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

Hyperbolic manifolds with a large number of systoles

Cayo Dória Emanoel M. S. Freire  and  Plinio G. P. Murillo Cayo Dória
Universidade Federal de Sergipe, Departamento de Matemática.
Av. Marcelo Déda Chagas s/n, 49100-000. São Cristóvão - SE, Brazil
cayo@mat.ufs.br Emanoel M. S. Freire
IMPA
Estrada Dona Castorina, 110
22460-320 Rio de Janeiro, Brazil
emanoel.m.s.freire@gmail.com Plinio G. P. Murillo
Universidade Federal Fluminense, Instituto de Matemática e Estatística.
Rua Prof. Marcos Waldemar de Freitas Reis, s/n, Bloco H, Campus do Gragoatã, 24210-201. Niterói-RJ, Brazil.
pliniom@id.uff.br
Abstract.

In this article, for any n4n\geq 4 we construct a sequence of compact hyperbolic nn-manifolds {Mi}\{M_{i}\} with number of systoles at least as vol(Mi)1+13n(n+1)ϵ\mathrm{vol}(M_{i})^{1+\frac{1}{3n(n+1)}-\epsilon} for any ϵ>0\epsilon>0. In dimension 3, the bound is improved to vol(Mi)43ϵ\mathrm{vol}(M_{i})^{\frac{4}{3}-\epsilon}. These results generalize previous work of Schmutz for n=2n=2, and Dória-Murillo for n=3n=3 to higher dimensions.

1991 Mathematics Subject Classification:
53C22, 11F06
The second author was supported by a CAPES research grant.

1. Introduction

A hyperbolic nn-manifold is a nn-dimensional manifold without boundary, equipped with a complete Riemannian metric of constant curvature 1-1. Any hyperbolic manifold is isometric to a quotient space M=Γ\nM=\Gamma\backslash\mathbb{H}^{n}, where n\mathbb{H}^{n} is the hyperbolic nn-space and Γ\Gamma is a torsion-free discrete group of isometries of n\mathbb{H}^{n}. In recent years a lot of progress has been made in the study of hyperbolic manifolds with extremal properties. For example, such spaces with minimal volume [Bel14], minimal diameter [BCP21], large systolic length [Mur19] and large kissing number (see below the definitions of systolic length and kissing number). In this article we are interested in hyperbolic manifolds with large kissing number, and their relation with its systolic length and volume.

It is well known that any finite volume hyperbolic manifold MM contains closed geodesics, and there exists at least one of minimal length, which is called a systole of MM. The length of any systole of MM will be called the systolic length of MM and denoted by sys(M)\mathrm{sys}(M) .

We define the kissing number Kiss(M)\mathrm{Kiss}(M) of MM as the number of systoles of MM. This is a well-defined invariant since, in negative curvature there are finitely many closed geodesics with the same length. This terminology was introduced by Schmutz Schaller, and it was inspired by the classical kissing number of lattices arising in sphere packings (see [Sch96a],[Sch96b],[Sch95]).

In general, it follows from a classical result of Anosov [Ano83] that a generic Riemannian manifold has at most one systole. For a closed hyperbolic nn-manifold MM, it is possible to bound Kiss(M)\mathrm{Kiss}(M) by above in terms of sys(M)\mathrm{sys}(M) or vol(M)\mathrm{vol}(M). It follows from works by Keen [Kee74] and Buser [Bus80] that there exist constants Cn,Dn>0C_{n},D_{n}>0 depending only on nn such that if sys(M)Cn\mathrm{sys}(M)\leq C_{n}, then

(1.1) Kiss(M)Dnvol(M).\mathrm{Kiss}(M)\leq D_{n}\mathrm{vol}(M).

Recently, Bourque and Petri proved an upper bound of Kiss(M)\mathrm{Kiss}(M) which holds for any systolic length (see [BP22, Theorem 1]). More precisely,

(1.2) Kiss(M)Anvol(M)exp(n12sys(M))sys(M),\mathrm{Kiss}(M)\leq A_{n}\mathrm{vol}(M)\frac{\exp\left(\frac{n-1}{2}\mathrm{sys}(M)\right)}{\mathrm{sys}(M)},

for some constant An>0A_{n}>0 which depends only on n.n. Since sinh(x)x1\frac{\sinh(x)}{x}\rightarrow 1 whenever x0x\rightarrow 0, we can unify the inequalities (1.1) and (1.2) in

(1.3) Kiss(M)Anvol(M)sinh(n12sys(M))sys(M)\mathrm{Kiss}(M)\leq A^{\prime}_{n}\mathrm{vol}(M)\frac{\sinh\left(\frac{n-1}{2}\mathrm{sys}(M)\right)}{\mathrm{sys}(M)}

for some constant An>0A^{\prime}_{n}>0 which depends only on nn 111We thank the referee for providing us this remark.. However, if sys(M)\mathrm{sys}(M) is large, Inequality (1.2) implies that

(1.4) Kiss(M)Bnvol(M)2log(1+vol(M)),\mathrm{Kiss}(M)\leq B_{n}\frac{\mathrm{vol}(M)^{2}}{\log(1+\mathrm{vol}(M))},

(see [BP22, Corollary 1.2]). In dimension 2, this result was previously established in 20132013 by Parlier (see [Par13]). In [FP16], similar upper bounds were established for non-compact hyperbolic surfaces of finite area. Whether a version of (1.2) and (1.4) holds for non-compact finite volume hyperbolic manifolds in n3n\geq 3 remains open.

These restrictions for Kiss(M)\mathrm{Kiss}(M), and the aforementioned result by Anosov motivated us to study the following question formulated in [Pet20]: Let n2n\geq 2 and

Kn(v)=max{Kiss(M)M is a hyperbolic n-manifold of vol(M)v}K_{n}(v)=\max\{\mathrm{Kiss}(M)\mid M\mbox{ is a hyperbolic $n$-manifold of }\mathrm{vol}(M)\leq v\}
Question 1.

How does Kn(v)K_{n}(v) grows as a function of vv?

Although this question is independent of the size of sys(M)\mathrm{sys}(M), it is interesting to understand Kiss(M)\mathrm{Kiss}(M) according to whether sys(M)\mathrm{sys}(M) is small or large. The first result in this article shows that there exist a sequence of closed hyperbolic manifolds NiN_{i} with bounded systolic length, and Kiss(Ni)\mathrm{Kiss}(N_{i}) growing at least as linearly with vol(Ni)\mathrm{vol}(N_{i}).

Theorem A.

Let n2n\geq 2 and A>0A>0. Then, there exist positive real numbers B,CB,C with ABA\leq B, and a sequence {Ni}\{N_{i}\} of closed hyperbolic nn-manifolds such that

  • Asys(Ni)BA\leq\mathrm{sys}(N_{i})\leq B for all i>0i>0,

  • vol(Ni)\mathrm{vol}(N_{i})\to\infty

  • Kiss(Ni)Cvol(Ni)\mathrm{Kiss}(N_{i})\geq C\cdot\mathrm{vol}(N_{i}).

In particular,

lim supvlogKn(v)logv1\limsup\limits_{v\to\infty}\frac{\log K_{n}(v)}{\log v}\geq 1

for any n2n\geq 2.

This result shows that the exponent one of vol(M)\mathrm{vol}(M) in (1.2) is the best possible that we can obtain. The manifolds in Theorem A are obtained by taking cyclic covers of a fix hyperbolic manifold with positive first Betti number. The use of this technique is well known in spectral geometry (see e.g [Ran74]). The proof of Theorem (A) is given in Section 6.

The main part of the article is devoted to give an answer to Question 1 independently on the size of the systoles. For n=2n=2, it follows from results by Schmutz in [Sch95] that

(1.5) lim supvlogK2(v)logv1+13.\limsup\limits_{v\to\infty}\dfrac{\log K_{2}(v)}{\log v}\geq 1+\frac{1}{3}.

To prove this result, in [op. cit.] the author constructed a sequence SiS_{i} of closed (also non-compact of finite area) hyperbolic surfaces with large kissing number as congruence coverings of a fixed arithmetic hyperbolic surface. It is worth to note that the surfaces Si{S_{i}} also satisfy

sys(Si)43log(area(Si))i.\mathrm{sys}(S_{i})\sim\frac{4}{3}\log(\mathrm{area}(S_{i}))\xrightarrow{i\rightarrow\infty}\infty.

More generally, if a sequence MiM_{i} of non-diffeomorphic closed hyperbolic nn-manifolds has Kiss(Mi)\mathrm{Kiss}(M_{i}) growing super linearly in vol(Mi)\mathrm{vol}(M_{i}) (i.e. Kiss(Mi)Cvol(Mi)1+ε\mathrm{Kiss}(M_{i})\geq C\mathrm{vol}(M_{i})^{1+\varepsilon} for some constants C,ε>0C,\varepsilon>0), then sys(Mi)\mathrm{sys}(M_{i}) grows logarithmic in vol(Mi)\mathrm{vol}(M_{i}). Indeed, it follows from (1.3) since vol(Mi)\mathrm{vol}(M_{i})\to\infty, and sinhxx>M\dfrac{\sinh x}{x}>M implies x>log(M)x>\log(M) for any M>0M>0. Hence

(1.6) sys(Mi)2εn1log(vol(Mi))+2n1log(2CAn(n1)).\mathrm{sys}(M_{i})\geq\frac{2\varepsilon}{n-1}\log(\mathrm{vol}(M_{i}))+\frac{2}{n-1}\log\left(\frac{2C}{A_{n}(n-1)}\right).

In [Mur19], the third author showed that congruence coverings of closed arithmetic hyperbolic nn-manifold of the first type have systole with length growing logarithmically in the volume, and determined the precise growth ratio. It is then natural to investigate the kissing number of such manifolds, and to ask whether they can provide a version of (1.5) in higher dimension. In this direction we obtain the following

Theorem B.

For any n2n\geq 2, there exists a compact arithmetic hyperbolic n-manifold of the first type MM, and a sequence of congruence coverings MjMM_{j}\to M of arbitrarily large degree such that

(1.7) Kiss(Mj)Cvol(Mj)1+13n(n+1)log(vol(Mj))\mathrm{Kiss}(M_{j})\geq C\ \frac{\mathrm{vol}(M_{j})^{1+\frac{1}{3n(n+1)}}}{\log(\mathrm{vol}(M_{j}))}

for some constant C>0C>0 independent of MjM_{j}. In particular,

lim supvlogKn(v)logv1+13n(n+1)\limsup\limits_{v\to\infty}\frac{\log K_{n}(v)}{\log v}\geq 1+\frac{1}{3n(n+1)}

for any n2.n\geq 2.

Although the systolic length of congruence coverings of arithmetic manifolds is large, exhibiting the systoles in these spaces is a much more delicate problem. We overcome this by constructing MjM_{j} containing a totally geodesic surface SjS_{j} whose systoles are also systoles of MjM_{j}.

Theorem C.

For any n3n\geq 3, the manifold MM obtained in Theorem B contains a closed totally geodesic surface SS such that for any jj, the congruence coverings MjMM_{j}\to M contains a congruence covering SjSS_{j}\to S satisfying sys(Sj)=sys(Mj).\mathrm{sys}(S_{j})=\mathrm{sys}(M_{j}).

This is the key part of the proof of Theorem B (see Corollary 5.7). The result then follows from an argument of high multiplicity inspired by [Sch95]. The proof is presented in Section 6.

The last part of the article reserves a special attention to dimension 33. In [DM21], the first and last authors constructed non-compact hyperbolic manifolds NiN_{i} satisfying

logKiss(Ni)logvol(Ni)1+427.\frac{\log\mathrm{Kiss}(N_{i})}{\log\mathrm{vol}(N_{i})}\gtrsim 1+\frac{4}{27}.

