R}^2} \phi(x,y)\psi(a-x,b-y)e^{2\pi\! i x(b-y)} \,dx\,dy.$$

For \(\phi\in{\cal S}({\mathbb R}^2)\) we define \(\eta(\phi)=\eta(\phi_B)\). This equals

$$\eta(\phi)=\int_{{\mathbb R}^2}\phi(a,b)\eta(a,b,0)\,da\,db.$$

By construction of the convolution product one gets \(\eta(\phi *_B\psi)=\eta (\phi)\eta(\psi)\), so η is an algebra homomorphism of the algebra \({\cal A}= ({\cal S}({\mathbb R}^2),*_B)\). One computes that \((\phi_B)^*=(\phi^*)_B\), where the involution on the left is the usual involution of functions g on the group B, given by \(g^*(b) ={\overline{g(b^{-1})}}\). The involution on the algebra \({\cal A}={\cal S}({\mathbb R}^2)\) on the other hand is defined by \(\phi^*(a,b)={\overline{\phi(\!\!-\!a,-b)}}e^{2\pi\! iab}\). So η is a homomorphism of *-algebras.

Lemma 10.2.2

The *-algebra homomorphism \(\eta: {\cal A}\to{\cal B}(V_\eta)\) is injective.

Proof

Assume \(\eta(\phi)=0\). Then, for \(h=(a,b,0)\in B\) and \(v,w\in V_\eta\) we have

$$\begin{aligned} 0&= {\left\langle{\eta(h)\eta(\phi)\eta(h^{-1})v,w}\right\rangle}\\ &= \int_{{\mathbb R}^2}{\left\langle{\eta\, \textit{(x,\;y,\;0)\;v,\;w}}\right\rangle}\phi(x,y)e^{2\pi\! i(ay-bx)}\,dx\,dy\end{aligned}$$

The continuous, rapidly decreasing function \({\left\langle{\eta(x,y,0)v,w}\right\rangle}\phi(x,y)\) therefore has zero Fourier transform, hence is zero. Varying v and w, this leads to φ being zero. □

In the special case \(\eta=\pi_1\) the lemma has the consequence that relations among the \(\pi_1(\phi)\) for \(\phi\in{\cal A}\) already hold in \({\cal A}\). We will make use of this principle for special elements as follows. For \(F,G\in{\cal S}({\mathbb R})\) let \(\tilde\phi_{F,G}(a,b)= {\overline{F(a+b)}}G(b),\) and \(\phi_{F,G}(a,b)={\cal F}_2\tilde\phi_{F,G}(a,b)\ \in\ {\cal S}({\mathbb R}^2),\) where \({\cal F}_2\) denotes taking Fourier transform in the second variable. An application of the properties of the Fourier transform yields

$$\begin{aligned} \pi_1(\phi_{F,G})f(x) &= \int_{{\mathbb R}^2} \phi_{F,G}(a,b)\pi_1(a,b,0)f(x)\,da\,db\\ &= \int_{{\mathbb R}^2} \phi_{F,G}(a,b)f(x+a)e^{2\pi\! i bx}\,da\,db\\ &= \int_{{\mathbb R}^2} {\cal F}_2\tilde\phi_{F,G}(a-x,b)f(a)e^{2\pi\! i bx}\,da\,db\\ &=\int_{\mathbb R}\tilde\phi_{F,G}(a-x,x)f(a)\,da\\ &= \int_{\mathbb R} {\overline{F(a)}}G(x)f(a)\,da= {\left\langle{f,F}\right\rangle}G(x).\end{aligned}$$

In particular, if the norm of F is one, then \(\pi_1(\phi_{F,F})\) is the orthogonal projection onto the one dimensional space \({\mathbb C} F\).

For \(h\in B={\cal H}/{\Gamma}\) and \(\phi\in{\cal S}({\mathbb R}^2)\) define \(L_h\phi\) by the formula \(\left(L_h\phi\right) _B=L_h(\phi_B)\). Explicitly, one computes that \(L_{(x,y,z)}\phi(a,b)= \phi(a-x,b-y)e^{2\pi\! i(bx-xy+z)}.\) For \(\phi,\psi\in{\cal A}\) we also write \(\phi\psi\) for the convolution product \(\phi *_B\psi\).

