This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Constructing multi-cusped hyperbolic manifolds that are isospectral and not isometric

Benjamin Linowitz Department of Mathematics
10 North Professor Street
Oberlin, OH 44074
[email protected]
Abstract.

In a recent paper Garoufalidis and Reid constructed pairs of 11-cusped hyperbolic 33-manifolds which are isospectral but not isometric. In this paper we extend this work to the multi-cusped setting by constructing isospectral but not isometric hyperbolic 33-manifolds with arbitrarily many cusps. The manifolds we construct have the same Eisenstein series, the same infinite discrete spectrum and the same complex length spectrum. Our construction makes crucial use of Sunada’s method and the Strong Approximation Theorem of Nori and Weisfeiler.

1. Introduction

In 1966 Kac [10] famously asked “Can one hear the shape of a drum?” In other words, can one deduce the shape of a planar domain given knowledge of the frequencies at which it resonates? Long before Kac had posed his question mathematicians had considered analogous problems in more general settings and sought to determine the extent to which the geometry and topology of a Riemannian manifold is determined by its Laplace eigenvalue spectrum.

Early constructions of isospectral non-isometric manifolds include 1616-dimensional flat tori (Milnor [13]), compact Riemann surfaces (Vignéras [22]) and lens spaces (Ikeda [9]). For an excellent survey of the long history of the construction of isospectral non-isometric manifolds we refer the reader to [7].

In this paper we consider a problem posed by Gordon, Perry and Schueth [8, Problem 1.2]: to construct complete, non-compact manifolds that are isospectral and non-isometric. This problem has received a great deal of attention in the case of surfaces. For example, Brooks and Davidovich [1] were able to use Sunada’s method [18] in order to construct a number of examples of isospectral non-isometric hyperbolic 22-orbifolds. For more examples, see [8].

In a recent paper Garoufalidis and Reid [4] constructed the first known examples of isospectral non-isometric 11-cusped hyperbolic 33-manifolds. The main result of this paper extends the work of Garoufalidis and Reid to the multi-cusped setting.

Theorem 1.1.

There exist finite volume orientable nn-cusped hyperbolic 33-manifolds that are isopectral and not isometric for arbitrarily large positive integers nn.

Moreover, the manifolds we construct will be shown to have the same Eisenstein series, the same infinite discrete spectrum and the same complex length spectrum.

The author would like to thank Dubi Kelmer, Emilio Lauret, Ben McReynolds, Djordje Milićević, Alan Reid and Ralf Spatzier for useful conversations concerning the material in this paper. The author is especially indebted to Jeff Meyer for his close reading of this paper and his many suggestions and comments. The work of the author is partially supported by NSF Grant Number DMS-1905437.

2. Preliminaries

Given a positive integer d2d\geq 2 we define 𝐇d{\bf H}^{d} to be dd-dimensional hyperbolic space, that is, the connected and simply connected Riemannian manifold of dimension dd having constant curvature 1-1. Let Γ\Gamma be a torsion-free discrete group of orientation preserving isometries of 𝐇d{\bf H}^{d} such that the quotient space 𝐇d/Γ{\bf H}^{d}/\Gamma has finite hyperbolic volume. Thus M=𝐇d/ΓM={\bf H}^{d}/\Gamma is a finite volume orientable hyperbolic dd-manifold.

There exists a compact hyperbolic dd-manifold MM^{\prime} with boundary (possibly empty) such that the complement MMM-M^{\prime} consists of at most finitely many disjoint unbounded ends of finite volume, the cusps of MM. Each cusp is homeomorphic to N×(0,)N\times(0,\infty) where NN is a compact Euclidean (d1)(d-1)-manifold.

Let Λ\Lambda denote the limit set of Γ\Gamma (i.e., the set of limit points of all the orbits of the action of Γ\Gamma on 𝐇d{\bf H}^{d}). A point cΛc\in\Lambda is called a parabolic limit point if it is the fixed point of some parabolic isometry γΓ\gamma\in\Gamma. The stabilizer Γc<Γ\Gamma_{c}<\Gamma of such a cc is called a maximal parabolic subgroup of Γ\Gamma. A cusp of Γ\Gamma is a Γ\Gamma-equivalence class of parabolic limit points and will be denoted by [c]Γ[c]_{\Gamma}. We will omit the subscript when the group is clear from context. The correspondence between cusps of MM and cusps of Γ\Gamma is given by the fact if CC is a cusp of MM then CC may be identified as C=Vc/ΓcC=V_{c}/\Gamma_{c} where Vc𝐇dV_{c}\subset{\bf H}^{d} is a precisely invariant horoball based at cc for some cusp [c][c] of Γ\Gamma.

3. Spectrum of the Laplacian

It is known that the space L2(M)L^{2}(M) has a decomposition

L2(M)=Ldisc2(M)Lcont2(M)L^{2}(M)=L^{2}_{disc}(M)\oplus L^{2}_{cont}(M)

where Ldisc2(M)L^{2}_{disc}(M) corresponds to the discrete spectrum of the Laplacian on MM and Lcont2(M)L^{2}_{cont}(M) corresponds to the continuous spectrum of MM. The discrete spectrum of MM is a collection of eigenvalues 0λ1λ20\leq\lambda_{1}\leq\lambda_{2}\leq\cdots where each λj\lambda_{j} occurs with a finite multiplicity. The continuous spectrum of MM is empty when MM is compact and otherwise is a union of finitely many intervals (one for each cusp of MM) of the form

[(d1)24,).\left[\frac{(d-1)^{2}}{4},\infty\right).

When MM is compact it is known that the discrete spectrum is infinite and obeys Weyl’s Asymptotic Law. The precise analogue of Weyl’s Asymptotic Law is in general not available when MM is not compact, though it is known in the case that Γ\Gamma is an arithmetic congruence group [17, 19, 20, 21].

The following elementary lemma will be useful in proving that certain manifolds have infinite discrete spectrum.

Lemma 3.1.

Let M=𝐇d/ΓM={\bf H}^{d}/\Gamma be a non-compact hyperbolic dd-manifold and M=𝐇d/ΓM^{\prime}={\bf H}^{d}/\Gamma^{\prime} be a finite cover of MM. If MM has an infinite discrete Laplace spectrum then so does MM^{\prime}.

Proof.

The eigenfunctions associated to the discrete Laplace spectrum of MM are the set of eigenfunctions of the Laplacian that are invariant under Γ\Gamma and which are L2L^{2}-integrable over some (and hence any) fundamental domain for Γ\Gamma. Any such function is also invariant under Γ\Gamma^{\prime}, and since the fundamental domain of Γ\Gamma^{\prime} is a finite union of fundamental domains of Γ\Gamma, the function will also be L2L^{2} integrable over a fundamental domain for Γ\Gamma^{\prime}. It follows that MM^{\prime} has an infinite discrete Laplace spectrum if MM does.∎

In order to discuss the spectrum of MM further we need to make clear the contribution of Eisenstein series. Let [c][c] be a cusp of Γ\Gamma with stabilizer Γc\Gamma_{c}. The Eisenstein series on MM associated to [c][c] is defined to be the convergent series

EM,c(w,s)=γΓc\Γy(σ1γw)s,w𝐇d,s𝐂,Re(s)>d1,E_{M,c}(w,s)=\sum_{\gamma\in\Gamma_{c}\backslash\Gamma}y(\sigma^{-1}\gamma w)^{s},\qquad w\in{\bf H}^{d},s\in\mathbf{C},\mathrm{Re}(s)>d-1,

where γΓ\gamma\in\Gamma represents a non-identity coset Γcγ\Gamma_{c}\gamma of Γc\Gamma_{c} in Γ\Gamma and σ\sigma is the orientation preserving isometry of hyperbolic space taking the point at infinity to the cusp point cc. Here we use the coordinates z=(x,y)𝐇d=𝐑d1×𝐑+z=(x,y)\in{\bf H}^{d}={\bf R}^{d-1}\times{\bf R}^{+} for the upper half-space.