In this case, the manifolds NiN_{i} are congruence coverings of Bianchi orbifolds. Once more, the technique includes ideas by Schmutz, but a new length-trace relation was needed, and also a result on averages of class numbers of imaginary binary forms proved by Sarnak. These results are no longer available in the compact setting. Instead, we have now been able to use the holonomy of closed geodesics. We are able to construct closed arithmetic hyperbolic 33-manifolds with a large number of systoles using the relation between length and trace of 2×22\times 2 matrices, and a result on the equidistribution of closed geodesics with holonomy in prescribed intervals proved by Sarnak and Wakayama in [SW99] (see also [MMO14]). We finish the article with the proof of the following theorem in Section 7.

Theorem D.

There exists a sequence {Mj}\{M_{j}\} of compact arithmetic hyperbolic 33-manifold with vol(Mj)\mathrm{vol}(M_{j}) going to infinity such that

Kiss(Mj)Cvol(Mj)4/3log(vol(Mj))\mathrm{Kiss}(M_{j})\geq C\frac{\mathrm{vol}(M_{j})^{4/3}}{\log(\mathrm{vol}(M_{j}))}

where C>0C>0 is a universal constant.

Comments and an open question. We recall that two positive sequences (aj)(a_{j}) and (bj)(b_{j}) satisfy the relation ajbja_{j}\gtrsim b_{j} (resp. ajbja_{j}\lesssim b_{j}) when for any δ>0\delta>0 there exists j0j_{0} such that ajbj(1δ)\frac{a_{j}}{b_{j}}\geq(1-\delta) (resp. ajbj1δ\frac{a_{j}}{b_{j}}\leq 1-\delta) for j>j0j>j_{0}. Hence, the sequences satisfy ajbja_{j}\sim b_{j} if and only if ajbja_{j}\lesssim b_{j} and ajbj.a_{j}\gtrsim b_{j}. A natural question arising from Theorem B is the following:

Question 2.

Is there a universal ε>0\varepsilon>0 such that for any n2n\geq 2, there is a sequence of closed hyperbolic nn-manifolds MjM_{j} with vol(Mj)\mathrm{vol}(M_{j})\to\infty and

Kiss(Mj)vol(Mj)1+εlog(vol(Mj))?\mathrm{Kiss}(M_{j})\gtrsim\frac{\mathrm{vol}(M_{j})^{1+\varepsilon}}{\log(\mathrm{vol}(M_{j}))}~{}?

We have already noticed that this would imply

(1.8) sys(Mj)2εn1log(vol(Mj))\mathrm{sys}(M_{j})\gtrsim\frac{2\varepsilon}{n-1}\log(\mathrm{vol}(M_{j}))

(see Inequality (1.6)). From the Appendix of [Mur19], for any n2n\geq 2 there is a sequence of compact arithmetic hyperbolic nn-manifolds MiM_{i} with vol(Mi)\mathrm{vol}(M_{i})\rightarrow\infty such that

(1.9) sys(Mi)8n(n+1)log(vol(Mi)),\mathrm{sys}(M_{i})\sim\frac{8}{n(n+1)}\log(\mathrm{vol}(M_{i})),

and the hypothetical bound (1.8) would be considerably larger than the growth in (1.9). While writing this article the authors do not know any improvement for (1.9).

Acknowledgements. We would like to thank the referees for their comments and suggestions which substantially improved the previous version of this article.

2. Preliminaries

2.1. Hyperbolic Manifolds

The hyperbolic nn-space is the complete simply connected nn-dimensional Riemannian manifold with constant sectional curvature equal to 1-1. The hyperboloid model of the hyperbolic nn-space is given by

n={xn+1;x02+x12++xn2=1,x0>0}\mathbb{H}^{n}=\{x\in\mathbb{R}^{n+1};\ -x_{0}^{2}+x_{1}^{2}+\cdots+x_{n}^{2}=-1,\ x_{0}>0\}

with the metric ds2=dx02+dx12++dxn2ds^{2}=-dx_{0}^{2}+dx_{1}^{2}+\cdots+dx_{n}^{2}.

The identity component SO(n,1)\mathrm{SO}(n,1)^{\circ} of the Lie group SO(n,1)\mathrm{SO}(n,1) is isomorphic to the orientation preserving isometries of n\mathbb{H}^{n}. Given a lattice ΓIsom+(n)\Gamma\subset\mathrm{Isom}^{+}(\mathbb{H}^{n}), i.e, a discrete subgroup having finite covolume with respect to the Haar measure, the associated quotient space M=Γ\nM=\Gamma\backslash\mathbb{H}^{n} is a finite volume hyperbolic orbifold. It is a manifold whenever Γ\Gamma is torsion-free.

2.2. Systole and Kissing number

We recall that an element γ\gamma in Isom+(n)\mathrm{Isom}^{+}(\mathbb{H}^{n}) is called

  • elliptic, if it has a fixed point on n\mathbb{H}^{n}.

  • parabolic, if it has exactly one fixed point in n\partial\mathbb{H}^{n}.

  • loxodromic, if it has two fixed points in n\partial\mathbb{H}^{n}.

The displacement at xnx\in\mathbb{H}^{n} of a loxodromic element γ\gamma is defined by l(γ,x):=d(x,γx)l(\gamma,x):=d(x,\gamma x). The displacement of γ\gamma (also called translation length) is defined by

l(γ):=infxnl(γ,x).l(\gamma):=\mathrm{inf}_{x\in\mathbb{H}^{n}}l(\gamma,x).

A systole of a hyperbolic orbifold M=Γ\nM=\Gamma\backslash\mathbb{H}^{n} is any closed geodesic of shortest length in MM, and its length is denoted by sys(M)\mathrm{sys}(M). The kissing number Kiss(M)\mathrm{Kiss}(M) is defined as the number of free homotopy classes of oriented closed geodesics in MM that realize sys(M)\mathrm{sys}(M). It is well known that when MM is a manifold, there exists a one-to-one correspondence between parametrized closed geodesic in MM up to unit speed reparametrization and conjugacy classes of loxodromic elements in Γ\Gamma. Moreover, the length of the closed geodesic corresponds to the translation length of the loxodromic element. This relation allows to study sys(M)\mathrm{sys}(M) and Kiss(M)\mathrm{Kiss}(M) through the number of conjugacy classes of loxodromic elements in Γ\Gamma that realize sys(M)\mathrm{sys}(M) as their translation length.

2.3. Clifford algebras and the Spin group

In this section we will recall the construction of spinor groups. These are the algebraic groups associated to fundamental groups of arithmetic hyperbolic orbifolds. For further details, we refer the reader to [Mur19, Section 2], and the references therein.

Let kk be a field with char k2k\neq 2, EE a nn-dimensional vector space over kk, and ff a non-degenerate quadratic form on EE with associated bilinear form Φ\Phi. The Clifford algebra of ff, denoted by 𝒞(f,k)\mathscr{C}(f,k), is a unitary associative algebra over kk given by the quotient 𝒞(f,k)=T(E)/𝔞f\mathscr{C}(f,k)=T(E)/\mathfrak{a}_{f}, where T(E)T(E) denotes the tensor algebra of EE, and 𝔞f\mathfrak{a}_{f} is the two-sided ideal of T(E)T(E) generated by the elements xy+yx2Φ(x,y)x\otimes y+y\otimes x-2\Phi(x,y). If we choose an orthogonal basis e1,,ene_{1},\dots,e_{n} of EE with respect to f,f, we have in 𝒞(f,k)\mathscr{C}(f,k) the relations ev2=f(ev)e_{v}^{2}=f(e_{v}) and eveμ=eμeve_{v}e_{\mu}=-e_{\mu}e_{v} for μ,v{1,,n},μv.\mu,v\in\{1,\ldots,n\},\ \mu\neq v. Let 𝒫n\mathscr{P}_{n} be the power set of {1,,n}\{1,...,n\}. For M={μ1,,μν}𝒫nM=\{\mu_{1},\dots,\mu_{\nu}\}\in\mathscr{P}_{n} with μ1<<μν,\mu_{1}<\dots<\mu_{\nu}, we define eM=eμ1eμνe_{M}=e_{\mu_{1}}\cdot...\cdot e_{\mu_{\nu}}, where we adopt the convention e=1e_{\emptyset}=1. The 2n2^{n} elements eMe_{M}, M𝒫nM\in\mathscr{P}_{n} determine a basis for 𝒞(f,k)\mathscr{C}(f,k). Hence, any s𝒞(f,k)s\in\mathscr{C}(f,k) is written uniquely as

s=s1+M𝒫n,MsMeM,s=s_{\mathbb{R}}\cdot 1+\sum_{\mathclap{M\in\mathscr{P}_{n},~{}M\neq\emptyset}}s_{M}e_{M},

with s,sMks_{\mathbb{R}},s_{M}\in k. We call the coefficient ss_{\mathbb{R}} as the real part of ss. We identify kk with k1k\cdot 1, and EE with the kk-linear subspace in 𝒞(f,k)\mathscr{C}(f,k) generated by e1,e2,,ene_{1},e_{2},\ldots,e_{n}. The algebra 𝒞(f,k)\mathscr{C}(f,k) has an anti-involution *. On the elements eMe_{M} this map acts by eM=(1)ν(ν1)/2eMe_{M}^{*}=(-1)^{\nu(\nu-1)/2}\ e_{M}, where ν=|M|\nu=|M|. The span of the elements eMe_{M} with |M||M| even is a subalgebra 𝒞+(f)\mathscr{C}^{+}(f) of 𝒞(f),\mathscr{C}(f), called the even Clifford subalgebra of ff. The spin group of ff is defined as

Spinf(k):={s𝒞+(f,k)|sEsE,ss=1}.\mathrm{Spin}_{f}(k):=\left\{s\in\mathscr{C}^{+}(f,k)\ \Big{|}\ sEs^{*}\subseteq E,\ ss^{*}=1\right\}.

In the case k=k=\mathbb{R}, E=n+1E=\mathbb{R}^{n+1} and f=x02+x12++xn2f=-x_{0}^{2}+x_{1}^{2}+\cdots+x_{n}^{2}, the corresponding spin group is denoted by Spin(n,1)\mathrm{Spin}(n,1). For an element sSpin(n,1)s\in\mathrm{Spin}(n,1) the linear map φs:EE\varphi_{s}:E\to E given by φs(x)=sxs\varphi_{s}(x)=sxs^{*} lies in SO(n,1)\mathrm{SO}(n,1)^{\circ}, and the map sφss\to\varphi_{s} is a two-sheeted covering of SO(n,1)\mathrm{SO}(n,1)^{\circ}, with kernel {±1}\{\pm 1\}. Since the image of a lattice under a finite covering map is also a lattice, in order to produce hyperbolic orbifolds we will implicitly contruct lattices in Spin(n,1)\mathrm{Spin}(n,1) and project them to SO(n,1)\mathrm{SO}(n,1)^{\circ}. Furthermore, we will abuse terminology saying that an element sSpin(n,1)s\in\mathrm{Spin}(n,1) is elliptic (resp. parabolic or loxodromic) if φs\varphi_{s} is elliptic (resp. parabolic or loxodromic).

3. Length inequality in Spin(1,n)\mathrm{Spin}(1,n)

In order to obtain explicitly the systole of a closed manifold M=Γ\nM=\Gamma\backslash\mathbb{H}^{n}, we need to find a hyperbolic element γ0Γ\gamma_{0}\in\Gamma such that (γ0)(γ)\ell(\gamma_{0})\leq\ell(\gamma) for any nontrivial element γΓ\gamma\in\Gamma. As it has been observed in [Mur19], it is useful to estimate the displacement of hyperbolic elements using information about the real part of elements in the spin group. Now we will give a more precise version of this relation. The main tool will be the connection between spin and Vahlen groups established by Elstrodt, Grunewald and Mennicke [EGM87], and a characterization of the translation length by Waterman [Wat93].