Lemma 10.2.3

In the algebra \({\cal A}\) , one has the relations

$$\phi_{F,G}\phi_{H,L}={\left\langle{L,F}\right\rangle}\phi_{H,G}\ {\rm and}\ \phi_{F,G}^*=\phi_{G,F}.$$

For \(h\in B\) one has \(L_h\phi_{F,G}=\phi_{F,\pi_1(h)G}\).

Proof

Using the formula \(\pi_1(\phi_{F,G})f= {\left\langle{f,F}\right\rangle}G\) one computes for \(f\in L^2({\mathbb R})\),

$$\begin{aligned} \pi_1(\phi_{F,G}\phi_{H,L})f &= \pi_1(\phi_{F,G})\pi_1(\phi_{H,L})f\\ &= {\langle{\pi_1(\phi_{H,L})f,F}\rangle}G\\ &= {\left\langle{{\left\langle{f,H}\right\rangle}L,F}\right\rangle}G= {\left\langle{L,F}\right\rangle}{\left\langle{f,H}\right\rangle}G.\end{aligned}$$

This implies the first claim as π1 is injective. The second follows from

$$\langle{\pi_1(\phi_{F,G})f\!,h}\rangle={\left\langle{f,F}\right\rangle}{\left\langle{G,h}\right\rangle}={\langle{f,\pi_1(\phi_{G,F})h}\rangle}.$$

For the last claim compute

$$\begin{aligned} \pi_1(L_h\phi_{F,G})f &= \pi_1(h)\pi_1(\phi_{F,G})f= \pi_1(h){\left\langle{f\!,F}\right\rangle}G\\ &= {\left\langle{f\!,F}\right\rangle}\pi_1(h)G=\pi_1(\phi_{F,\pi_1(h)G})f.\end{aligned}$$

The lemma follows. □

As \({\cal S}({\mathbb R})\) is dense in \(L^2({\mathbb R})\), by the orthonormalizing scheme one can find a sequence F j in \({\cal S}({\mathbb R})\) that forms an orthonormal base of \(L^2({\mathbb R})\). For \(j,k\in{\mathbb N}\) write \(\phi_{j,k}\) for \(\phi_{F_j,F_k}\). Lemma 10.2.3 implies that \(\phi_{j,k}\phi_{s,t}=\delta_{j,t}\phi_{s,k}\) (Kronecker-delta), and \(\phi_{j,j}^2=\phi_{j,j}=\phi_{j,j}^*\).

Let now \((\eta,V_\eta)\) again denote an irreducible unitary representation with central character χ1. By integration, η gives an algebra homomorphism \({\cal A}\to{\cal B}(V_\eta)\).

Lemma 10.2.4

The \(\eta(\phi_{j,j})_{j\in{\mathbb N}}\) are non-zero projections. They are pairwise orthogonal. Write V j for the image of \(\eta(\phi_{j,j})\). Then \(\eta(\phi_{j,k})\) is an isometry from V j to V k and it annihilates every V l for \(l\ne j\).

Proof

Since \(\phi_{j,j}^2=\phi_{j,j}=\phi_{j,j}^*\), the same holds for \(\eta(\phi_{j,j})\) hence the latter are projections. They are non-zero by Lemma 10.2.2. For \(v\in V_j\) one computes \({\langle{\eta(\phi_{j,k})v,\eta(\phi_{j,k})v}\rangle}= {\langle{\eta(\phi_{k,j}\phi_{j,k})v,w}\rangle} = {\langle{\eta(\phi_{j,j})v,w}\rangle}={\langle{v,v}\rangle}.\) The claim follows. □

Now choose \(v_1\in V_1\) of norm one and set \(v_j=\eta(\phi_{1,j})v_1\). Then (v j ) is an orthonormal system in \(V_\eta\). Define an isometry \(T: L^2({\mathbb R})\to V_\eta\) by mapping F j to v j . Then

$$\begin{aligned} \eta(h)T(F_j)&= \eta(h)v_j= \eta(h)\eta(\phi_{j,j})v_j\\ &= \eta(L_h\phi_{j,j})v_j=\eta(\phi_{F_j,\pi_1(h)F_j})v_j.\end{aligned}$$

By the fact that (F j ) is an orthonormal basis, one concludes

$$\pi_1(h)F_j=\sum_k{\langle{\pi_1(h)F_j,F_k}\rangle}F_k.$$

So that by Lemma 10.2.3

$$\begin{aligned} \eta(h)T(F_j)&= \sum_k{\langle{\pi_1(h)F_j,F_k}\rangle}\eta(\phi_{F_j,F_k})v_j\\ &=\sum_k{\langle{\pi_1(h)F_j,F_k}\rangle}v_k= T(\pi_1(h)F_j).\end{aligned}$$