Let c1,,cκc_{1},\dots,c_{\kappa} be representatives of a full set of inequivalent cusps of Γ\Gamma. To ease notation we will temporarily refer to the Eisenstein series associated to the ii-th cusp by Ei(w,s)E_{i}(w,s). The constant term of Ei(w,s)E_{i}(w,s) with respect to cjc_{j} is denoted Eij(w,s)E_{ij}(w,s) and satisfies

Eij(w,s)=δijy(σj1w)s+ϕij(s)y(σj1w)d1s,E_{ij}(w,s)=\delta_{ij}y(\sigma_{j}^{-1}w)^{s}+\phi_{ij}(s)y(\sigma_{j}^{-1}w)^{d-1-s},

where σj\sigma_{j} is the orientation preserving isometry of hyperbolic space taking the point at infinity to the cusp point cjc_{j} and where the coefficients ϕij(s)\phi_{ij}(s) define the scattering matrix Φ(s)=(ϕij)\Phi(s)=(\phi_{ij}). We define the scattering determinant to be the function φ(s)=detΦ(s)\varphi(s)=\det\Phi(s). The Eisenstein series Ej(w,s)E_{j}(w,s), the scattering matrix Φ(s)\Phi(s) and the scattering determinant ϕ(s)\phi(s) have meromorphic extensions to the complex plane. The poles of φ(s)\varphi(s) are poles of the Eisenstein series and all lie in the half-plane Re(s)<d12\mathrm{Re}(s)<\frac{d-1}{2}, except for at most finitely many poles in the interval (d12,d1](\frac{d-1}{2},d-1]. The latter poles are related to the discrete spectrum as follows. Taking the residue of Ej(w,s)E_{j}(w,s) at one of the latter poles yields an eigenfunction of the Laplacian with eigenvalue s(d1s)s(d-1-s). The subset of the discrete spectrum arising from residues of poles of Eisenstein series (equivalently, of φ(s)\varphi(s)) is called the residual spectrum. If tt is such a pole then we define the multiplicity at tt to be the order of the pole at tt, plus the dimension of the eigenspace in the case when tt contributes to the residual spectrum as described above. This discussion motivates the following definition.

Definition 3.2.

Let M1,M2M_{1},M_{2} be nn-cusped hyperbolic dd-manifolds (for some positive integer nn) of finite volume with scattering determinants φ1(s),φ2(s)\varphi_{1}(s),\varphi_{2}(s). We say that M1M_{1} and M2M_{2} are isospectral if

  • M1M_{1} and M2M_{2} have the same discrete spectrum, counting multiplicities;

  • φ1(s)\varphi_{1}(s) and φ2(s)\varphi_{2}(s) have the same set of poles and multiplicities.

The scattering determinant is in general very difficult to compute explicitly, although it has been worked out in several special case. For example, the scattering determinants associated to Hilbert modular groups over number fields have been computed in terms of Dedekind zeta functions by Efrat and Sarnak [3] and Masri [12].

4. Cusps of finite covers of hyperbolic manifolds

We begin with a group theoretic lemma. Let GG be a group, gg be an element of GG, and H,KH,K be subgroups of GG. We define the double coset HgKHgK by

HgK={hgk:hH,kK}.HgK=\{hgk:h\in H,k\in K\}.
Lemma 4.1.

There is a bijection between the cosets of HH in HgKHgK and the cosets of gKg1HgKg^{-1}\cap H in gKg1gKg^{-1}.

Proof.

Recall that HgKHgK is the union of the cosets HgkHgk as kk varies over the elements of KK. As right cosets of HH in GG, two cosets Hgk1Hgk_{1} and Hgk2Hgk_{2} intersect if and only if they are equal. Observe that Hgk1=Hgk2Hgk_{1}=Hgk_{2} if and only if there is an element hHh\in H such that gk1=hgk2gk_{1}=hgk_{2}, or equivalently, if and only if k1k21g1Hgk_{1}k_{2}^{-1}\in g^{-1}Hg (and thus is an element of Kg1HgK\cap g^{-1}Hg). This shows that Hgk1=Hgk2Hgk_{1}=Hgk_{2} if and only if (Kg1Hg)k1=(Kg1Hg)k2(K\cap g^{-1}Hg)k_{1}=(K\cap g^{-1}Hg)k_{2}. We have therefore shown that the map ff given by f(Hgk)=(Kg1Hg)kf(Hgk)=(K\cap g^{-1}Hg)k is a bijection between the cosets of HH in HgKHgK and of Kg1HgK\cap g^{-1}Hg in KK. We can now conjugate by gg to obtain a bijection between the cosets of HH in HgKHgK and the cosets of (gKg1H)(gKg^{-1}\cap H) in gKg1gKg^{-1}. ∎

Let Γ\Gamma be a discrete subgroup of Isom+(𝐇d)\mathrm{Isom}^{+}({\bf H}^{d}) and x,y𝐇dx,y\in\partial{\bf H}^{d} be Γ\Gamma-equivalent. Let GG be a subgroup of Γ\Gamma of finite index. We now define the set

Γx,y={γΓ:γxGy}.\Gamma_{x,y}=\{\gamma\in\Gamma:\gamma x\in G\cdot y\}.
Lemma 4.2.

There is an equality of sets Γx,y=GγPx\Gamma_{x,y}=G\gamma P_{x}, where Px=StabΓ(x)P_{x}=\mathrm{Stab}_{\Gamma}(x) and γ\gamma is any element of Γ\Gamma such that γx=y\gamma x=y.

Proof.

That any element of GγPxG\gamma P_{x} lies in Γx,y\Gamma_{x,y} is clear. Suppose therefore that δΓx,y\delta\in\Gamma_{x,y} and that δx=gy=g(γx)\delta x=gy=g(\gamma x). Then (gγ)1δx=x(g\gamma)^{-1}\delta x=x, hence γ1g1δPx\gamma^{-1}g^{-1}\delta\in P_{x} and there exists pPxp\in P_{x} such that γ1g1δ=p\gamma^{-1}g^{-1}\delta=p. This implies that δ=gγpGγPx\delta=g\gamma p\in G\gamma P_{x} and completes the proof of the lemma. ∎

Let M=𝐇d/ΓM={\bf H}^{d}/\Gamma and N=𝐇d/GN={\bf H}^{d}/G be non-compact hyperbolic dd-manifolds of finite volume and

π:NM\pi:N\longrightarrow M

be a covering. Let cc represent a cusp of Γ\Gamma and P=StabΓ(c)P=\mathrm{Stab}_{\Gamma}(c).

Definition.

The preimage of a cusp of MM is always a union of cusps of NN. We say a cusp of MM remains a cusp of NN relative to π\pi when the preimage of that cusp has precisely one cusp of NN. Algebraically, this is equivalent to [c]Γ=[c]G[c]_{\Gamma}=[c]_{G}.

Lemma 4.3.

Suppose cc is a cusp representative of both Γ\Gamma and GG and that [c]Γ=[c]G[c]_{\Gamma}=[c]_{G}. Then there is an equality of sets Γ=GP\Gamma=GP.

Proof.