Proposition 3.1.

For any loxodromic element rSpin(1,n)r\in\mathrm{Spin}(1,n) we have that

(r)2cosh1(|r|).\ell(r)\geq 2\cosh^{-1}(|r_{\mathbb{R}}|).
Proof.

Let us consider k=k=\mathbb{R} and the quadratic forms

Q0(y0,y1,y2)\displaystyle Q_{0}(y_{0},y_{1},y_{2}) =y02y12y22\displaystyle=y_{0}^{2}-y_{1}^{2}-y_{2}^{2}
Q(x1,x2,,xn2)\displaystyle Q(x_{1},x_{2},\ldots,x_{n-2}) =x12xn22\displaystyle=-x_{1}^{2}-\ldots-x_{n-2}^{2}
Q~(y0,y1,y2,x1,,xn2)\displaystyle\tilde{Q}(y_{0},y_{1},y_{2},x_{1},\ldots,x_{n-2}) =Q0(y0,y1,y2)+Q(x1,,xn2).\displaystyle=Q_{0}(y_{0},y_{1},y_{2})+Q(x_{1},\ldots,x_{n-2}).

It is clear that SpinQ~()=Spin(1,n)\mathrm{Spin}_{\tilde{Q}}(\mathbb{R})=\mathrm{Spin}(1,n). By [EGM87, Theorem 4.1], any element rSpin(1,n)r\in\mathrm{Spin}(1,n) is the image under a group isomorphism ψ\psi of a 2×22\times 2 matrix with coefficients in the Clifford algebra 𝒞(Q)\mathscr{C}(Q). More precisely, and following the notation in [loc. cit] a direct computation shows that

(3.1) r\displaystyle r =ψ((abcd))\displaystyle=\psi\left(\begin{pmatrix}a&b\\ c&d\\ \end{pmatrix}\right)
=a˙12(1+f0f1)+b˙12(f0f2f1f2)+c˙12(f0f2+f1f2)+d˙12(1f0f1)\displaystyle=\dot{a}\frac{1}{2}(1+f_{0}f_{1})+\dot{b}\frac{1}{2}(f_{0}f_{2}-f_{1}f_{2})+\dot{c}\frac{1}{2}(f_{0}f_{2}+f_{1}f_{2})+\dot{d}\frac{1}{2}(1-f_{0}f_{1})

where a,b,c,d𝒞(Q)a,b,c,d\in\mathscr{C}(Q). Here f0,f_{0}, f1f_{1} and f2f_{2} denote a basis of orthogonal elements for Q0Q_{0}, and \cdot denotes the KK-algebra homomorphism from 𝒞(Q)\mathscr{C}(Q) to 𝒞(Q~)\mathscr{C}(\tilde{Q}) given by

x˙=(M𝒫nxMeM)=M𝒫nxM(f0f1f2)ξMeM,\dot{x}=\left(\sum_{M\in\mathscr{P}_{n}}x_{M}e_{M}\right)^{\cdot}=\sum_{M\in\mathscr{P}_{n}}x_{M}(f_{0}f_{1}f_{2})^{\xi_{M}}e_{M},

where

ξM={0for|M|0mod 21for|M|1mod 2,\xi_{M}=\left\{\begin{array}[]{rll}0\ \text{for}\ |M|\equiv 0\ \text{mod}\ 2\\ \\ 1\ \text{for}\ |M|\equiv 1\ \text{mod}\ 2,\\ \end{array}\right.

(see [EGM87, Section 2]). The important point for us is that f0,f1,f2f_{0},f_{1},f_{2} are not elements in the center of 𝒞(Q~)\mathscr{C}(\tilde{Q}), and the real part of x˙\dot{x} is equal to that of xx, for any x𝒞(Q)x\in\mathscr{C}(Q). Looking at (3.1), this implies that the real part of rr is determined by the real part of a+da+d. More precisely

(3.2) r=(a+d)/2.r_{\mathbb{R}}=(a+d)_{\mathbb{R}}/2.

On the other hand, by [Wat93, Lemma 14]

(3.3) (a+d)=(λ+λ1)i=1n32cos(θi),(a+d)_{\mathbb{R}}=\left(\lambda+\lambda^{-1}\right)\prod_{i=1}^{\lfloor\frac{n-3}{2}\rfloor}\cos(\theta_{i}),

where λ\lambda is the multiplier of rr, and 2θ12\theta_{1}, 2θ2,,2θn322\theta_{2},\ldots,2\theta_{\lfloor\frac{n-3}{2}\rfloor} denote its rotational angles. In this case, λ=e(r)2\lambda=e^{\frac{\ell(r)}{2}} and we get by (3.2) and (3.3) that

r=cosh(l(r)2)i=1n32cos(θi).r_{\mathbb{R}}=\cosh\left(\frac{l(r)}{2}\right)\prod_{i=1}^{\lfloor\frac{n-3}{2}\rfloor}\cos(\theta_{i}).

Hence, |r|cosh(l(r)2)|r_{\mathbb{R}}|\leq\cosh\left(\frac{l(r)}{2}\right), and the proposition follows. ∎

4. Arithmetic subgroups of Spin(1,n)\mathrm{Spin}(1,n)

4.1. Arithmetic Hyperbolic Manifolds

We recall that a discrete subgroup ΓSpin(1,n)\Gamma\subset\mathrm{Spin}(1,n) is arithmetic if there exist a number field kk, a kk-algebraic group H, and an epimorphism φ:H(k)Spin(1,n)\varphi:\mathrm{\textbf{H}}(k\otimes_{\mathbb{Q}}\mathbb{R})\rightarrow\mathrm{Spin}(1,n) with compact kernel such that φ(H(𝒪k))\varphi(\mathrm{\textbf{H}}(\mathcal{O}_{k})) is commensurable to Γ\Gamma. Here, 𝒪k\mathcal{O}_{k} denotes the ring of integers of kk and H(𝒪k)=HGLn(𝒪k)\mathrm{\textbf{H}}(\mathcal{O}_{k})=\mathrm{\textbf{H}}\cap\mathrm{GL}_{n}(\mathcal{O}_{k}) with respect to some fixed embedding of H into GLn\mathrm{GL}_{n}. The field kk is the field of definition of Γ\Gamma. Any algebraic kk-group H satisfying these properties is called admissible. A hyperbolic orbifold M=Γ\nM=\Gamma\backslash\mathbb{H}^{n} such that Γ\Gamma is an arithmetic subgroup of Isom+(n)\mathrm{Isom}^{+}(\mathbb{H}^{n}) is called an arithmetic hyperbolic orbifold. It follows from Borel and Harish–Chandra’s Theorem that any arithmetic hyperbolic orbifold has finite volume [BHC62].

4.2. Arithmetic groups of the first type

The admissibility condition implies that kk is a totally real number field, H is a simply-connected algebraic kk-group, and by fixing an embedding kk\subset\mathbb{R} we can assume that H()=Spin(n,1)\mathrm{\textbf{H}}(\mathbb{R})=\mathrm{Spin}(n,1) (see [BE12, Sec. 2.2] and [Eme09, Sec. 13.1]). Suppose kk is a totally real number field, and ff is a quadratic form over kk. The admissibility of the algebraic kk-group H=Spinf\mathrm{\textbf{H}}=\mathrm{Spin}_{f} is equivalent to the fact that ff has signature (n,1)(n,1) over \mathbb{R}, and fσf^{\sigma} is definite for any non-trivial embedding σ:k\sigma:k\rightarrow\mathbb{R}, where fσf^{\sigma} is the quadratic form defined on the same kk-space of ff, given by fσ(x)=σ(f(x))f^{\sigma}(x)=\sigma(f(x)). A quadratic form satisfying these conditions will be called an admissible quadratic form. The arithmetic groups commensurable to Spinf(𝒪k)\mathrm{Spin}_{f}(\mathcal{O}_{k}) with ff admissible, are called arithmetic groups of the first type. For nn even, any arithmetic subgroup of Spin(n,1)\mathrm{Spin}(n,1) is of the first type. There is a second class in odd dimensions, arising from skew-hermitian forms over division quaternion algebras. For n=7n=7 there is a third class, arising from certain Cayley algebras. In this work we will deal with arithmetic hyperbolic orbifolds of the first type in dimensions n2n\geq 2.

4.3. Congruence subgroups

Let Γ\Gamma be an arithmetic subgroup of Spin(n,1)\mathrm{Spin}(n,1) commensurable with φ(H(𝒪k))\varphi(\mathrm{\textbf{H}}(\mathcal{O}_{k})), and M=Γ\nM=\Gamma\backslash\mathbb{H}^{n} the corresponding hyperbolic arithmetic orbifold. If I𝒪kI\subset\mathcal{O}_{k} is a non-zero ideal of 𝒪k\mathcal{O}_{k}, the principal congruence subgroup of Γ\Gamma associated to II is the subgroup Γ(I):=Γφ(H(I))\Gamma(I):=\Gamma\cap\varphi(\mathrm{\textbf{H}}(I)), where

H(I):=ker(H(𝒪k)πIH(𝒪k/I)),\mathrm{\textbf{H}}(I):=\mathrm{ker}\left(\mathrm{\textbf{H}}(\mathcal{O}_{k})\xrightarrow{\pi_{I}}\mathrm{\textbf{H}}(\mathcal{O}_{k}/I)\right),

and πI\pi_{I} denotes the reduction modulo II map. The associated principal congruence covering is MI=Γ(I)\nMM_{I}=\Gamma(I)\backslash\mathbb{H}^{n}\rightarrow M. Since Γ(I)\Gamma(I) is a normal finite-index subgroup of Γ\Gamma, the covering MIMM_{I}\rightarrow M is a regular finite-sheeted covering map. More generally, a discrete subgroup Λ<Γ\Lambda<\Gamma is called a congruence subgroup if Γ(I)Λ\Gamma(I)\subset\Lambda for some ideal I𝒪kI\subset\mathcal{O}_{k}.

Let ff be an admissible quadratic form over a totally real number field kk of degree dd. We can describe the group Γ=Spinf(𝒪k)\Gamma=\mathrm{Spin}_{f}(\mathcal{O}_{k}) and its principal congruence subgroups Γ(I)\Gamma(I) in the following way. Denote by e1,e2,,en+1e_{1},e_{2},\ldots,e_{n+1} an orthogonal basis with respect to ff. Then under the linear representation given by left multiplication in 𝒞+(f,)\mathscr{C}^{+}(f,\mathbb{R}) we get

Γ={s=|M| even sMeM|sM𝒪k and ss=1}\Gamma=\left\{s=\sum_{\mathclap{|M|\mbox{ even }}}s_{M}e_{M}~{}|~{}s_{M}\in\mathcal{O}_{k}\mbox{ and }ss^{*}=1\right\}

and

Γ(I)={s=|M| evensMeMΓ|sMI for M and s1I}.\Gamma(I)=\left\{s=\sum_{\mathclap{|M|\mbox{ even}}}s_{M}e_{M}\in\Gamma~{}|~{}s_{M}\in I\mbox{ for }M\neq\emptyset\mbox{ and }s_{\mathbb{R}}-1\in I\right\}.

(see [Mur19, Sec. 2.4]). To simplify the discussion, we will denote by 𝒬\mathcal{Q} the 𝒪k\mathcal{O}_{k}-order in 𝒞+(f,)\mathscr{C}^{+}(f,\mathbb{R}) given by

𝒬={s=|M| evensMeM|sM𝒪k}.\mathcal{Q}=\left\{s=\sum_{\mathclap{|M|\mbox{ even}}}s_{M}e_{M}~{}|~{}s_{M}\in\mathcal{O}_{k}\right\}.