Hence T is an \({\cal H}\)-homomorphism onto a non-zero closed subspace of \(V_\eta\). As η is irreducible, T must be surjective, so T is unitary and \(\eta\) is equivalent to π1. This finishes the proof of the Theorem of Stone and von Neumann. □

10.3 The Plancherel Theorem for \({\cal H}\)

We have an identification \({\cal H}\cong{\mathbb R}^3\). We interpret \(f\in{\cal S}({\mathbb R}^3)\) as a function on \({\cal H}\), and we write \({\cal S}({\cal H})\) for this space of functions.

Theorem 10.3.1

(Plancherel Theorem). Let \(f\in{\cal S}({\cal H})\). For every \(t\in{\mathbb R}^\times\) the operator \(\pi_t(f)\) is a Hilbert-Schmidt operator, and we have

$$\int_{{\mathbb R}^\times} {\left|\hspace{-1pt}\left| {\pi_t(f)}\right|\hspace{-1pt} \right|}_{HS}^2\, |t|\, dt=\int_{\cal H} |f(h)|^2\, dh.$$

It follows that the Plancherel measure of \({\cal H}\) in the sense of Theorem 8.5.3 equals \(|t|\,dt\) and that the set of one dimensional representations of \({\cal H}\) has Plancherel measure zero.

Proof

This is proved in [Dei05], Chap. 12. □

10.4 The Standard Lattice

Let \({\Lambda}\) be the set of all \((a,b,c)\in{\cal H}\), where \(a,b,c\) are integers. The multiplication and inversion keep this set stable, so \({\Lambda}\) is a discrete subgroup.

Lemma 10.4.1

\({\Lambda}\) is a uniform lattice in \({\cal H}\).

Proof

Let \(K=[0,1]\times[0,1]\times[0,1]\subset{\cal H}\). Then K is a compact subset. We claim that the projection map \(K\to{\cal H}/{\Lambda}\) is surjective. This means we have to show that for every \(h\in{\cal H}\) there exists \({\lambda}\in{\Lambda}\) such that \(h{\lambda}\in K\). Let \(h=(x,y,z)\in{\cal H}\), and let \({\lambda}=(a,b,c)\in{\Lambda}\). Then \(h{\lambda}=(x+a,y+b,z+c+xb)\). So we can find \(a,b\in{\mathbb Z}\) such that \(x+a\) and \(y+b\) lie in the unit interval. So we may assume as well that \(x,y\in [0,1]\). Assuming that, we only consider such \({\lambda}\), which are central, i.e., of the form \({\lambda}=(0,0,c)\) and it becomes clear that we can find \({\lambda}\) such that \(h {\lambda}\in K\). □

By Theorem 9.2.2 we conclude that as \({\cal H}\)-representation one has the decomposition

$$L^2({\Lambda}{\backslash}{\cal H})\ \cong\ \bigoplus_{\pi\in\widehat{\cal H}}N(\pi)\pi,$$

with finite multiplicities \(N(\pi)\).

We want to apply the trace formula to compute the numbers \(N(\pi)\). For this recall the characters of \({\cal H}\) given by \(\chi_{r,s}(a,b,c)=e^{2\pi\! i(ra+sb)}\), where \(r,s\in{\mathbb R}\).

Theorem 10.4.2

For a character \(\chi_{r,s}\) of \({\cal H}\) the multiplicity \(N(\chi_{r,s})\) in \(L^2({\Lambda}{\backslash} {\cal H})\) is equal to one if r and s are both integers. Otherwise it is zero. For \(t\in{\mathbb R}\), \(t\ne 0\), the multiplicity \(N(\pi_t)\) is \(|t|\) if t is an integer and zero otherwise. So the decomposition of the representation \(\left(L^2({\Lambda}{\backslash} {\cal H}), R\right)\) reads

$$R\ \cong\ \bigoplus_{r,s\in{\mathbb Z}}\chi_{r,s}\ \oplus\ \bigoplus_{^{k\in{\mathbb Z}}_{k\ne 0}}|k|\pi_k.$$

Proof

We first consider the subspace H of \(L^2({\Lambda}{\backslash} {\cal H})\), which is invariant under the action of the center \(Z\subset{\cal H}\). This is isomorphic to \(L^2\left({\mathbb Z}^2{\backslash} {\mathbb R}^2\right)\) as \({\Lambda}{\backslash} {\cal H}/Z\cong{\mathbb Z}^2{\backslash} {\mathbb R}^2\) and the representation of \({\cal H}\) on H factors through the representation of \({\mathbb R}^2\) on \(L^2\left({\mathbb Z}^2{\backslash} {\mathbb R}^2\right)\). By the theory of Fourier series this gives us the value of \(N\left(\chi_{r,s}\right)\) as in the theorem. The difficult part is to determine \(N(\pi_t)\).