That GPΓGP\subseteq\Gamma is clear as both GG and PP are subgroups of Γ\Gamma. Now let γΓ\gamma\in\Gamma. Since Γc=Gc\Gamma c=Gc there exists an element gGg\in G such that γc=gc\gamma c=gc. It follows that (g1γ)c=c(g^{-1}\gamma)c=c, hence g1γPg^{-1}\gamma\in P and there exists pPp\in P such that g1γ=pg^{-1}\gamma=p. This implies that γ=gp\gamma=gp, concluding the proof. ∎

Theorem 4.4.

Let {d1,,dm}\{d_{1},\dots,d_{m}\} represent the GG-orbits on the elements of 𝐇d\partial{\bf H}^{d} belonging to the cusp [c][c] of Γ\Gamma. Then

[Γ:G]=i=1m[StabΓ(di):StabΓ(di)G].[\Gamma:G]=\sum_{i=1}^{m}[\mathrm{Stab}_{\Gamma}(d_{i}):\mathrm{Stab}_{\Gamma}(d_{i})\cap G].
Proof.

Write Γ\Gamma as a disjoint union of cosets GγiG\gamma_{i}:

Γ=i=1rGγi.\Gamma=\bigcup_{i=1}^{r}G\gamma_{i}.

Since Γ\Gamma acts transitively on [c][c], every element of [c][c] is in the GG orbit of γid1\gamma_{i}d_{1} for some ii. For each j{1,,m}j\in\{1,\dots,m\}, fix δjΓ\delta_{j}\in\Gamma such that δjd1=dj\delta_{j}d_{1}=d_{j}. By Lemma 4.2, Γd1,dj=GδjStabΓ(d1)\Gamma_{d_{1},d_{j}}=G\delta_{j}\mathrm{Stab}_{\Gamma}(d_{1}). Lemma 4.1 shows that Γd1,dj\Gamma_{d_{1},d_{j}} is the union of nn cosets of GG, where nn is the index of δjStabΓ(d1)δj1G\delta_{j}\mathrm{Stab}_{\Gamma}(d_{1})\delta_{j}^{-1}\cap G in δjStabΓ(d1)δj1\delta_{j}\mathrm{Stab}_{\Gamma}(d_{1})\delta_{j}^{-1}. As δjStabΓ(d1)δj1=StabΓ(δjd1)=StabΓ(dj)\delta_{j}\mathrm{Stab}_{\Gamma}(d_{1})\delta_{j}^{-1}=\mathrm{Stab}_{\Gamma}(\delta_{j}d_{1})=\mathrm{Stab}_{\Gamma}(d_{j}), we see that n=[StabΓ(dj):StabΓ(dj)G]n=[\mathrm{Stab}_{\Gamma}(d_{j}):\mathrm{Stab}_{\Gamma}(d_{j})\cap G].

Putting all of this together, we see that Γ\Gamma is the disjoint union of Γd1,dj\Gamma_{d_{1},d_{j}} as jj varies over {1,,m}\{1,\dots,m\}. Since each of these is the disjoint union of [StabΓ(dj):StabΓ(dj)G][\mathrm{Stab}_{\Gamma}(d_{j}):\mathrm{Stab}_{\Gamma}(d_{j})\cap G] cosets of GG, we conclude that

[Γ:G]=i=1m[StabΓ(di):StabΓ(di)G],[\Gamma:G]=\sum_{i=1}^{m}[\mathrm{Stab}_{\Gamma}(d_{i}):\mathrm{Stab}_{\Gamma}(d_{i})\cap G],

which completes our proof.∎

Corollary 4.5.

We have an equality of indices [Γ:G]=[StabΓ(d):StabΓ(d)G][\Gamma:G]=[\mathrm{Stab}_{\Gamma}(d):\mathrm{Stab}_{\Gamma}(d)\cap G] for all cusps [d][d] of GG if and only if every cusp of MM remains a cusp of NN.

Proof.

We first prove that if every cusp of MM remains a cusp of NN then [Γ:G]=[StabΓ(d):StabG(d)][\Gamma:G]=[\mathrm{Stab}_{\Gamma}(d):\mathrm{Stab}_{G}(d)] for all cusps [d][d] of GG. Fix a cusp [d][d] of GG and define P=StabΓ(d)P=\mathrm{Stab}_{\Gamma}(d). We must show that [P:PG]=[Γ:G][P:P\cap G]=[\Gamma:G]. To that end, suppose that p1,p2Pp_{1},p_{2}\in P. Then

Gp1Gp2\displaystyle Gp_{1}\cap Gp_{2}\neq\emptyset Gp1=Gp2\displaystyle\iff Gp_{1}=Gp_{2}
p1=gp2 for some gG\displaystyle\iff p_{1}=gp_{2}\text{ for some }g\in G
p1p21=g\displaystyle\iff p_{1}p_{2}^{-1}=g
p1p21PG\displaystyle\iff p_{1}p_{2}^{-1}\in P\cap G
(PG)p1=(PG)p2.\displaystyle\iff(P\cap G)p_{1}=(P\cap G)p_{2}.

We have therefore exhibited a bijection between the cosets of GG in GP=ΓGP=\Gamma (the equality follows from Lemma 4.3) and the cosets of (PG)(P\cap G) in PP, hence [Γ:G]=[P:PG][\Gamma:G]=[P:P\cap G].

As the reverse direction is an immediate consequence of Theorem 4.4, our proof is complete. ∎

Corollary 4.6.

Suppose that NN is a normal cover of MM. Let [c][c] be a cusp of Γ\Gamma and [d][d] be a cusp of GG contained in [c][c]. The number of cusps of GG contained in [c][c] is

[Γ:G][StabΓ(d):StabΓ(d)G].\frac{[\Gamma:G]}{[\mathrm{Stab}_{\Gamma}(d):\mathrm{Stab}_{\Gamma}(d)\cap G]}.
Proof.

In light of Theorem 4.4 it suffices to prove that if [di],[dj][d_{i}],[d_{j}] are cusps of GG contained in the cusp [c][c] of Γ\Gamma then [StabΓ(di):StabΓ(di)G]=[StabΓ(dj):StabΓ(dj)G][\mathrm{Stab}_{\Gamma}(d_{i}):\mathrm{Stab}_{\Gamma}(d_{i})\cap G]=[\mathrm{Stab}_{\Gamma}(d_{j}):\mathrm{Stab}_{\Gamma}(d_{j})\cap G]. To that end, let γΓ\gamma\in\Gamma be such that γdi=dj\gamma d_{i}=d_{j}. Then

StabΓ(dj)=StabΓ(γdi)=γStabΓ(di)γ1,\mathrm{Stab}_{\Gamma}(d_{j})=\mathrm{Stab}_{\Gamma}(\gamma d_{i})=\gamma\mathrm{Stab}_{\Gamma}(d_{i})\gamma^{-1},

hence, as G=γGγ1G=\gamma G\gamma^{-1}, we have

[StabΓ(dj):StabΓ(dj)G]=[γStabΓ(di)γ1:γStabΓ(di)γ1γGγ1]=[StabΓ(di):StabΓ(di)G],[\mathrm{Stab}_{\Gamma}(d_{j}):\mathrm{Stab}_{\Gamma}(d_{j})\cap G]=[\gamma\mathrm{Stab}_{\Gamma}(d_{i})\gamma^{-1}:\gamma\mathrm{Stab}_{\Gamma}(d_{i})\gamma^{-1}\cap\gamma G\gamma^{-1}]=[\mathrm{Stab}_{\Gamma}(d_{i}):\mathrm{Stab}_{\Gamma}(d_{i})\cap G],

which completes the proof. ∎

5. Eisenstein series

Theorem 5.1.