4.4. Length inequality for Γ(α)\Gamma(\alpha)

For a principal ideal I=(α)I=(\alpha), α𝒪k\alpha\in\mathcal{O}_{k}, we denote Γ(I)\Gamma(I) simply by Γ(α)\Gamma(\alpha). In the same way, we denote by N(α)\mathrm{N}(\alpha) the norm of the ideal (α)(\alpha). We will present a series of results relating the real part of elements in Γ(α)\Gamma(\alpha) with α\alpha that will be necessary in the course of the investigation.

Lemma 4.1.

Let α𝒪k\alpha\in\mathcal{O}_{k} be a nonzero element. For any sΓ(α)s\in\Gamma(\alpha) we have the equality

s=1+12α2ζs_{\mathbb{R}}=1+\frac{1}{2}\alpha^{2}\zeta

for some ζ𝒪k.\zeta\in\mathcal{O}_{k}.

Proof.

By definition, we can write s=1+αts=1+\alpha t for some t𝒬t\in\mathcal{Q}. Since s=1+αts^{*}=1+\alpha t^{*} we have

1=ss=1+α(t+t)+α2tt.1=ss^{*}=1+\alpha(t+t^{*})+\alpha^{2}tt^{*}.

Now, t𝒬t\in\mathcal{Q} implies that (tt)(tt^{*})_{\mathbb{R}} lies in 𝒪k\mathcal{O}_{k}. By taking the equality of real parts, and observing that 2t=(t+t)2t_{\mathbb{R}}=(t+t^{*})_{\mathbb{R}}, the lemma follows with ζ=(tt)\zeta=-(tt^{*})_{\mathbb{R}}. ∎

The following complement of Lemma 4.1 will also be useful.

Lemma 4.2.

Let α𝒪k\alpha\in\mathcal{O}_{k} be a nonzero element and sΓs\in\Gamma such that ssα𝒬s-s_{\mathbb{R}}\in\alpha\mathcal{Q}. For any rΓ(α)r\in\Gamma(\alpha) we have the equality

(sr)=s+12α2ζ(sr)_{\mathbb{R}}=s_{\mathbb{R}}+\frac{1}{2}\alpha^{2}\zeta

for some ζ𝒪k.\zeta\in\mathcal{O}_{k}.

Proof.

Again, we have r=1+αtr=1+\alpha t for some t𝒬t\in\mathcal{Q} with 2t=αξ2t_{\mathbb{R}}=\alpha\xi for some ξ𝒪k.\xi\in\mathcal{O}_{k}. Since s=s+αus=s_{\mathbb{R}}+\alpha u for some u𝒬u\in\mathcal{Q} we can write

sr=s(1+αt)=s+α(s+αu)t=s+αst+α2ut.sr=s(1+\alpha t)=s+\alpha(s_{\mathbb{R}}+\alpha u)t=s+\alpha s_{\mathbb{R}}t+\alpha^{2}ut.

Since ut𝒬ut\in\mathcal{Q}, then (ut)(ut)_{\mathbb{R}} lies in 𝒪k\mathcal{O}_{k}. When we take the real part in the last equality we finish the proof with ζ=sξ+2(ut).\zeta=s_{\mathbb{R}}\xi+2(ut)_{\mathbb{R}}.

Proposition 4.3.

Let α𝒪k\alpha\in\mathcal{O}_{k} be a nonzero element. For any loxodromic element rΓ(α)r\in\Gamma(\alpha) we have

|r|>122d1N(α)21,|r_{\mathbb{R}}|>\frac{1}{2^{2d-1}}\mathrm{N}(\alpha)^{2}-1,

where d=[k:]d=[k:\mathbb{Q}].

Proof.

See [Mur19, Lemma 4.1]. ∎

4.5. The congruence subgroup Γτ(α)\Gamma_{\tau}(\alpha)

An important example of congruence subgroup which will play a special role in this work is the following. Fix an element α𝒪k\alpha\in\mathcal{O}_{k}. Suppose that τ(𝒪k/α𝒪k)×\tau\in(\mathcal{O}_{k}/\alpha\mathcal{O}_{k})^{\times} is an element of order 2, and define

Γτ(α)={γΓγΓ(α)orγτ(modα𝒬)}.\Gamma_{\tau}(\alpha)=\{\gamma\in\Gamma\mid\gamma\in\Gamma(\alpha)\hskip 5.69054pt\mbox{or}\hskip 5.69054pt\gamma\equiv\tau(\mathrm{mod}~{}\alpha\mathcal{Q})\}.

The group Γτ(α)\Gamma_{\tau}(\alpha) is a normal subgroup of Γ\Gamma. Indeed, let t𝒪kt\in\mathcal{O}_{k} such that τ=t+α𝒪k\tau=t+\alpha\mathcal{O}_{k}. We note that {1+α𝒬,t+α𝒬}\{1+\alpha\mathcal{Q},t+\alpha\mathcal{Q}\} is a central subgroup of (𝒬/α𝒬)×\left(\mathcal{Q}/\alpha\mathcal{Q}\right)^{\times}, hence Γτ(α)\Gamma_{\tau}(\alpha) is normal since it is the preimage of this normal subgroup under the natural projection map Γ(𝒬/α𝒬)×\Gamma\to\left(\mathcal{Q}/\alpha\mathcal{Q}\right)^{\times} .

5. Hyperbolic manifolds with a systole lying in a surface

The goal of this section is to show that, under certain conditions, the manifold Γτ(α)\n\Gamma_{\tau}(\alpha)\backslash\mathbb{H}^{n} has a systole contained in a totally geodesic surface.

5.1. A totally geodesic surface embedded in Γτ(α)\n\Gamma_{\tau}(\alpha)\backslash\mathbb{H}^{n}

Let kk be a totally real number field, and (E,f)(E,f) be an admissible (n+1)(n+1)-dimensional quadratic space over kk (see subsection 4.2). Since ff has signature (n,1)(n,1) and kk\subset\mathbb{R}, by the Gram-Schmidt process and the Law of Inertia, there exists a basis {e0,e1,,en}\{e_{0},e_{1},\ldots,e_{n}\} of EE such that in this basis ff has the diagonal form f=a0x02+a1x12++anxn2f=-a_{0}x_{0}^{2}+a_{1}x_{1}^{2}+\cdots+a_{n}x_{n}^{2}, with aika_{i}\in k and positive for all ii. In fact, we can suppose that the coefficients are in 𝒪k\mathcal{O}_{k} if we replace eie_{i} by dieid_{i}e_{i} where di𝒪kd_{i}\in\mathcal{O}_{k} is a denominator of aia_{i}. By admissibility, fσf^{\sigma} is positive definite for any non-trivial Galois embedding σ:k\sigma:k\rightarrow\mathbb{R}, hence σ(a0)=σ(a0)=fσ(e0)<0\sigma(a_{0})=-\sigma(-a_{0})=-f^{\sigma}(e_{0})<0 and σ(ai)=fσ(ei)>0\sigma(a_{i})=f^{\sigma}(e_{i})>0 for all i=1,,ni=1,\ldots,n.

Let EE^{\prime} be the subspace generated by {e0,e1,e2}\{e_{0},e_{1},e_{2}\}, and f:Ekf^{\prime}:E^{\prime}\rightarrow k the restriction of ff to EE^{\prime}. The inclusion EEE^{\prime}\rightarrow E defines a natural inclusion Γ=Spinf(𝒪k)Spinf(𝒪k)=Γ\Gamma^{\prime}=\mathrm{Spin}_{f^{\prime}}(\mathcal{O}_{k})\hookrightarrow\mathrm{Spin}_{f}(\mathcal{O}_{k})=\Gamma. For any α𝒪k\alpha\in\mathcal{O}_{k} and τ(𝒪k\α𝒪k)×\tau\in(\mathcal{O}_{k}\backslash\alpha\mathcal{O}_{k})^{\times} of order two, by definition we get an inclusion

Γτ(α)Γτ(α).\Gamma^{\prime}_{\tau}(\alpha)\hookrightarrow\Gamma_{\tau}(\alpha).

Consider an isometric embedding of 2\mathbb{H}^{2} into n\mathbb{H}^{n} equivariant with respect to the actions of Γ\Gamma^{\prime} and Γ\Gamma and the inclusions above. For any α\alpha and τ\tau as before, we obtain a totally geodesic embedding

(5.1) Sα,τMα,τ,S_{\alpha,\tau}\hookrightarrow M_{\alpha,\tau},

where Sα,τ=Γτ(α)\2S_{\alpha,\tau}=\Gamma^{\prime}_{\tau}(\alpha)\backslash\mathbb{H}^{2} and Mα,τ=Γτ(α)\nM_{\alpha,\tau}=\Gamma_{\tau}(\alpha)\backslash\mathbb{H}^{n}. This implies in particular that sys(Mα,τ)sys(Sα,τ)\mathrm{sys}(M_{\alpha,\tau})\leq\mathrm{sys}(S_{\alpha,\tau}).

Proposition 5.1.

Let α𝒪k\alpha\in\mathcal{O}_{k} be a nonzero element and sΓs\in\Gamma^{\prime} such that ssα𝒬s-s_{\mathbb{R}}\in\alpha\mathcal{Q}. Then τ=s¯\tau=\overline{s_{\mathbb{R}}} has order two in (𝒪k/α𝒪k)×(\mathcal{O}_{k}/\alpha\mathcal{O}_{k})^{\times}. Furthermore, for any loxodromic element γΓτ(α)Γ(α)\gamma\in\Gamma_{\tau}(\alpha)\setminus\Gamma(\alpha) we have

|γ|>122d1N(α)2|s|.|\gamma_{\mathbb{R}}|>\frac{1}{2^{2d-1}}\mathrm{N}(\alpha)^{2}-|s_{\mathbb{R}}|.
Proof.

Indeed, since ss is contained in a quaternion algebra we have s=s+αus=s_{\mathbb{R}}+\alpha u for some u𝒬u\in\mathcal{Q} with u+u=0u^{*}+u=0. Hence, the equation 1=ss=s2+α2uu1=ss^{*}=s_{\mathbb{R}}^{2}+\alpha^{2}uu^{*} implies that s2=1(modα)s_{\mathbb{R}}^{2}=1(\mathrm{mod}~{}\alpha). Since the index [Γτ(α):Γ(α)]=2[\Gamma_{\tau}(\alpha):\Gamma(\alpha)]=2, we need to estimate the real part of any product γ=sr\gamma=sr with rΓ(α)r\in\Gamma(\alpha). In this case, by Lemma 4.2, we get

(5.2) γ=s+12α2ζ.\gamma_{\mathbb{R}}=s_{\mathbb{R}}+\frac{1}{2}\alpha^{2}\zeta.

Now, for any non-trivial archimedean place σ\sigma of kk we know that |σ(γ)|1|\sigma(\gamma_{\mathbb{R}})|\leq 1, and |σ(s)|1|\sigma(s_{\mathbb{R}})|\leq 1 ([Mur19, Equation 8]). Therefore, by applying σ\sigma to Equation (5.2) we get

|σ(α2ζ)|=2|σ(γ)σ(s)|4.|\sigma(\alpha^{2}\zeta)|=2|\sigma(\gamma_{\mathbb{R}})-\sigma(s_{\mathbb{R}})|\leq 4.

Once more, by (5.2), and the fact that ζ𝒪k\zeta\in\mathcal{O}_{k}, we obtain that

|γ|\displaystyle|\gamma_{\mathbb{R}}| 12|α|2|ζ||s|\displaystyle\geq\frac{1}{2}|\alpha|^{2}|\zeta|-|s_{\mathbb{R}}|
=12σid|σ(α)2σ(ζ)|N(α)2|N(ζ)||s|\displaystyle=\frac{1}{2\prod_{\sigma\neq id}|\sigma(\alpha)^{2}\sigma(\zeta)|}\mathrm{N}(\alpha)^{2}|\mathrm{N}(\zeta)|-|s_{\mathbb{R}}|
122d1N(α)2|s|.\displaystyle\geq\frac{1}{2^{2d-1}}\mathrm{N}(\alpha)^{2}-|s_{\mathbb{R}}|.