Let \(h\in C_c^\infty({\cal H})\) and set \(f=h*h^*\). Then the trace formula says,

$$\hspace{1pt}{\rm tr}\hspace{2pt} R(f)=\sum_{\pi\in\widehat{\cal H}}N(\pi){\hspace{1pt}{\rm tr}\hspace{2pt}}\pi(f)=\sum_{[{\lambda}]}{\rm vol}\left({\Lambda}_{\lambda}{\backslash}{\cal H}_{\lambda}\right)\, {\cal O}_{\lambda}(f),$$

where the sum is taken over all conjugacy classes \([{\lambda}]\) in \({\Lambda}\), \({\Lambda}_{\lambda}\) and \({\cal H}_{\lambda}\) denote the centralizers of \({\lambda}\) in \({\Lambda}\) and \({\cal H}\), respectively, and \({\cal O}_{\lambda}(f)=\int_{{\cal H}_{\lambda}{\backslash}{\cal H}} f(x^{-1}{\lambda} x)\, dx\). First consider \(\pi\in\widehat{\cal H}\) with trivial central character, say \(\pi=\chi_ r,s\). Then

$$\begin{aligned} {\hspace{1pt}{\rm tr}\hspace{2pt}}\pi(f)&=\chi_{r,s}(f)=\int_{\mathbb R}\int_{\mathbb R}\int_{\mathbb R} f(a,b,c) e^{2\pi\! i (ar+bs)}\,da\,db\,dc\\ &={\cal F}_1{\cal F}_2{\cal F}_3f(\!\!-\!\!r,-s,0),\end{aligned}$$

where \({\cal F}_i\) denotes Fourier transform on \({\mathbb R}\) applied to the i-th component of \({\cal H}\). Next the center \(Z_{\Lambda}\) of \({\Lambda}\), the set of all \((0,0,k)\) for \(k\in{\mathbb Z}\), is contained in the center of \({\cal H}\) and therefore acts trivially on \(L^2({\Lambda} {\backslash}{\cal H})\). This implies that for \(\pi\in\widehat{\cal H}\) one has that \(N(\pi)\ne 0\) implies \(\pi(Z_{\Lambda})=\{1\}\). Therefore, \(N(\pi_t)\ne 0\) implies that \(t\in{\mathbb Z}\). For \(t=k\in{\mathbb Z}\) one computes

$$\begin{aligned} \pi_k(f)\phi(x) &= \int_{\mathbb R}\int_{\mathbb R}\left(\int_{\mathbb R} f(a,b,c)e^{2\pi\! ikc}\,dc \right) e^{2\pi\! ikbx}\phi(a+x)\, da\,db\\ &= \int_{\mathbb R} {\cal F}_2{\cal F}_3f(a-x,-kx,-k)\phi(a)\,da\end{aligned}$$

So \(\pi_k(f)\) is an integral operator with kernel

$$k(x,y)={\cal F}_2{\cal F}_3f(y-x,-kx,-k).$$

Analogously, the kernel of \(\pi_k(h)\) is \({\cal F}_2{\cal F}_3h(y-x,-kx,-k)\). The latter kernel is in \({\cal S}({\mathbb R}^2)\) and therefore is admissible. By Proposition 9.3.1 and Fourier inversion, we have

$$\hspace{1pt}{\rm tr}\hspace{2pt}\pi_k(f)=\int_{\mathbb R}{\cal F}_2{\cal F}_3f(0,-kx,-k)\,dx= \frac 1{|k|}{\cal F}_3f(0,0,-k).$$

Together we get

$$\hspace{1pt}{\rm tr}\hspace{2pt} R(f)= \sum_{r,s\in{\mathbb Z}}{\cal F} f(r,s,0)+\sum_{k\in{\mathbb Z}{\smallsetminus}\{0\}} \frac{N(\pi_k)}{|k|}{\cal F}_3f(0,0,k),$$

where \({\cal F} f={\cal F}_1{\cal F}_2{\cal F}_3 f\).