Let M=𝐇d/ΓM={\bf H}^{d}/\Gamma be a non-compact hyperbolic dd-manifold and N=𝐇d/GN={\bf H}^{d}/G be a finite cover of MM with covering degree nn. If a cusp [c][c] of Γ\Gamma is also a cusp of GG (i.e., the preimage in NN of the corresponding cusp of MM is a single cusp) then EM,c(w,s)=EN,c(w,s)E_{M,c}(w,s)=E_{N,c}(w,s).

Proof.

Let cc represent a fixed cusp of Γ\Gamma and P=StabΓ(c)P=\mathrm{Stab}_{\Gamma}(c). We begin our proof by noting that Theorem 4.4 shows that [Γ:G]=[P:PG][\Gamma:G]=[P:P\cap G], hence we may select a collection of coset representatives for PGP\cap G in PP which is also a collection of coset representatives for GG in Γ\Gamma. Let {δ1,,δn}P\{\delta_{1},\dots,\delta_{n}\}\subset P be such a collection.

An arbitrary term of EM,c(w,s)E_{M,c}(w,s) is of the form y(σ1γw)sy(\sigma^{-1}\gamma w)^{s} where γΓ\gamma\in\Gamma represents a non-identity coset PγP\gamma of PP in Γ\Gamma and σ\sigma is the orientation preserving isometry of hyperbolic space taking the point at infinity to the cusp point cc. Here we use the coordinates z=(x,y)𝐇d=𝐑d1×𝐑+z=(x,y)\in{\bf H}^{d}={\bf R}^{d-1}\times{\bf R}^{+} for the upper half-space. Using our decomposition of Γ\Gamma into cosets of GG we see that there exists δj\delta_{j} and gGg\in G such that γ=δjg\gamma=\delta_{j}g. Because δjP\delta_{j}\in P, the coset Pγ=PδjgP\gamma=P\delta_{j}g is equal to the coset PgPg as cosets of P\ΓP\backslash\Gamma. In particular this implies that we may choose representatives for the cosets P\ΓP\backslash\Gamma to all lie in GG. Note that for all g1,g2Gg_{1},g_{2}\in G we have

Pg1=Pg2\displaystyle Pg_{1}=Pg_{2} g1g21P\displaystyle\iff g_{1}g_{2}^{-1}\in P
g1g21PG\displaystyle\iff g_{1}g_{2}^{-1}\in P\cap G
(PG)g1=(PG)g2.\displaystyle\iff(P\cap G)g_{1}=(P\cap G)g_{2}.

It follows that

EM,c(w,s)=γP\Γy(σ1γw)s=gPG\Gy(σ1gw)s=EN,c(w,s).E_{M,c}(w,s)=\sum_{\gamma\in P\backslash\Gamma}y(\sigma^{-1}\gamma w)^{s}=\sum_{g\in P\cap G\backslash G}y(\sigma^{-1}gw)^{s}=E_{N,c}(w,s).

The following is an immediate consequence of Theorem 5.1.

Corollary 5.2.

Suppose that MM is a cusped orientable finite volume hyperbolic dd-manifold and that M1,M2M_{1},M_{2} are finite covers of MM with the same covering degree and having the property that every cusp of MM remains a cusp of MiM_{i} (i=1,2i=1,2). Then all of the Eisenstein series of M1M_{1} and M2M_{2} are equal.

6. Congruence covers and pp-reps

Let MM be a non-compact finite volume orientable hyperbolic 33-manifold. Let c1,,cκc_{1},\dots,c_{\kappa} represent a complete set of inequivalent cusps of π1(M)\pi_{1}(M) and PiP_{i} be the subgroup of π1(M)\pi_{1}(M) that fixes cic_{i}.

Definition 6.1.

A surjective homomorphism ρ:π1(M)PSL(2,p)\rho:\pi_{1}(M)\rightarrow\operatorname{PSL}(2,p) is called a pp-rep if, for all ii, ρ(Pi)\rho(P_{i}) is non-trivial and all non-trivial elements of ρ(Pi)\rho(P_{i}) are parabolic elements of PSL(2,p)\operatorname{PSL}(2,p).

We remark that if ρ:π1(M)PSL(2,p)\rho:\pi_{1}(M)\rightarrow\operatorname{PSL}(2,p) is a pp-rep then ρ(Pi)\rho(P_{i}) must be a subgroup of PSL(2,p)\operatorname{PSL}(2,p) of order pp.

Theorem 6.2.

Let MM be a 11-cusped, non-arithmetic, finite volume orientable hyperbolic 33-manifold with pp-reps ρ:π1(M)PSL(2,7)\rho:\pi_{1}(M)\rightarrow\operatorname{PSL}(2,7) and ρ:π1(M)PSL(2,11)\rho^{\prime}:\pi_{1}(M)\rightarrow\operatorname{PSL}(2,11). Let kk be a number field with ring of integers 𝒪k\mathcal{O}_{k} and degree not divisible by 33. Assume that the faithful discrete representation of π1(M)\pi_{1}(M) can be conjugated to lie in PSL(2,𝒪k)\operatorname{PSL}(2,\mathcal{O}_{k}). There exist infinitely many prime powers qq and covers MqM_{q} of MM such that:

  1. (i)

    the composite homomorphism

    ρq:=ρι:π1(Mq)π1(M)PSL(2,7)\rho_{q}:=\rho\circ\iota:\pi_{1}(M_{q})\hookrightarrow\pi_{1}(M)\rightarrow\operatorname{PSL}(2,7)

    is a pp-rep,

  2. (ii)

    the degree over MM of the cover MqM_{q} is 112(q3q)\frac{11}{2}(q^{3}-q),

  3. (iii)

    the number of cusps of MqM_{q} is at least q+1q+1, and

  4. (iv)

    MqM_{q} has an infinite discrete spectrum.

Proof.

We begin by constructing a finite cover M~\widetilde{M} of MM which has an infinite discrete spectrum. The manifold MqM_{q} will arise as a finite cover of M~\widetilde{M} and will therefore have an infinite discrete spectrum by virtue of Lemma 3.1. To that end, let HH be an index 1111 subgroup of PSL(2,11)\operatorname{PSL}(2,11). Such a subgroup is well-known to exist, and the cover of MM associated to the pullback subgroup of HH by ρ\rho^{\prime} is a degree 1111 cover of MM. Denote this cover by M~\widetilde{M}. We claim that M~\widetilde{M} has one cusp. Let PP be the subgroup of π1(M)\pi_{1}(M) stabilizing the cusp of MM. As was commented above, ρ(P)\rho^{\prime}(P) must be a cyclic subgroup of PSL(2,11)\operatorname{PSL}(2,11) of order 1111. Since HH has index 1111 in PSL(2,11)\operatorname{PSL}(2,11) and |PSL(2,11)|=660=223511|\operatorname{PSL}(2,11)|=660=2^{2}\cdot 3\cdot 5\cdot 11 it must be the case that ρ(P)H\rho^{\prime}(P)\cap H is trivial. It follows that [P:Pπ1(M~)]=11=[π1(M):π1(M~)][P:P\cap\pi_{1}(\widetilde{M})]=11=[\pi_{1}(M):\pi_{1}(\widetilde{M})], hence M~\widetilde{M} has one cusp by Corollary 4.5. It now follows from [4, Theorem 2.4] that M~\widetilde{M} has an infinite discrete spectrum. We note that [4, Theorem 2.4] has two hypotheses: that M~\widetilde{M} be non-arithmetic and that M~\widetilde{M} not be the minimal element in its commensurability class. That M~\widetilde{M} is non-arithmetic is clear, since it is a finite cover of MM, which is non-arithmetic. It is equally clear that M~\widetilde{M} is not the minimal element of its commensurability class, since such an element cannot be a finite cover of another hyperbolic 33-manifold.