5.2. Embeddings of quadratic fields in 𝒞+(f,k)\mathscr{C}^{+}(f^{\prime},k)

A direct computation shows that the Clifford algebra 𝒞+(f,k)\mathscr{C}^{+}(f^{\prime},k) is a quaternion algebra of Γ\Gamma^{\prime} (see [MR03, Section 12.2]). In fact, it coincides with the invariant quaternion algebra. It is well known that closed geodesics in S=Γ\2S^{\prime}=\Gamma^{\prime}\backslash\mathbb{H}^{2} are related with quadratic extensions of kk that embed in 𝒞+(f,k)\mathscr{C}^{+}(f^{\prime},k). In this subsection we will recall the important properties of this connection that will be useful in the sequel.

For any s𝒞+(f,k)s\in\mathscr{C}^{+}(f^{\prime},k) the reduced trace a=s+sa=s+s^{*}, and the reduced norm b=ssb=ss^{*} are elements of kk. Hence ss is a root of the quadratic polynomial g(x)=x2ax+bk[x]g(x)=x^{2}-ax+b\in k[x]. If gg is irreducible over kk, and LkL\supset k is the quadratic extension where gg splits, then for any fixed root α\alpha of gg in LL, there exists a unique monomorphism ϕ:L𝒞+(f,k)\phi:L\to\mathscr{C}^{+}(f^{\prime},k) such that ϕ(α)=s\phi(\alpha)=s, ϕk\phi\mid_{k} is the identity, and ϕ(σ(x))=ϕ(x)\phi(\sigma(x))=\phi(x)^{*} for the non-trivial Galois automorphism σ:LL\sigma:L\rightarrow L of LL over kk. In particular, via the identification of LL with ϕ(L)\phi(L), the map σ\sigma coincides with the restriction of to LL.

The following proposition is a well-known fact about quaternion algebras which we recall here for reader convenience.

Proposition 5.2.

Let sΓs\in\Gamma^{\prime} be a loxodromic element. There exist a quadratic extension L=k(D)L=k(\sqrt{D}) for some DkD\in k positive, and a kk-homomorphism ψ:L𝒞+(f)\psi:L\to\mathscr{C}^{+}(f^{\prime}) such that s=s+ψ(D).s=s_{\mathbb{R}}+\psi(\sqrt{D}).

Proof.

Consider the irreducible polynomial

(5.3) g(x)=x2(s+s)x+1{}g(x)=x^{2}-(s+s^{*})x+1

over kk. Since ss is loxodromic, gg has two distinct real roots λ\lambda and λ1\lambda^{-1}, and let LL be the quadratic extension k(λ)k(\lambda). Without loss of generality, suppose that |λ|>1|\lambda|>1. We note that λ\lambda and ss are roots of gg, thus there is a unique kk-homomorphism ϕ:L𝒞+(f)\phi:L\to\mathscr{C}^{+}(f^{\prime}) with ϕ(λ)=s\phi(\lambda)=s. If we define θ=λsL\theta=\lambda-s_{\mathbb{R}}\in L, the equality λ+σ(λ)=2s\lambda+\sigma(\lambda)=2s_{\mathbb{R}} implies θ+θ=0\theta+\theta^{*}=0, and then θ2=s21=:D\theta^{2}=s_{\mathbb{R}}^{2}-1=:D. Moreover, ss loxodromic implies that |2s|>2|2s_{\mathbb{R}}|>2, thus D>0D>0. To finish the proof, if θ>0\theta>0 we take ψ=ϕ\psi=\phi, otherwise we consider ψ=ϕ\psi=\phi^{*}. ∎

By Proposition 5.2, we can write s=ψ(λ0)s=\psi(\lambda_{0}) where λ0=s+D\lambda_{0}=s_{\mathbb{R}}+\sqrt{D}. Thus we have an isomorphism between the cyclic group generated by λ0\lambda_{0} in L=k(D)L=k(\sqrt{D}), and the cyclic group generated by ss in Γ\Gamma^{\prime}. For each nn\in\mathbb{N} we can write λ0n+1=tn+unD\lambda_{0}^{n+1}=t_{n}+u_{n}\sqrt{D} with tn,un𝒪kt_{n},u_{n}\in\mathcal{O}_{k}. In the next result we obtain asymptotic relations between un,tnu_{n},t_{n} and ss_{\mathbb{R}}.

Lemma 5.3.

If λ=x0+D\lambda=x_{0}+\sqrt{D} is a unit in 𝒪k[D]\mathcal{O}_{k}[\sqrt{D}] and λn+1=tn+unD,\lambda^{n+1}=t_{n}+u_{n}\sqrt{D}, then for each n1n\geq 1, we have

(5.4) tn=2nx0n+1+O(x0n)t_{n}=2^{n}x_{0}^{n+1}+O(x_{0}^{n})

and

(5.5) un=2nx0n+O(x0n1),u_{n}=2^{n}x_{0}^{n}+O(x_{0}^{n-1}),

where the O notation is considered with respect to x0.x_{0}.

Proof.

Since D=x021D=x_{0}^{2}-1, we have the following relations

tn=(x021)un1+x0tn1 and un=x0un1+tn1.t_{n}=(x_{0}^{2}-1)u_{n-1}+x_{0}t_{n-1}\mbox{ and }u_{n}=x_{0}u_{n-1}+t_{n-1}.

Hence,

tn=x0unun1 and un=2x0un1un2 for all n2.t_{n}=x_{0}u_{n}-u_{n-1}\mbox{ and }u_{n}=2x_{0}u_{n-1}-u_{n-2}\mbox{ for all }n\geq 2.

We prove (5.5) by induction. For n=0,1n=0,1 the relation is trivial. Assuming valid for any 1k<n1\leq k<n, it follows that

un=2x0(2n1x0n1+O(x0n2))2n2x0n2+O(x0n3)=2nx0n+O(x0n1).u_{n}=2x_{0}(2^{n-1}x_{0}^{n-1}+O(x_{0}^{n-2}))-2^{n-2}x_{0}^{n-2}+O(x_{0}^{n-3})=2^{n}x_{0}^{n}+O(x_{0}^{n-1}).

Now, we obtain (5.4) from the relation tn=x0unun1t_{n}=x_{0}u_{n}-u_{n-1}. ∎

Recall that a real algebraic integer λ>1\lambda>1 is a Salem number if λ1\lambda^{-1} is a Galois conjugate of λ\lambda, and the other conjugates of λ\lambda lie on the unit circle. It is well known that the roots of the characteristic polynomial associated to loxodromic elements in Γ\Gamma^{\prime} are Salem numbers (see [GH01]). In this work, we will only deal with Salem numbers of degree four. The next subsection contains the main result that will be necessary for our purpose.

5.3. Salem numbers of degree four

For the interest of this work, it is important to develop some results about Salem numbers of low degree. Let μ\mu be a Salem number of degree four. The field K=(μ+μ1)K=\mathbb{Q}(\mu+\mu^{-1}) is a totally real number subfield of (μ)\mathbb{Q}(\mu), with nontrivial \mathbb{Q}-isomorphism τ:KK\tau:K\to K. Since [(μ):K]=2[\mathbb{Q}(\mu):K]=2, there exists a unique nontrivial KK-isomorphism σ:(μ)(μ)\sigma:\mathbb{Q}(\mu)\to\mathbb{Q}(\mu) such that σ(μ)=μ1\sigma(\mu)=\mu^{-1}. Hence, the four embeddings of (μ)\mathbb{Q}(\mu) into \mathbb{C} are the inclusion, σ\sigma, τ\tau and τ¯\overline{\tau}, where τ\tau is the extension of the nontrivial \mathbb{Q}-morphism of KK into (μ)\mathbb{Q}(\mu). In particular, we can assume that τ(μ)=eiν\tau(\mu)=e^{i\nu} for some ν(0,π)\nu\in(0,\pi).

Now, suppose that there exists D𝒪KD\in\mathcal{O}_{K} such that (μ)=K(D)\mathbb{Q}(\mu)=K(\sqrt{D}) and μ=t+uD\mu=t+u\sqrt{D} for some t,u𝒪Kt,u\in\mathcal{O}_{K}. Since σ\sigma is the non-trivial KK-automorphism of (μ)\mathbb{Q}(\mu) then σ(D)=D\sigma(\sqrt{D})=-\sqrt{D}, and

(5.6) 1=μμ1=μσ(μ)=(t+uD)(tD)=t2u2D1=\mu\cdot\mu^{-1}=\mu\cdot\sigma(\mu)=(t+u\sqrt{D})\cdot(t-\sqrt{D})=t^{2}-u^{2}D
(5.7) 2τ(t)=τ(2t)=τ(μ+σ(μ))=τ(μ)+τ(μ1)=2cos(ν).2\tau(t)=\tau(2t)=\tau(\mu+\sigma(\mu))=\tau(\mu)+\tau(\mu^{-1})=2\cos(\nu).

Thus t2u2D=1t^{2}-u^{2}D=1 and |τ(t)|<1|\tau(t)|<1. For geometric reasons, it is important to get Salem numbers in this form such that τ(t)\tau(t) is not very small. The next proposition shows that we can assume that this property is true up to a small power of μ\mu.

Proposition 5.4.

Let μ>1\mu>1 be a Salem number of degree four. With the previous notations, if μ=t+uD\mu=t+u\sqrt{D}, t,u𝒪Kt,u\in\mathcal{O}_{K} there exists m{0,1,2}m\in\{0,1,2\} such that μm+1=tm+umD\mu^{m+1}=t_{m}+u_{m}\sqrt{D} with τ(tm)2>12.\tau(t_{m})^{2}>\frac{1}{2}.

Proof.

For each m{0,1,2}m\in\{0,1,2\}, if μm+1=tm+umD\mu^{m+1}=t_{m}+u_{m}\sqrt{D} with tm,um𝒪Kt_{m},u_{m}\in\mathcal{O}_{K}, then

μ2(m+1)=(tm2+Dum2)+2tmumD=(2tm21)+2tmumD.\mu^{2(m+1)}=(t_{m}^{2}+Du_{m}^{2})+2t_{m}u_{m}\sqrt{D}=(2t_{m}^{2}-1)+2t_{m}u_{m}\sqrt{D}.