Next we consider the geometric side of the trace formula. Let \({\lambda}\in{\Lambda}\). First assume that \({\lambda}\) lies in the center of \({\Lambda}\). Then \({\lambda}=(0,0,k)\) for some \(k\in{\mathbb Z}\), and the centralizer \({\cal H}_{\lambda}\) equals \({\cal H}\), so that the orbital integral equals \({\cal O}_{\lambda}(f)= f({\lambda})= f(0,0,k).\) Next let \({\lambda}=(r,s,t)\in{\Lambda}={\mathbb Z}^3\) with \((r,s)\ne (0,0)\). Let \(h=(a,b,c)\in{\cal H}\), then \(h{\lambda} h^{-1}=(r,s,t+as-br).\) So h lies in the centralizer \({\cal H}_{\lambda}\) if and only if as = br. It follows that \({\cal H}_{\lambda}\) is the subspace generated by the vectors \((r,s,0)\) and \((0,0,c)\) as a subspace of \({\mathbb R}^3\). On \({\cal H}_{\lambda}\) we pick the Haar measure that comes from the euclidean length in \({\mathbb R}^3\); then the volume of \({\Lambda}_{\lambda}{\backslash}{\cal H}_{\lambda}\) is the minimal value \(\sqrt{a^2+b^2}\) where \(a,b\in{\mathbb Z}\), not both zero, and as = br. This minimum is taken in \(a=\frac{r}{{\rm gcd}(r,s)}\) and \(b=\frac{s}{{\rm gcd}(r,s)}\), where \({\rm gcd}(r,s)\) denotes the greatest common divisor of r and s. Therefore,

$$\rm vol({\Lambda}_{\lambda}{\backslash}{\cal H}_{\lambda})=\frac{\sqrt{r^2+s^2}}{{\rm gcd}(r,s)}.$$

The orbital integral equals

$$\cal O_{\lambda}(f)=\int_{U^\perp}f(r,s,t+xs-yr)\,d(x,y),$$

where U is the subspace of all \(\left(\begin{smallmatrix}a\\ b\end{smallmatrix} \right)\in{\mathbb R}^2\) with as-br = 0. Then \(U={\mathbb R}\left(\begin{smallmatrix}r \\ s\end{smallmatrix}\right)\) and the orthogonal space is spanned by the norm one vector \(\frac 1{\sqrt{r^2+s^2}}\left(\begin{smallmatrix}s\\ -r\end{smallmatrix}\right).\) So we get

$$\cal O_{\lambda}(f) = \int_{\mathbb R} f\left(r,s,t+x\frac{s^2+r^2}{\sqrt{r^2+s^2}}\right)\,dx= \frac 1{\sqrt{r^2+s^2}}\,{\cal F}_3f(r,s,0).$$

The number of conjugacy classes with given \((r,s,*)\) equals \({\rm gcd}(r,s)\), hence

$$\hspace{1pt}{\rm tr}\hspace{2pt} R(f)=\sum_{k\in{\mathbb Z}}f(0,0,k) +\sum_{(r,s)\ne (0,0)}{\cal F}_3f(r,s,0).$$

Using the Poisson summation formula in the first two variables we get

$$\sum_{(r,s)\in{\mathbb Z}^2}{\cal F}_3f(r,s,0)=\sum_{(r,s)\in{\mathbb Z}^2}{\cal F} f(r,s,0)$$

with \({\cal F}={\cal F}_1{\cal F}_2{\cal F}_3\) and using the formula in the third variable gives

$$\sum_{k\in{\mathbb Z}}f(0,0,k)=\sum_{k\in{\mathbb Z}}{\cal F}_3 f(0,0,k).$$

Combining this with the two expressions for \({\hspace{1pt}{\rm tr}\hspace{2pt}} R(f)\) we conclude

$$\sum_{k\ne 0}\frac{N(\pi_k)}{|k|}{\cal F}_3f(0,0,k)=\sum_{k\ne 0}{\cal F}_3 f(0,0,k).$$

By Lemma 9.3.5 this implies \(N(\pi_k)=|k|\). The theorem is proven. □

10.5 Exercises and Notes

Exercise 10.1

Let \({\cal H}\) be the Heisenberg group, and let \(\cal Z\) be its center. Show that every normal subgroup of \({\cal H}\) lies in the center \({\cal Z}\) or contains the center.