We claim that π1(M~)\pi_{1}(\widetilde{M}) also admits a pp-rep to PSL(2,7)\operatorname{PSL}(2,7). In particular, we will show the homomorphism to PSL(2,7)\operatorname{PSL}(2,7) obtained by composing the inclusion map π1(M~)π1(M)\pi_{1}(\widetilde{M})\hookrightarrow\pi_{1}(M) with ρ:π1(M)PSL(2,7)\rho:\pi_{1}(M)\rightarrow\operatorname{PSL}(2,7) is a pp-rep. To see this, note that because gcd(11,|PSL(2,7)|)=1\gcd(11,|\operatorname{PSL}(2,7)|)=1, the map gg11g\mapsto g^{11} is a bijection from PSL(2,7)\operatorname{PSL}(2,7) to itself, hence our claim follows from the fact that for every γπ1(M)\gamma\in\pi_{1}(M) the element γ11\gamma^{11} lies in π1(M~)\pi_{1}(\widetilde{M}).

Given a proper, non-zero ideal II of 𝒪k\mathcal{O}_{k} we have a composite homomorphism

ϕI:π1(M~)PSL(2,𝒪k)PSL(2,𝒪k/I)\phi_{I}:\pi_{1}(\widetilde{M})\longrightarrow\operatorname{PSL}(2,\mathcal{O}_{k})\longrightarrow\operatorname{PSL}(2,\mathcal{O}_{k}/I)

called the level II congruence homomorphism. It follows from the Strong Approximation Theorem of Nori [14] and Weisfeiler [23] that for all but finitely many prime ideals 𝔭\mathfrak{p} of 𝒪k\mathcal{O}_{k} the level 𝔭\mathfrak{p} congruence homomorphism ϕ𝔭\phi_{\mathfrak{p}} is surjective.

By Dirichlet’s Theorem on Primes in Arithmetic Progressions we may choose a prime pp satisfying p5(mod168)p\equiv 5\pmod{168} which does not divide the discriminant of kk. Let 𝔭\mathfrak{p} be a prime ideal of 𝒪k\mathcal{O}_{k} lying above pp which has inertia degree ff satisfying gcd(f,3)=1\gcd(f,3)=1. Note that the existence of such a prime ideal 𝔭\mathfrak{p} follows from the well-known equality in algebraic number theory

[k:𝐐]=i=1ge(𝔭i/p)f(𝔭i/p),[k:\mathbf{Q}]=\sum_{i=1}^{g}e(\mathfrak{p}_{i}/p)f(\mathfrak{p}_{i}/p),

where p𝒪k=𝔭1𝔭gp\mathcal{O}_{k}=\mathfrak{p}_{1}\cdots\mathfrak{p}_{g}, e(𝔭i/p)e(\mathfrak{p}_{i}/p) denotes the ramification degree of 𝔭i\mathfrak{p}_{i} over pp and f(𝔭i/p)f(\mathfrak{p}_{i}/p) denotes the inertia degree of 𝔭i\mathfrak{p}_{i} over pp. In particular our assertion follows from the hypothesis that [k:𝐐][k:\mathbf{Q}] not be divisible by 33 and the fact that all of the ramification degrees e(𝔭i/p)e(\mathfrak{p}_{i}/p) are equal to one (since pp doesn’t divide the discriminant of kk and thus does not ramify in kk).

We observed above that it follows from the Strong Approximation Theorem that for all but finitely many primes the associated congruence homomorphism is surjective. In light of our use of Dirichlet’s Theorem on Primes in Arithmetic Progressions in the previous paragraph we may assume that 𝔭\mathfrak{p} was selected so that ϕ𝔭\phi_{\mathfrak{p}} is surjective. Let MqM_{q} be the cover of M~\widetilde{M} associated to the kernel of ϕ𝔭\phi_{\mathfrak{p}}. The cover MqM_{q} of M~\widetilde{M} is normal of degree

|PSL(2,𝒪k/𝔭)|=|PSL(2,pf)|=p3fpf2,|\operatorname{PSL}(2,\mathcal{O}_{k}/\mathfrak{p})|=|\operatorname{PSL}(2,p^{f})|=\frac{p^{3f}-p^{f}}{2},

which proves (ii) upon setting q=pfq=p^{f}.

Assertion (iii) follows from assertion (ii) and Corollary 4.6 since the image under ϕ𝔭\phi_{\mathfrak{p}} of a cusp stabilizer PiP_{i} will be an abelian subgroup of PSL(2,pf)\operatorname{PSL}(2,p^{f}) and thus will have order at most pf(pf1)2\frac{p^{f}(p^{f}-1)}{2} by the classification of subgroups of PSL(2,q)\operatorname{PSL}(2,q) (see [2]).

We now prove assertion (i). We will abuse notation and denote by ρ\rho the pp-rep from π1(M~)\pi_{1}(\widetilde{M}) onto PSL(2,7)\operatorname{PSL}(2,7). Because this pp-rep was obtained by composing the inclusion of π1(M~)\pi_{1}(\widetilde{M}) into π1(M)\pi_{1}(M) with the pp-rep from π1(M)\pi_{1}(M) onto PSL(2,7)\operatorname{PSL}(2,7) (which was also denoted ρ\rho), it suffices to prove assertion (i) with M~\widetilde{M} in place of MM. Let N=p3fpf2=[π1(M~):π1(Mq)]N=\frac{p^{3f}-p^{f}}{2}=[\pi_{1}(\widetilde{M}):\pi_{1}(M_{q})]. As ρq(π1(Mq))\rho_{q}(\pi_{1}(M_{q})) contains ρq(γN)=ρ(γN)=ρ(γ)N\rho_{q}(\gamma^{N})=\rho(\gamma^{N})=\rho(\gamma)^{N} for all γπ1(M~)\gamma\in\pi_{1}(\widetilde{M}) and ρ:π1(M)PSL(2,7)\rho:\pi_{1}(M)\rightarrow\operatorname{PSL}(2,7) is surjective, the surjectivity of ρq\rho_{q} follows from the fact (easily verifiable in SAGE [16]) that PSL(2,7)\operatorname{PSL}(2,7) is generated by the NNth powers of its elements whenever p5(mod168)p\equiv 5\pmod{168} and gcd(f,3)=1\gcd(f,3)=1.

Let P0P_{0} be the subgroup of π1(Mq)\pi_{1}(M_{q}) which fixes some cusp of MqM_{q} and PP be the subgroup of π1(M~)\pi_{1}(\widetilde{M}) fixing the corresponding cusp of M~\widetilde{M}. Because ρ:π1(M~)PSL(2,7)\rho:\pi_{1}(\widetilde{M})\rightarrow\operatorname{PSL}(2,7) is a pp-rep, ρ(P)\rho(P) consists entirely of parabolic elements and therefore is a subgroup of PSL(2,7)\operatorname{PSL}(2,7) of order 77. Note that [P:P0]=d[P:P_{0}]=d for some divisor dd of NN. We will show that NN, and thus dd, is not divisible by 77. Because pp was chosen so that p5(mod168)p\equiv 5\pmod{168}, we also have p5(mod7)p\equiv 5\pmod{7} (since 168=2337168=2^{3}\cdot 3\cdot 7). It is now an easy exercise in elementary number theory to show that N=p3fpf2N=\frac{p^{3f}-p^{f}}{2} is not divisible by 77 whenever gcd(f,3)=1\gcd(f,3)=1. Having shown that gcd(d,7)=1\gcd(d,7)=1, we observe that if γP\gamma\in P has non-trivial image in PSL(2,7)\operatorname{PSL}(2,7) then γdP0\gamma^{d}\in P_{0} and thus ρq(γd)=ρ(γ)d\rho_{q}(\gamma^{d})=\rho(\gamma)^{d} is non-trivial in PSL(2,7)\operatorname{PSL}(2,7). Since ρq(P0)\rho_{q}(P_{0}) is a subgroup of ρ(P)\rho(P) and thus also consists entirely of parabolic elements, this proves assertion (i). ∎

7. Sunada’s Method for constructing isospectral manifolds

We begin this section by recalling the statement of Sunada’s theorem [18].