Hence, t2m+1=2tm21t_{2m+1}=2t_{m}^{2}-1 and τ(t2m+1)=2τ(tm)21\tau(t_{2m+1})=2\tau(t_{m})^{2}-1. Hence τ(tm)2>12\tau(t_{m})^{2}>\frac{1}{2} if, and only if, τ(t2m+1)>0.\tau(t_{2m+1})>0. On the other hand,

τ(2t2m+1)=τ(μ2(m+1))+μ2(m+1))=2cos(2(m+1)ν),\tau(2t_{2m+1})=\tau(\mu^{2(m+1)})+\mu^{-2(m+1)})=2\cos(2(m+1)\nu),

where m+1{1,2,3}m+1\in\{1,2,3\}. Then, it remains to show that cos(2kν)>0\cos(2k\nu)>0 for some k{1,2,3}k\in\{1,2,3\}. Indeed, consider the sets S1=(0,π4)(3π4,π),S2=(3π8,5π8),S3=(π4,3π8)(5π8,3π4)S_{1}=(0,\frac{\pi}{4})\cup(\frac{3\pi}{4},\pi),S_{2}=(\frac{3\pi}{8},\frac{5\pi}{8}),S_{3}=(\frac{\pi}{4},\frac{3\pi}{8})\cup(\frac{5\pi}{8},\frac{3\pi}{4}). For each νSj\nu\in S_{j}, we have cos(2jν)>0\cos(2j\nu)>0. The lemma is now proven since [0,π](S1S2S3)[0,\pi]-(S_{1}\cup S_{2}\cup S_{3}) only contains rational multiples of π\pi and Salem numbers do not have conjugates of finite order. ∎

5.4. Congruence subgroups with explicit elements of minimal displacement

The purpose of this section is to construct hyperbolic manifolds with systole lying in a totally geodesic surface. More specifically, we are looking for conditions on τ\tau and α\alpha such that the manifold Mτ,αM_{\tau,\alpha} has a systole in Sτ,αS_{\tau,\alpha} (see Section 5.1). Since sys(Mα,τ)sys(Sα,τ)\mathrm{sys}(M_{\alpha,\tau})\leq\mathrm{sys}(S_{\alpha,\tau}), it is necessary to bound sys(Mα,τ)\mathrm{sys}(M_{\alpha,\tau}) from below. Proposition 4.3 and Proposition 5.1 show that this require a lower bound for the norm of α\alpha in the base field kk, which at the same time implies that the Galois conjugates of α\alpha cannot be very small. We are able to find such α\alpha when kk is a real quadratic number field.

In the sequel, we consider the definitions of Γ\Gamma, Γ\Gamma^{\prime} as in Section 5.1, and kk a real quadratic field, with σ:k\sigma:k\rightarrow\mathbb{R} the nontrivial embedding of kk into \mathbb{R}.

Lemma 5.5.

Let sΓs\in\Gamma^{\prime} be a primitive loxodromic element with s>0s_{\mathbb{R}}>0. There exists l{2,5,8}l\in\{2,5,8\} which depends only on ss_{\mathbb{R}} such that sl+1=(sl+1)+αluls^{l+1}=(s^{l+1})_{\mathbb{R}}+\alpha_{l}u_{l} with ul𝒬u_{l}\in\mathcal{Q}, αl𝒪k\alpha_{l}\in\mathcal{O}_{k} and

1|σ(αl)|5.1\leq|\sigma(\alpha_{l})|\leq 5.

Moreover, if t=st=s_{\mathbb{R}}, then the following asymptotic relation

αl=Clt23(l+1)+O(t23(l+1)2),\alpha_{l}=C_{l}t^{\frac{2}{3}(l+1)}+O\left(t^{\frac{2}{3}(l+1)-2}\right),

holds for some constant Cl>0C_{l}>0 which depends only on ll.

Proof.

By Proposition 5.2 and the discussion in Section 5.1, the element ss corresponds to a Salem number λ0=t0+D\lambda_{0}=t_{0}+\sqrt{D}, and L=k(λ0)L=k(\lambda_{0}) is a quadratic extension of kk. Since kk is a real quadratic number field, λ0\lambda_{0} is a Salem number of degree four. By Proposition 5.4, there exists m{0,1,2}m\in\{0,1,2\} such that λ0m+1=tm+umD\lambda_{0}^{m+1}=t_{m}+u_{m}\sqrt{D} with |σ(tm)|2>12|\sigma(t_{m})|^{2}>\frac{1}{2}. For convenience, we can rewrite

λ=λ0m+1=t+E\lambda=\lambda_{0}^{m+1}=t+\sqrt{E}

where t=tmt=t_{m} and E=um2DE=u_{m}^{2}D. It is straightforward to check that

λ3=(4t33t)+(4t21)E.\lambda^{3}=(4t^{3}-3t)+(4t^{2}-1)\sqrt{E}.

If ll is given by l=3(m+1)1,l=3(m+1)-1, then

λ0l+1=tl+ulD=tl+αlE\lambda_{0}^{l+1}=t_{l}+u_{l}\sqrt{D}=t_{l}+\alpha_{l}\sqrt{E}

where αl=4tm21\alpha_{l}=4t_{m}^{2}-1 and E=um2DE=u_{m}^{2}D. Since 12<|σ(tm)|2<1\frac{1}{2}<|\sigma(t_{m})|^{2}<1 we conclude that 1<|σ(αl)|<51<|\sigma(\alpha_{l})|<5. The asymptotic behaviour of αl\alpha_{l} follows directly from (5.4) and the equality m+1=13(l+1).m+1=\frac{1}{3}(l+1).

We will now prove that for any primitive element sΓs\in\Gamma^{\prime} producing a closed geodesic with length sufficiently large, some power sls^{l} with ll uniformly bounded realizes the systole of some congruence hyperbolic nn-manifold.

Proposition 5.6.

There exists a universal constant L>0L>0 such that for any loxodromic element sΓs\in\Gamma^{\prime} with s>L,s_{\mathbb{R}}>L, we can find l{2,5,8}l\in\{2,5,8\} depending only on ss_{\mathbb{R}} with sl+1(sl+1)αl𝒬s^{l+1}-(s^{l+1})_{\mathbb{R}}\in\alpha_{l}\mathcal{Q}, for some αl𝒪k\alpha_{l}\in\mathcal{O}_{k} and

(sl+1)(r) for all loxodromic element rΓτl(αl),\ell(s^{l+1})\leq\ell(r)\mbox{ for all loxodromic element }r\in\Gamma_{\tau_{l}}(\alpha_{l}),

where τl\tau_{l} is the class of (sl+1)(s^{l+1})_{\mathbb{R}} modulo αl\alpha_{l}.

Proof.

Fix sΓs\in\Gamma^{\prime} loxodromic with real part ss_{\mathbb{R}}. By Lemma 5.5 there exists l{2,5,8}l\in\{2,5,8\} depending only on ss_{\mathbb{R}} such that sl+1(sl+1)αl𝒬s^{l+1}-(s^{l+1})_{\mathbb{R}}\in\alpha_{l}\mathcal{Q} for some αl𝒪k\alpha_{l}\in\mathcal{O}_{k} with 1|σ(αl)|51\leq|\sigma(\alpha_{l})|\leq 5. Let tl=(sl+1)t_{l}=(s^{l+1})_{\mathbb{R}}, it follows from Proposition 5.1 that tl21(modαl)t_{l}^{2}\equiv 1(\mathrm{mod}~{}\alpha_{l}). Hence, if we denote by τl\tau_{l} the class of tlt_{l}, we have

Γτl(αl)=Γ(αl)sl+1Γ(αl).\Gamma_{\tau_{l}}(\alpha_{l})=\Gamma(\alpha_{l})\cup s^{l+1}\Gamma(\alpha_{l}).

If rΓ(αl)r\in\Gamma(\alpha_{l}), since |αl|1|\alpha_{l}^{\prime}|\geq 1, Proposition 4.3 gives us that

|r|18|σ(αl)|2αl2118αl21.|r_{\mathbb{R}}|\geq\frac{1}{8}|\sigma(\alpha_{l})|^{2}\alpha_{l}^{2}-1\geq\frac{1}{8}\alpha_{l}^{2}-1.

Since (sl+1)=2cosh1(|tl|)\ell(s^{l+1})=2\cosh^{-1}(|t_{l}|), by Proposition 3.1, in order to show that (r)(sl+1)\ell(r)\geq\ell(s^{l+1}) it is sufficient to guarantee that |r|tl|r_{\mathbb{R}}|\geq t_{l}. By Lemma 5.5 and Equation (5.4) we have

18αl21=C(s)43(l+1)+O((s)43(l+1)4) and tl2l(s)l+1.\frac{1}{8}\alpha_{l}^{2}-1=C(s_{\mathbb{R}})^{\frac{4}{3}(l+1)}+O\left((s_{\mathbb{R}})^{\frac{4}{3}(l+1)-4}\right)\mbox{ and }t_{l}\sim 2^{l}(s_{\mathbb{R}})^{l+1}.

Hence, 18αl21tl\frac{1}{8}\alpha_{l}^{2}-1\geq t_{l} whenever ss_{\mathbb{R}} is sufficiently large.

Analogously, if rsl+1Γ(αl)r\in s^{l+1}\Gamma(\alpha_{l}) and rsl+1r\neq s^{l+1}, we have by Lemma 4.2 that

|r|18|σ(αl)|2αl2tl18αl2tl.|r_{\mathbb{R}}|\geq\frac{1}{8}|\sigma(\alpha_{l})|^{2}\alpha_{l}^{2}-t_{l}\geq\frac{1}{8}\alpha_{l}^{2}-t_{l}.

And |r|>tl|r_{\mathbb{R}}|>t_{l} whenever αl2>16tl,\alpha_{l}^{2}>16t_{l}, which holds whenever ss_{\mathbb{R}} is large enough. ∎

The previous result has Theorem C mentioned in Introduction as geometric counterpart. We recall it as a corollary, adapted to the terminology used so far.

Corollary 5.7.

With the notation as in Proposition 5.6, the hyperbolic manifold Mτl,αl=Γτl(αl)\nM_{\tau_{l},\alpha_{l}}=\Gamma_{\tau_{l}}(\alpha_{l})\backslash\mathbb{H}^{n} contains the totally geodesic surface Sτl,αl=Γτl(αl)\2S_{\tau_{l},\alpha_{l}}=\Gamma^{\prime}_{\tau_{l}}(\alpha_{l})\backslash\mathbb{H}^{2} with

sys(Sτl,αl)=sys(Mτl,αl).\mathrm{sys}(S_{\tau_{l},\alpha_{l}})=\mathrm{sys}(M_{\tau_{l},\alpha_{l}}).
Proof.

Indeed, by Proposition 5.6, we exhibit a loxodromic element in Γτl(αl)\Gamma^{\prime}_{\tau_{l}}(\alpha_{l}), namely sl+1s^{l+1}, of minimal displacement in Γτl(αl)\Gamma_{\tau_{l}}(\alpha_{l}). In particular this element is also of minimal displacement in Γτl(αl)\Gamma^{\prime}_{\tau_{l}}(\alpha_{l}). Hence, its length is at same time the systolic length of Sτl,αlS_{\tau_{l},\alpha_{l}} and Mτl,αl.M_{\tau_{l},\alpha_{l}}.

6. Proof of Theorem A and Theorem B

Any closed geodesic on a hyperbolic manifold MM is parametrized by a constant speed curve from the circle to MM and we can identify the geodesic with the equivalence class of such parametrization up to reparametrization. Let γ:𝐒1=/M\gamma:\mathbf{S}^{1}=\mathbb{R}/\mathbb{Z}\rightarrow M be a closed geodesic, we say that γ\gamma is primitive if γ\gamma is injective, i.e. if γ\gamma is an embedding. Any closed geodesic δ\delta is a kk-fold iterate of some primitive geodesic γ\gamma, i.e. there exists kk\in\mathbb{N} such that δ(t)=γ(kt)\delta(t)=\gamma(kt) (up to reparametrizations of δ\delta and γ\gamma). We note that kk is uniquely determined by the relation (δ)=k(γ)\ell(\delta)=k\ell(\gamma) and because of this, we call it as the order of δ\delta.

Let π:MN\pi:M\to N be a covering map between two hypebolic nn-orbifolds MM and NN, we recall that a closed geodesic c:𝐒1Nc:\mathbf{S}^{1}\to N lifts to MM if there is a closed geodesic c~:𝐒1M\tilde{c}:\mathbf{S}^{1}\to M such that c=πc~c=\pi\circ\tilde{c}. In this case, we say that c~\tilde{c} is a lift of cc.
We note that the deck group Deck(π)={gIsom(M)πg=π}\mathrm{Deck}(\pi)=\{g\in\mathrm{Isom}(M)\mid\pi\circ g=\pi\} is always finite whenever MM and NN have finite volume. In the sequel, we consider the natural action of Deck(π)\mathrm{Deck(\pi)} on the set of closed geodesics of MM.