Exercise 10.2

On \({\mathbb R}^2\) consider the bilinear form

$$\begin{aligned} b(v,w) \stackrel{\rm def}{=}\, v^t\left(\begin{array}{@{}rr@{}}0 & 1 \\ -1 & 0\end{array} \right)\,w.\end{aligned}$$

Let L be the group \({\mathbb R}^2\times{\mathbb R}\) with the multiplication \((v,t)(w,s){\ \stackrel{{\rm def}}{=}\,} (v+w,s+t+\frac 12 b(v,w)).\) Show: The map \(\psi:{\cal H}\to L\), defined by \(\psi(a,b,c)= ((a,b)^t,c-\frac 12 ab)\) is a continuous group isomorphism with continuous inverse.

Exercise 10.3

Let L be defined as in Exercise 10.2. The group \(G={\rm SL}_2({\mathbb R})\) acts on L via \(g(v,t)=(gv,t)\). Show that G acts by group homomorphisms, which fix the center of L point-wise. Conclude that for every \(g\in G\) one has \(\pi_1\circ g\cong\pi_1\). Conclude that for every \(g\in G\) there is a \(T(g)\in U(L^2({\mathbb R}))\), which is unique up to scalar multiplication, such that \(\pi_1(gl)= T(g)\pi_1(l) T(g)^{-1},\) where we consider π1 as a representation of L via Exercise 10.2. Show that the map \(g\mapsto T(g)\) is a group homomorphism \(G\to U(L^2({\mathbb R}))/{\mathbb T}\).

Exercise 10.4

Let \({\rm Aut}({\cal H})\) be the group of all continuous group automorphisms of \({\cal H}\). Every \(\phi\in{\rm Aut}({\cal H})\) preserves the center \({\cal Z}\), hence induces an element of \({\rm Aut}({\cal H}/{\cal Z})={\rm Aut}({\mathbb R}^2)={\rm GL}_2({\mathbb R})\). Show that the ensuing map \({\rm Aut}({\cal H})\to{\rm GL}_2({\mathbb R})\) is surjective.

Exercise 10.5

For \(h\in{\cal H}\) let \(\phi_h\in{\rm Aut}({\cal H})\) be defined by \(\phi_h(x)=hxh^{-1}\). Every such automorphism is called an inner automorphism . Let \({\rm Inn}({\cal H})\) be the group of all inner automorphisms. Show that \({\rm Inn}({\cal H})\cong{\mathbb R}^2\), and that, together with the last exercise, one gets an exact sequence

$$1\to{\mathbb R}^2\to{\rm Aut}({\cal H})\to{\rm GL}_2({\mathbb R})\to 1.$$

Exercise 10.6

Let \(N\subset{\rm GL}_n({\mathbb R})\) be the group of all upper triangular matrices with ones on the diagonal. Show that \({\Gamma}=N\cap{\rm GL}_n({\mathbb Z})\) is a uniform lattice in N.

Exercise 10.7

The \((2n+1)\)-dimensional Heisenberg group \({\cal H}_n\) is defined as the group of all matrices of the form \(\left(\begin{smallmatrix}1 & x^t & z \\ & 1 & y \\ & & 1\end{smallmatrix}\right)\) in \({\rm GL}_{n+2}({\mathbb R})\) where \(x,y\in{\mathbb R}^n\) and \(z\in{\mathbb R}\). Determine its unitary dual along the lines of the Stone-von Neumann Theorem.

Exercise 10.8

Let L be as in Exercise 10.2. Let \(\Lambda\subset{\mathbb R}^2\) be a lattice with \(b({\Lambda}\times{\Lambda})\subset 2{\mathbb Z}\). Show that \({\Lambda}\times{\Lambda}\times{\mathbb Z}\) is a uniform lattice in L. Use the proof of Theorem 10.4.2 to find the decomposition of \(L^2({\Lambda}{\backslash} L)\).

10.5.1 Notes

As shown in Exercise 10.3, the Stone-von Neumann Theorem yields a group homomorphism \({\rm SL}_2({\mathbb R})\to U(L^2({\mathbb R}))/{\mathbb T}\). There is a unique non-trivial covering group \({\rm SL}_2^2({\mathbb R})\) of degree 2. On this group the homomorphism induces a proper unitary representation \({\rm SL}_2^2({\mathbb R})\to U(L^2({\mathbb R}))\). This representation is known as the Weil representation . It is used to explain the behavior of theta-series, in particular with respect to lifting of automorphic forms [How79, LV80, Wei64].