Given a finite group GG with subgroups H1H_{1} and H2H_{2} we say that H1H_{1} and H2H_{2} are almost conjugate if, for all gGg\in G,

#(H1[g])=#(H2[g])\#(H_{1}\cap[g])=\#(H_{2}\cap[g])

where [g][g] denotes the conjugacy class of gg in GG.

Theorem 7.1 (Sunada).

Let MM be a Riemannian manifold and ρ:π1(M)G\rho:\pi_{1}(M)\rightarrow G be a surjective homomorphism. The coverings MH1M^{H_{1}} and MH2M^{H_{2}} of MM with fundamental groups ρ1(H1)\rho^{-1}(H_{1}) and ρ1(H2)\rho^{-1}(H_{2}) are isospectral.

The following is a group theoretic lemma of Prasad and Rajan [15, Lemma 1] which they used to reprove Sunada’s theorem. In what follows, if GG is a group and VV is a GG-module then VGV^{G} is the submodule of invariants of GG.

Lemma 7.2.

Suppose that GG is a finite group with almost conjugate subgroups H1H_{1} and H2H_{2}. Assume that VV is a representation space of GG over a field kk of characteristic zero. Then there exists an isomorphism i:VH1VH2i:V^{H_{1}}\rightarrow V^{H_{2}}, commuting with the action of any endomorphism Δ\Delta of VV which commutes with the action of GG on VV; i.e. the following diagram commutes:

VH1V^{H_{1}}VH2V^{H_{2}}VH1V^{H_{1}}VH2V^{H_{2}}Δ\DeltaiiΔ\Deltaii
Theorem 7.3.

Let M=𝐇3/ΓM={\bf H}^{3}/\Gamma be a cusped finite volume orientable hyperbolic 33-manifold that is non-arithmetic and that is the minimal element in its commensurability class (i.e., Γ=Comm(Γ)\Gamma=\mathrm{Comm}(\Gamma) where Comm()\mathrm{Comm}(\cdot) denotes the commensurator). Let M0=𝐇3/Γ0M_{0}={\bf H}^{3}/\Gamma_{0} be a finite cover of MM, GG be a finite group and H1,H2H_{1},H_{2} be non-conjugate almost conjugate subgroups of GG. Suppose that Γ\Gamma admits a homomorphism onto GG such that the induced composite homomorphism Γ0ΓG\Gamma_{0}\hookrightarrow\Gamma\rightarrow G is also onto. Let M1,M2M_{1},M_{2} be the finite covers of M0M_{0} associated to the pullback subgroups of H1H_{1} and H2H_{2} and assume that M1M_{1} and M2M_{2} both have the same number of cusps as M0M_{0}. Then M1M_{1} and M2M_{2} are are isospectral, have the same complex length spectra, are non-isometric and have infinite discrete spectra.

Proof.

Our proof will largely follow the proof of the analogous result of Garoufalidis and Reid [4, Theorem 3.1].

We begin by proving that the manifolds M1M_{1} and M2M_{2} are non-isometric. Let Γ1,Γ2\Gamma_{1},\Gamma_{2} be such that M1=𝐇3/Γ1M_{1}={\bf H}^{3}/\Gamma_{1} and M2=𝐇3/Γ2M_{2}={\bf H}^{3}/\Gamma_{2}. If M1M_{1} and M2M_{2} are isometric then there exists gIsom(𝐇3)g\in\mathrm{Isom}({\bf H}^{3}) such that gΓ1g1=Γ2g\Gamma_{1}g^{-1}=\Gamma_{2}. Such an element gg necessarily lies in the commensurator Comm(Γ)\mathrm{Comm}(\Gamma) of Γ\Gamma, and since Γ=Comm(Γ)\Gamma=\mathrm{Comm}(\Gamma) we see that gΓg\in\Gamma. By hypothesis there exists a surjective homomorphism ρ:ΓG\rho:\Gamma\rightarrow G. Projecting onto GG we see that ρ(g)H1ρ(g)1=H2\rho(g)H_{1}\rho(g)^{-1}=H_{2}, which contradicts our hypothesis that H1H_{1} and H2H_{2} be non-conjugate.

To prove that M1M_{1} and M2M_{2} are isospectral we must show that their scattering determinants have the same poles with multiplicities and that they have the same discrete spectrum. Since M1M_{1} and M2M_{2} have the same covering degree over M0M_{0}, that their scattering determinants have the same poles with multiplicities follows immediately from Theorem 5.1, which in fact shows that all of their Eisenstein series coincide. That M1M_{1} and M2M_{2} have the same discrete spectrum follows from Lemma 7.2 with k=𝐂k=\mathbf{C}, V=Ldisc2(M0)V=L^{2}_{disc}(M_{0}) and Δ\Delta the Laplacian.

That M1M_{1} and M2M_{2} have the same complex length spectra follows from the proof given by Sunada [18, Section 4].

That M1M_{1} and M2M_{2} have infinite discrete spectra follows from [4, Theorem 2.4].

8. Proof of Theorem 1.1

In light of Theorems 6.2 and 7.3 it suffices to exhibit a non-arithmetic, 11-cusped finite volume hyperbolic 33-manifold MM which is the minimal element in its commensurability class and which admits pp-reps onto PSL(2,7)\operatorname{PSL}(2,7) and PSL(2,11)\operatorname{PSL}(2,11).

To prove this assertion, let MM be a hyperbolic 33-manifold as in the previous paragraph and assume that π1(M)\pi_{1}(M) can be conjugated to lie in PSL(2,𝒪k)\operatorname{PSL}(2,\mathcal{O}_{k}) for some number field kk whose degree is not divisible by 33. (We will construct such a manifold below.) It follows from Theorem 6.2 that there exist infinitely many prime powers qq and covers MqM_{q} of MM such that composing the inclusion π1(Mq)π1(M)\pi_{1}(M_{q})\hookrightarrow\pi_{1}(M) with the pp-rep π1(M)PSL(2,7)\pi_{1}(M)\rightarrow\operatorname{PSL}(2,7) yields a pp-rep and such that MqM_{q} has at least q+1q+1 cusps.

We have seen that there is a surjective homomorphism ρ:π1(Mq)PSL(2,7)\rho:\pi_{1}(M_{q})\rightarrow\operatorname{PSL}(2,7). It is well known that PSL(2,7)\operatorname{PSL}(2,7) contains a pair of non-conjugate, almost conjugate subgroups of index 77. Call these subgroups H1H_{1} and H2H_{2} and observe that since |PSL(2,7)|=168|PSL(2,7)|=168, it must be that H1H_{1} and H2H_{2} have order 2424. Let Mi=𝐇3/ΓiM_{i}={\bf H}^{3}/\Gamma_{i} (i=1,2i=1,2) be the manifold covers of MqM_{q} associated to H1H_{1} and H2H_{2}.