Lemma 6.1.

Let π:MN\pi:M\to N be a covering map between the hyperbolic nn-orbifolds MM and NN of finite volume, and let G=Deck(π)G=\mathrm{Deck}(\pi).

  1. (1)

    If γ~1,γ2~\tilde{\gamma}_{1},\tilde{\gamma_{2}} are closed geodesics on MM which are liftings of two distinct closed geodesics γ1,γ2\gamma_{1},\gamma_{2} on NN respectively, then the orbits Gγ~1G\cdot\tilde{\gamma}_{1} and Gγ~2G\cdot\tilde{\gamma}_{2} are disjoint.

  2. (2)

    If γ\gamma is a closed geodesic on NN of order kk which lifts, then for any lifting γ~\tilde{\gamma} its isotropy group Gγ~G_{\tilde{\gamma}} has at most kk elements.

Proof.

Indeed, if γ~1=gγ~2\tilde{\gamma}_{1}=g\circ\tilde{\gamma}_{2} (up to reparametrization of γ1~\tilde{\gamma_{1}}, and γ2~\tilde{\gamma_{2}}) then γ1=πγ~1=πgγ~2=πγ~2=γ2,\gamma_{1}=\pi\circ\tilde{\gamma}_{1}=\pi\circ g\circ\tilde{\gamma}_{2}=\pi\circ\tilde{\gamma}_{2}=\gamma_{2}, which proves (1)(1). For (2)(2) we can suppose that M=Λ\nM=\Lambda^{\prime}\backslash\mathbb{H}^{n} and N=Λ\nN=\Lambda\backslash\mathbb{H}^{n} where Λ<Λ\Lambda^{\prime}<\Lambda. With this identification, the group GG can be seen as NΛ(Λ)/Λ\mathrm{N}_{\Lambda}(\Lambda^{\prime})/\Lambda^{\prime}, where NΛ(Λ)\mathrm{N}_{\Lambda}(\Lambda^{\prime}) denotes the normalizer of Λ\Lambda^{\prime} in Λ\Lambda. Moreover, a closed geodesic on MM can be associated with a conjugacy class [γ][\gamma^{\prime}] of a loxodromic element γΛ\gamma^{\prime}\in\Lambda^{\prime}. The action of GG on the set of closed geodesics is given by λΛ[γ]=[λ1γλ]\lambda\Lambda^{\prime}\cdot[\gamma^{\prime}]=[\lambda^{-1}\gamma^{\prime}\lambda]. If [γ][\gamma^{\prime}] denotes a closed geodesic of order nn on NN, we can use the same notation for its lifting on MM since γΛ\gamma\in\Lambda^{\prime}. Hence λΛ[γ]=[γ]\lambda\Lambda^{\prime}\cdot[\gamma^{\prime}]=[\gamma^{\prime}] means that λ1γλ=λ11γλ1\lambda^{-1}\gamma^{\prime}\lambda=\lambda_{1}^{-1}\gamma^{\prime}\lambda_{1} for some λ1Λ\lambda_{1}\in\Lambda^{\prime}, then λ1λ1\lambda_{1}\lambda^{-1} commutes with γ\gamma^{\prime}. By hypothesis, γ=η0n\gamma^{\prime}=\eta_{0}^{n}, and for results in hyperbolic geometry we have that the centralizer of γ\gamma^{\prime} is the cyclic group generated by η0\eta_{0}. Therefore, λΛ{η0iΛ0ik1}\lambda\Lambda^{\prime}\in\{\eta_{0}^{i}\Lambda^{\prime}\mid 0\leq i\leq k-1\}. ∎

Remark 1.

Let MM be a closed hyperbolic nn-manifold and let ΣM\Sigma\subset M be a totally geodesic submanifold. If α,β\alpha,\beta are distinct primitive closed geodesics on Σ\Sigma then α\alpha and β\beta are distinct primitive closed geodesics in MM. Indeed, if α\alpha is a kk-folded iterated of α0\alpha_{0} for some primitive α0:𝐒1M\alpha_{0}:\mathbf{S}^{1}\to M, we have α0(0)Σ\alpha_{0}(0)\in\Sigma and α0(0)Tα0(0)Σ\alpha_{0}^{\prime}(0)\in T_{\alpha_{0}(0)}\Sigma, thus α0\alpha_{0} is a closed geodesic on Σ\Sigma and then α=α0\alpha=\alpha_{0}. In particular, if sys(Σ)=sys(M)\mathrm{sys}(\Sigma)=\mathrm{sys}(M) then Kiss(M)Kiss(Σ)\mathrm{Kiss}(M)\geq\mathrm{Kiss}(\Sigma).

Proof of Theorem A.

For each n2n\geq 2 we can consider a fixed closed arithmetic hyperbolic nn-manifold of the first type MM. By [Xue92, Main Theorem] there exists a sequence MjM_{j} of congruence coverings of MM such that β1(Mj)\beta_{1}(M_{j})\to\infty, since vol(Mj)\mathrm{vol}(M_{j})\to\infty. Moreover, it follows from [Mur17, Theorem 6.1] that sys(Mj)\mathrm{sys}(M_{j})\to\infty. Hence, given A>0A>0 we can suppose that we have a closed hyperbolic nn-manifold MM with sys(M)>A\mathrm{sys}(M)>A and β1(M)>0\beta_{1}(M)>0.

If we write M=Γ\nM=\Gamma\backslash\mathbb{H}^{n} where Γπ1(M)\Gamma\simeq\pi_{1}(M), then there exists an epimorphism ϕ:Γ\phi:\Gamma\to\mathbb{Z}. Let NtMN_{t}\to M be the cyclic cover of degree tt obtained by the kernel of the map

Γ/t.\Gamma\to\mathbb{Z}\to\mathbb{Z}/t\mathbb{Z}.

Let γΓ\gamma\in\Gamma be a nontrivial loxodromic element of minimal displacement satisfying ϕ(γ)=0\phi(\gamma)=0. If α\alpha corresponds to the closed geodesic induced by γ\gamma, then α\alpha is lifted for all covering NtN_{t}, thus Asys(Nt)(α)A\leq\mathrm{sys}(N_{t})\leq\ell(\alpha) for all t2t\geq 2.

For any tt, there exists a closed geodesic αt\alpha_{t} on MM whose lifting αt~\tilde{\alpha_{t}} in NtN_{t} satisfies sys(Nt)=(αt~)=(αt)\mathrm{sys}(N_{t})=\ell(\tilde{\alpha_{t}})=\ell(\alpha_{t}). Let Gt/tG_{t}\simeq\mathbb{Z}/t\mathbb{Z} be the deck group of the covering NtMN_{t}\to M. We claim that the isotropy subgroup of αt~\tilde{\alpha_{t}} under the action of GtG_{t} has cardinality at most cc for some constant c>0c>0 which does not depend on tt. Indeed, αt=βm\alpha_{t}=\beta^{m} for some primitive closed geodesic β\beta. Hence,

msys(M)m(β)=(αt)=sys(Nt)(α).m\cdot\mathrm{sys}(M)\leq m\cdot\ell(\beta)=\ell(\alpha_{t})=\mathrm{sys}(N_{t})\leq\ell(\alpha).

Therefore, by Lemma 6.1 the order of the isotropy group of αt~\tilde{\alpha_{t}} by the action of the deck group GtG_{t} is bounded from above by c=(α)sys(M)>0.c=\frac{\ell(\alpha)}{\mathrm{sys}(M)}>0.

For any pp prime with p>cp>c we conclude that the orbit Gpαp~G_{p}\cdot\tilde{\alpha_{p}} has pp elements. Putting all the information above together, we have proved that for any p>cp>c, Kiss(Np)p\mathrm{Kiss}(N_{p})\geq p and vol(Np)=pvol(M)\mathrm{vol}(N_{p})=p\cdot\mathrm{vol}(M). Thus, the theorem is proved with B=(α)B=\ell(\alpha) and C=vol(M)C=\mathrm{vol}(M). ∎

We are now ready to present the proof of Theorem B.

Proof of Theorem B.

Let ϕ(x)\phi^{\prime}(x) denote the number of conjugacy classes of loxodromic elements in Γ\Gamma^{\prime} with reduced trace equal to xx. By the Prime Geodesic Theorem there is a sequence xix_{i}\rightarrow\infty such that (see [Sch95])

ϕ(xi)xilog(xi).\phi^{\prime}(x_{i})\geq\frac{x_{i}}{\log(x_{i})}.

For each ii with xix_{i} large enough, xi=2(si)x_{i}=2(s_{i})_{\mathbb{R}} for some loxodromic element siΓs_{i}\in\Gamma^{\prime}. Let us now consider

mi=min{l|l{2,5,8}satisfiesProposition 5.6 for (si)}.m_{i}=\textrm{min}\{l\ |\ l\in\{2,5,8\}\ \textrm{satisfies}\ \textrm{Proposition~{}\ref{inedisp}}\mbox{ for }(s_{i})_{\mathbb{R}}\}.

Furthermore, take τmi\tau_{m_{i}} and αmi\alpha_{m_{i}} as given in Lemma 5.5 and Proposition 5.6. Then the manifold Mi=Γτmi(αmi)\nM_{i}=\Gamma_{\tau_{m_{i}}}(\alpha_{m_{i}})\backslash\mathbb{H}^{n} and the totally geodesic surface Si=Γτmi(αmi)\2S_{i}=\Gamma^{\prime}_{\tau_{m_{i}}}(\alpha_{m_{i}})\backslash\mathbb{H}^{2} satisfy (see Remark 1)

(6.1) Kiss(Mi)Kiss(Si)xilog(xi).\mathrm{Kiss}(M_{i})\geq\mathrm{Kiss}(S_{i})\geq\frac{x_{i}}{\log(x_{i})}.

On the other hand, take the isometry group Gi=Γ/Γτmi(αmi)G_{i}=\Gamma/\Gamma_{\tau_{m_{i}}}(\alpha_{m_{i}}) acting on the set of closed geodesics of MiM_{i}. By Lemma 6.1(1), if we denote by γ1,,γKiss(Si)\gamma_{1},\dots,\gamma_{\mathrm{Kiss}(S_{i})} the systoles of SiS_{i} embedded in MiM_{i}, the orbit sets GiγjG_{i}\gamma_{j}, j{1,,Kiss(Si)}j\in\{1,\dots,\mathrm{Kiss}(S_{i})\} are pairwise disjoint. It follows that

(6.2) Kiss(Mi)\displaystyle\mathrm{Kiss}(M_{i}) j=1Kiss(Si)|Giγj|\displaystyle\geq\sum_{j=1}^{\mathrm{Kiss}(S_{i})}|G_{i}\gamma_{j}|
(6.3) =j=1Kiss(Si)|Gi||(Gi)γj|,\displaystyle=\sum_{j=1}^{\mathrm{Kiss}(S_{i})}\dfrac{|G_{i}|}{|(G_{i})_{\gamma_{j}}|},

where (Gi)γj(G_{i})_{\gamma_{j}} denotes the isotropy group of γj\gamma_{j} under the action of GiG_{i}. By Lemma 6.1(2), |(Gi)γj||(G_{i})_{\gamma_{j}}| is at most the order of γj\gamma_{j}, and this is smaller than a fixed constant C>0C>0 since mi8m_{i}\leq 8. Therefore, we get from (6.1) and (6.3) that

Kiss(Mi)CKiss(Si)|Gi|Cxilog(xi)|Gi|.\mathrm{Kiss}(M_{i})\geq C\mathrm{Kiss}(S_{i})\cdot|G_{i}|\geq C\frac{x_{i}}{\log(x_{i})}|G_{i}|.