Fix i{1,2}i\in\{1,2\} and let [d][d] be a cusp of Γi\Gamma_{i}. Let Pi=StabΓi(d)P_{i}=\mathrm{Stab}_{\Gamma_{i}}(d) and P=Stabπ1(Mq)(d)P=\mathrm{Stab}_{\pi_{1}(M_{q})}(d). Because the homomorphism ρ:π1(Mq)PSL(2,7)\rho:\pi_{1}(M_{q})\rightarrow\operatorname{PSL}(2,7) is a pp-rep, ρ(P)\rho(P) is a cyclic subgroup of PSL(2,7)\operatorname{PSL}(2,7) of order 77. Since HiH_{i} has order 2424 it must be that ρ(P)Hi\rho(P)\cap H_{i} is trivial. In particular it follows that ρ(Pi)=1\rho(P_{i})=1 and consequently that [π1(Mq):Γi]=7=[P:Pi][\pi_{1}(M_{q}):\Gamma_{i}]=7=[P:P_{i}]. Corollary 4.5 now implies that every cusp of MqM_{q} remains a cusp of MiM_{i}. In particular this shows that M1M_{1} and M2M_{2} both have the same number of cusps as MqM_{q}, and this number can be made arbitrarily large by taking the prime power qq (from Theorem 6.2) to be arbitrarily large. Theorem 1.1 now follows from Theorem 7.3.

We now construct a non-arithmetic, 11-cusped finite volume hyperbolic 33-manifold MM which is the minimal element in its commensurability class and which admits pp-reps onto PSL(2,7)\operatorname{PSL}(2,7) and PSL(2,11)\operatorname{PSL}(2,11). We will additionally show that π1(M)\pi_{1}(M) can be conjugated to lie in PSL(2,𝒪k)\operatorname{PSL}(2,\mathcal{O}_{k}) where kk is a number field of degree 88.

Refer to caption
Figure 1. The knot K11n116.

To that end, let KK be the knot K11n116 of the Hoste-Thistlethwaite table shown in Figure 1. The manifold M=S3K=𝐇3/ΓM=S^{3}\setminus K={\bf H}^{3}/\Gamma has 11 cusp, volume 7.75445376027.7544537602\cdots and invariant trace field k=𝐐(t)k=\mathbf{Q}(t) where t=0.00106+0.9101192it=0.00106+0.9101192i is a root of the polynomial x82x7x6+4x53x3+x+1x^{8}-2x^{7}-x^{6}+4x^{5}-3x^{3}+x+1. It was proven in [6] that MM is the minimal element in its commensurability class (i.e., that Γ=Comm(Γ)\Gamma=\mathrm{Comm}(\Gamma) where Comm(Γ)\mathrm{Comm}(\Gamma) denotes the commensurator of Γ\Gamma). The work of Margulis [11] shows that this implies MM must be non-arithmetic. Moreover, a computation in Snap [5] shows that Γ\Gamma has presentation

Γ=a,b,c | aaCbAccBB, aacbCbAAB,\Gamma=\langle a,b,c\text{ }|\text{ }aaCbAccBB,\text{ }aacbCbAAB\rangle,

and peripheral structure

μ=CbAcb,λ=AAbCCbacb.\mu=CbAcb,\qquad\lambda=AAbCCbacb.

Here A=a1,B=b1,C=c1A=a^{-1},B=b^{-1},C=c^{-1}. In terms of matrices, we may represent Γ\Gamma as a subgroup of PSL(2,𝒪k)\operatorname{PSL}(2,\mathcal{O}_{k}) via

a=(t2+t1t73t6+4t5t4+t2tt2+t10),a=\begin{pmatrix}-t^{2}+t-1&t^{7}-3t^{6}+4t^{5}-t^{4}+t^{2}-t\\ -t^{2}+t-1&0\end{pmatrix},
b=(t7+2t62t53t3+2t23t1t62t5+t4+3t32t2+3t+2t7+3t65t5+4t44t3+2t22t1t73t6+5t54t4+4t3t2+t+2),b=\begin{pmatrix}-t^{7}+2t^{6}-2t^{5}-3t^{3}+2t^{2}-3t-1&t^{6}-2t^{5}+t^{4}+3t^{3}-2t^{2}+3t+2\\ -t^{7}+3t^{6}-5t^{5}+4t^{4}-4t^{3}+2t^{2}-2t-1&t^{7}-3t^{6}+5t^{5}-4t^{4}+4t^{3}-t^{2}+t+2\end{pmatrix},

and

c=(t6+4t58t4+7t35t2t2t7+7t614t5+15t412t3+t2+3t1t53t4+4t33t2+tt7+4t69t5+11t49t3+3t2+t2).c=\begin{pmatrix}-t^{6}+4t^{5}-8t^{4}+7t^{3}-5t^{2}-t&-2t^{7}+7t^{6}-14t^{5}+15t^{4}-12t^{3}+t^{2}+3t-1\\ t^{5}-3t^{4}+4t^{3}-3t^{2}+t&-t^{7}+4t^{6}-9t^{5}+11t^{4}-9t^{3}+3t^{2}+t-2\end{pmatrix}.

We now show that Γ\Gamma admits pp-reps onto PSL(2,7)\operatorname{PSL}(2,7) and PSL(2,11)\operatorname{PSL}(2,11). We begin by exhibiting the pp-rep onto PSL(2,7)\operatorname{PSL}(2,7). As the discriminant of kk is 156166337156166337, which is not divisible by 77, we see that 77 is unramified in k/𝐐k/\mathbf{Q}. Using SAGE [16] we find that 7𝒪k=𝔭1𝔭2𝔭3,7\mathcal{O}_{k}=\mathfrak{p}_{1}\mathfrak{p}_{2}\mathfrak{p}_{3}, where the inertia degrees of the 𝔭i\mathfrak{p}_{i} are 1,2,51,2,5. We note that the prime 𝔭1\mathfrak{p}_{1} of norm 77 is equal to the principal ideal (t1)(t-1). Upon identifying 𝒪k/𝔭1\mathcal{O}_{k}/\mathfrak{p}_{1} with 𝐅7{\bf F}_{7} we obtain a homomorphism from Γ\Gamma to PSL(2,7)\operatorname{PSL}(2,7) by reducing the matrix entries of a,b,ca,b,c modulo 𝔭1\mathfrak{p}_{1}. The images of a,b,ca,b,c in PSL(2,7)\operatorname{PSL}(2,7) are represented by

a=(6160),b=(1635),c=(3405),a=\begin{pmatrix}6&1\\ 6&0\end{pmatrix},\qquad b=\begin{pmatrix}1&6\\ 3&5\end{pmatrix},\qquad c=\begin{pmatrix}3&4\\ 0&5\end{pmatrix},

while the images of μ,λ\mu,\lambda in PSL(2,7)\operatorname{PSL}(2,7) are represented by the parabolic matrices

μ=(0455),λ=(2513).\mu=\begin{pmatrix}0&4\\ 5&5\end{pmatrix},\qquad\lambda=\begin{pmatrix}2&5\\ 1&3\end{pmatrix}.

It remains only to show that the homomorphism we have defined, call it ρ7\rho_{7}, is surjective. Our proof of this will make use of the following easy lemma.

Lemma 8.1.

Let pp be a prime. The group SL(2,p)\operatorname{SL}(2,p) is generated by the matrices

T=(1101),U=(1011).T=\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\qquad U=\begin{pmatrix}1&0\\ 1&1\end{pmatrix}.
Proof.

The lemma follows from the fact that SL(2,𝐙)\operatorname{SL}(2,\mathbf{Z}) is generated by the matrices in the lemma’s statement. To see this, note that the usual generators of SL(2,𝐙)\operatorname{SL}(2,\mathbf{Z}) are

S=(0110),T=(1101),S=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix},\qquad T=\begin{pmatrix}1&1\\ 0&1\end{pmatrix},

and S=T1UT1S=T^{-1}UT^{-1}.∎

Surjectivity of our homomorphism ρ7:ΓPSL(2,7)\rho_{7}:\Gamma\rightarrow\operatorname{PSL}(2,7) now follows from the fact that

(1101)=ρ7(b)1ρ7(a)2ρ7(b)1ρ7(a)ρ7(b)1\begin{pmatrix}1&1\\ 0&1\end{pmatrix}=\rho_{7}(b)^{-1}\rho_{7}(a)^{-2}\rho_{7}(b)^{-1}\rho_{7}(a)\rho_{7}(b)^{-1}

and

(1011)=ρ7(c)ρ7(a)1ρ7(b)ρ7(c)2.\begin{pmatrix}1&0\\ 1&1\end{pmatrix}=\rho_{7}(c)\rho_{7}(a)^{-1}\rho_{7}(b)\rho_{7}(c)^{2}.

We have just shown that Γ\Gamma admits a pp-rep onto PSL(2,7)\operatorname{PSL}(2,7). We now show that Γ\Gamma admits a pp-rep onto PSL(2,11)\operatorname{PSL}(2,11) as well. In kk we have the factorization 11𝒪k=𝔭1𝔭2𝔭311\mathcal{O}_{k}=\mathfrak{p}_{1}\mathfrak{p}_{2}\mathfrak{p}_{3} where the inertia degrees of the 𝔭i\mathfrak{p}_{i} are 1,1,61,1,6. We may assume without loss of generality that 𝔭1=(t4)\mathfrak{p}_{1}=(t-4). Identifying 𝒪k/𝔭1\mathcal{O}_{k}/\mathfrak{p}_{1} with 𝐅11{\bf F}_{11} we see that the images in PSL(2,11)\operatorname{PSL}(2,11) of a,b,ca,b,c are represented by the matrices

a=(9690),b=(4311),c=(10164),a=\begin{pmatrix}9&6\\ 9&0\end{pmatrix},\qquad b=\begin{pmatrix}4&3\\ 1&1\end{pmatrix},\qquad c=\begin{pmatrix}10&1\\ 6&4\end{pmatrix},

while the images of μ,λ\mu,\lambda in PSL(2,11)\operatorname{PSL}(2,11) are represented by the parabolic matrices

μ=(10101010),λ=(100610).\mu=\begin{pmatrix}10&10\\ 10&10\end{pmatrix},\qquad\lambda=\begin{pmatrix}10&0\\ 6&10\end{pmatrix}.

Finally, we show that our homomorphism ρ11:ΓPSL(2,11)\rho_{11}:\Gamma\rightarrow\operatorname{PSL}(2,11) is surjective by applying Lemma 8.1. To that end we simply note that

(1101)=ρ11(a)1ρ11(b)ρ11(c)1\begin{pmatrix}1&1\\ 0&1\end{pmatrix}=\rho_{11}(a)^{-1}\rho_{11}(b)\rho_{11}(c)^{-1}

and

(1011)=ρ11(c)ρ11(a)2.\begin{pmatrix}1&0\\ 1&1\end{pmatrix}=\rho_{11}(c)\rho_{11}(a)^{2}.

This completes the proof of Theorem 1.1.

References

  • [1] R. Brooks and O. Davidovich. Isoscattering on surfaces. J. Geom. Anal., 13(1):39–53, 2003.
  • [2] L. E. Dickson. Linear groups: With an exposition of the Galois field theory. with an introduction by W. Magnus. Dover Publications, Inc., New York, 1958.
  • [3] I. Efrat and P. Sarnak. The determinant of the Eisenstein matrix and Hilbert class fields. Trans. Amer. Math. Soc., 290(2):815–824, 1985.
  • [4] S. Garoufalidis and A. W. Reid. Constructing 1-cusped isospectral non-isometric hyperbolic 3-manifolds. J. Topol. Anal., 10(1):1–25, 2018.
  • [5] O. Goodman. Snap, the computer program. http://www.ms.unimelb.edu.au/\simsnap.
  • [6] O. Goodman, D. Heard, and C. Hodgson. Commensurators of cusped hyperbolic manifolds. Experiment. Math., 17(3):283–306, 2008.
  • [7] C. S. Gordon. Survey of isospectral manifolds. In Handbook of differential geometry, Vol. I, pages 747–778. North-Holland, Amsterdam, 2000.
  • [8] C. Gordon, P. Perry, and D. Schueth. Isospectral and isoscattering manifolds: a survey of techniques and examples. In Geometry, spectral theory, groups, and dynamics, volume 387 of Contemp. Math., pages 157–179. Amer. Math. Soc., Providence, RI, 2005.
  • [9] A. Ikeda. On lens spaces which are isospectral but not isometric. Ann. Sci. École Norm. Sup. (4), 13(3):303–315, 1980.
  • [10] M. Kac. Can one hear the shape of a drum? Amer. Math. Monthly, 73(4, part II):1–23, 1966.
  • [11] G. A. Margulis. Discrete subgroups of semisimple Lie groups, volume 17 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1991.
  • [12] R. Masri. The scattering matrix for the Hilbert modular group. Proc. Amer. Math. Soc., 137(8):2541–2555, 2009.
  • [13] J. Milnor, Eigenvalues of the Laplace operator on certain manifolds, Proc. Nat. Acad. Sci. U.S.A. 51 (1964), 542.
  • [14] M. V. Nori. On subgroups of GLn(𝐅p){\rm GL}_{n}({\bf F}_{p}). Invent. Math., 88(2):257–275, 1987.
  • [15] D. Prasad and C. S. Rajan. On an Archimedean analogue of Tate’s conjecture. J. Number Theory, 99(1):180–184, 2003.
  • [16] The Sage Developers. SageMath, the Sage Mathematics Software System (Version 8.8), 2019. https://www.sagemath.org.
  • [17] A. Selberg. Harmonic analysis and discontinuous groups in weakly symmetric Riemannian spaces with applications to Dirichlet series. J. Indian Math. Soc. (N.S.), 20:47–87, 1956.
  • [18] Toshikazu Sunada, Riemannian coverings and isospectral manifolds, Ann. of Math. (2) 121 (1985), no. 1, 169–186.
  • [19] A. B. Venkov. The asymptotic formula connected with the number of eigenvalues of the Laplace-Beltrami operator on the fundamental domain of the modular group PSL(2,Z)PSL(2,Z) that correspond to odd eigenfunctions. Dokl. Akad. Nauk SSSR, 233(6):1021–1023, 1977.
  • [20] A. B. Venkov. Artin-Takagi formula for the Selberg zeta function and the Roelcke hypothesis. Dokl. Akad. Nauk SSSR, 247(3):540–543, 1979.
  • [21] A. B. Venkov. Spectral theory of automorphic functions. Proc. Steklov Inst. Math., (4(153)):ix+163 pp. (1983), 1982. A translation of Trudy Mat. Inst. Steklov. 153 (1981).
  • [22] Marie-France Vignéras, Variétés riemanniennes isospectrales et non isométriques, Ann. of Math. (2) 112 (1980), no. 1, 21–32.
  • [23] B. Weisfeiler. Strong approximation for Zariski-dense subgroups of semisimple algebraic groups. Ann. of Math. (2), 120(2):271–315, 1984.