Since vol(Mi)=[Γ:Γτmi(αmi)]vol(Γ\n)=|Gi|vol(Γ\n)\mathrm{vol}(M_{i})=[\Gamma:\Gamma_{\tau_{m_{i}}}(\alpha_{m_{i}})]\mathrm{vol}(\Gamma\backslash\mathbb{H}^{n})=|G_{i}|\mathrm{vol}(\Gamma\backslash\mathbb{H}^{n}), the above inequality becomes

(6.4) Kiss(Mi)Cvol(Γ\n)xilog(xi)vol(Mi).\mathrm{Kiss}(M_{i})\geq\frac{C}{\mathrm{vol}(\Gamma\backslash\mathbb{H}^{n})}\frac{x_{i}}{\log(x_{i})}\mathrm{vol}(M_{i}).

The goal now is to bound xilog(xi)\frac{x_{i}}{\log(x_{i})} from below in terms of vol(Mi)\mathrm{vol}(M_{i}). Since Γτmi(αmi)\Gamma_{\tau_{m_{i}}}(\alpha_{m_{i}}) has index two in Γ(αmi)\Gamma(\alpha_{m_{i}}), then

(6.5) vol(Mi)=[Γ:Γτmi(αmi)]vol(Γ\n)=vol(Γ\n)2[Γ:Γ(αmi)].\mathrm{vol}(M_{i})=[\Gamma:\Gamma_{\tau_{m_{i}}}(\alpha_{m_{i}})]\mathrm{vol}(\Gamma\backslash\mathbb{H}^{n})=\frac{\mathrm{vol}(\Gamma\backslash\mathbb{H}^{n})}{2}[\Gamma:\Gamma(\alpha_{m_{i}})].

Note that the norm of the ideal (αmi)(\alpha_{m_{i}}) goes to infinity (see Lemma 5.5), thus we can apply the results in [Mur19, Section 5] along with the fact that mi8m_{i}\leq 8, to obtain the following

N(αmi)n(n+1)2\displaystyle\leq\mathrm{N}(\alpha_{m_{i}})^{\frac{n(n+1)}{2}} =O(|αmi|n(n+1)2)\displaystyle=O\left(|\alpha_{m_{i}}|^{\frac{n(n+1)}{2}}\right)
=O((xi2(mi+1)3)n(n+1)2)\displaystyle=O\left(\left(x_{i}^{\frac{2(m_{i}+1)}{3}}\right)^{\frac{n(n+1)}{2}}\right)
=O(xi3n(n+1)).\displaystyle=O\left(x_{i}^{3n(n+1)}\right).

By putting together the estimate obtained above with Equation (6.5), we find the lower bound

(6.6) xiC1vol(Mi)13n(n+1)x_{i}\geq C_{1}\cdot\mathrm{vol}(M_{i})^{\frac{1}{3n(n+1)}}

for some constant C1>0C_{1}>0. Finally, since the function xxlog(x)x\mapsto\frac{x}{\log(x)} is increasing for large xx, by  (6.4) we get that

Kiss(Mi)C2vol(Mi)1+13n(n+1)log(vol(Mi)),\mathrm{Kiss}(M_{i})\geq C_{2}\ \frac{\mathrm{vol}(M_{i})^{1+\frac{1}{3n(n+1)}}}{\log(\mathrm{vol}(M_{i}))},

for a constant C2>0C_{2}>0. This ends the proof of Theorem B. ∎

7. The three-dimensional case

7.1. Hyperbolic 3-manifolds

In the three-dimensional case it will be convenient to consider the upper-half model of the hyperbolic 33-space given by

3={(z,t)×;t>0},\mathbb{H}^{3}=\{(z,t)\in\mathbb{C}\times\mathbb{R};t>0\},

with the Riemannian metric ds2=dz2+dt2t2ds^{2}=\frac{dz^{2}+dt^{2}}{t^{2}}. Note that we realize 3\mathbb{H}^{3} as a subset of Hamilton’s quaternion algebra

={a+bi+cj+dk|a,b,c,d,i2=j2=k2=1},\mathcal{H}=\{a+bi+cj+dk|a,b,c,d\in\mathbb{R},i^{2}=j^{2}=k^{2}=-1\},

where we represent a point P3P\in\mathbb{H}^{3} as a Hamiltonian quaternion

P=(z,t):=x+yi+tj,P=(z,t):=x+yi+tj,

where z=x+iyz=x+iy. Moreover G=SL(2,)G=\mathrm{SL}(2,\mathbb{C}) acts by isometries on 3\mathbb{H}^{3}, and this action is described as follows. For each M=(abcd)SL(2,)M=\left(\begin{array}[]{cc}a&b\\ c&d\\ \end{array}\right)\in\mathrm{SL}(2,\mathbb{C})

PMP:=(aP+b)(cP+d)1,P\mapsto MP:=(aP+b)(cP+d)^{-1},

where the inverse is taken in the skew field of Hamilton’s quaternions. This action is not faithful since I-\mathrm{I} acts trivially, but the finite quotient PSL(2,)=SL(2,)/{±I}\mathrm{PSL(2,\mathbb{C})}=\mathrm{SL}(2,\mathbb{C})/\{\pm\mathrm{I}\} is isomorphic to Isom+(3)\mathrm{Isom}^{+}(\mathbb{H}^{3}) (see [EGM13, Chapter. 1]). As we have already observed with Spin groups, there is no loss of generality in identifying elements in SL(2,)\mathrm{SL}(2,\mathbb{C}) with their projection in PSL(2,)\mathrm{PSL}(2,\mathbb{C}). We recall that an element γSL(2,)\gamma\in\mathrm{SL}(2,\mathbb{C}) is said to be:

  • elliptic if γ\gamma is conjugate to (η00η1)\left(\begin{array}[]{cc}\eta&0\\ 0&\eta^{-1}\\ \end{array}\right),  with |η|=1,η±1.\mbox{ with }|\eta|=1,\eta\neq\pm 1.

  • parabolic if γ\gamma is conjugate to (1z01)\left(\begin{array}[]{cc}1&z\\ 0&1\\ \end{array}\right),  for some z,z0.\mbox{ for some }z\in\mathbb{C},\ z\neq 0.

  • loxodromic if γ\gamma is conjugate to (reiθ00r1eiθ)\left(\begin{array}[]{cc}re^{i\theta}&0\\ 0&r^{-1}e^{-i\theta}\\ \end{array}\right), r>1,r,θ.r>1,r,\theta\in\mathbb{R}.

We define the trace of γ=±(abcd)PSL(2,)\gamma=\pm\left(\begin{array}[]{cc}a&b\\ c&d\\ \end{array}\right)\in\mathrm{PSL}(2,\mathbb{C}) as

tr(γ):=μ(a+d),\mathrm{tr}(\gamma):=\mu\cdot(a+d),

where μ{±1}\mu\in\{\pm 1\} is chosen so that tr(γ)=reiθ\mathrm{tr}(\gamma)=re^{i\theta} with r0r\geq 0 and θ[0,π)\theta\in[0,\pi).

Consider the eigenvalues of γ\gamma as the eigenvalues of a lift to SL(2,)\mathrm{SL}(2,\mathbb{C}). Hence the roots of the characteristic polynomial relative to γ\gamma are

λγ±=tr(γ)±(tr(γ))242.\lambda^{\pm}_{\gamma}=\frac{\mathrm{tr}(\gamma)\pm\sqrt{(\mathrm{tr}(\gamma))^{2}-4}}{2}.

These are the eigenvalues of one of the representatives of γSL(2,)\gamma\in\mathrm{SL}(2,\mathbb{C}), and we denote by λγ\lambda_{\gamma} the eigenvalue with norm greater than one. We also choose the branch of the argument function Arg(z)\mathrm{Arg}(z) on (,0]\mathbb{C}\setminus(-\infty,0] with Arg(z)(π,π)\mathrm{Arg}(z)\in(-\pi,\pi). It is well-known that λγ\lambda_{\gamma} determines the translation length (γ)\ell(\gamma) of a γ\gamma. More precisely,

(7.1) (γ)=2log(|λγ|).\ell(\gamma)=2\log(|\lambda_{\gamma}|).

The holonomy of γ\gamma is defined as

(7.2) θ(γ):=2Arg(λγ),\theta(\gamma):=2\mathrm{Arg}(\lambda_{\gamma}),

We end this section by recalling how l(γ)l(\gamma) can be determined from tr(γ)\mathrm{tr}(\gamma).

Proposition 7.1.

For any loxodromic element γSL(2,)\gamma\in\mathrm{SL}(2,\mathbb{C}) we have

4cosh(l(γ)2)=|tr(γ)2|+|tr(γ)+2|.4\cosh\left(\frac{l(\gamma)}{2}\right)=|\mathrm{tr}(\gamma)-2|+|\mathrm{tr}(\gamma)+2|.

In particular

4cosh(l(γ))=|tr(γ)2|+|tr(γ)24|.4\cosh(l(\gamma))=|\mathrm{tr}(\gamma)^{2}|+|\mathrm{tr}(\gamma)^{2}-4|.
Proof.

See [DM21, Proposition. 2.1], c.f [Gen15, Lemma. 5.1]. ∎

7.2. Arithmetic Kleinian groups

A Kleinian group is a discrete group of PSL(2,)\mathrm{PSL}(2,\mathbb{C}). Let kk be a number field with exactly one complex place and let 𝒜\mathcal{A} be a quaternion algebra over kk ramified at all real places. A Kleinian group Γ\Gamma is arithmetic if it is commensurable with the projection Pρ(𝒪1):=ρ(𝒪1)/{±I}P\rho(\mathcal{O}^{1}):=\rho(\mathcal{O}^{1})/\{\pm\mathrm{I}\}, where ρ\rho be a kk-embedding of 𝒜\mathcal{A} into M2()M_{2}(\mathbb{C}) and 𝒪1\mathcal{O}^{1} denotes the group of elements of reduced norm one of an order 𝒪\mathcal{O} of 𝒜\mathcal{A}. When ΓPρ(𝒪1)\Gamma\subset P\rho(\mathcal{O}^{1}) we say that Γ\Gamma is derived from a quaternion algebra.

A hyperbolic 33-orbifold Γ\3\Gamma\backslash\mathbb{H}^{3} is arithmetic if Γ\Gamma is an arithmetic Kleinian group. For explicity examples, consider for each square-free positive integer d>0,d>0, the Kleinian group PSL(2,𝒪d)\mathrm{PSL}(2,\mathcal{O}_{d}), where 𝒪d\mathcal{O}_{d} is the ring of integers of (d).\mathbb{Q}(\sqrt{-d}). These groups are known as Bianchi groups (see [EGM13, Chapter 7]).

Let 𝒜\mathcal{A} be a quaternion algebra over a number field kk, with ring of integers 𝒪k\mathcal{O}_{k}. For any ideal I𝒪kI\subset\mathcal{O}_{k}, and any order 𝒪𝒜\mathcal{O}\subset\mathcal{A} we have an ideal I𝒪I\mathcal{O} defined by

I𝒪={jtjwj|tjIandwj𝒪}.I\mathcal{O}=\left\{\sum_{j}t_{j}w_{j}|\ t_{j}\in I\ \textrm{and}\ w_{j}\in\mathcal{O}\right\}.

The principal congruence subgroup of 𝒪1\mathcal{O}^{1} of level II is then given by

𝒪1(I)={γ𝒪1|γ1I𝒪}.\mathcal{O}^{1}(I)=\{\gamma\in\mathcal{O}^{1}\ |\ \gamma-1\in I\mathcal{O}\}.

Suppose that Γ<PSL(2,)\Gamma<\mathrm{PSL}(2,\mathbb{C}) is a Kleinian group. We will denote by Γ~\tilde{\Gamma} the preimage of Γ\Gamma by the natural projection of SL(2,)\mathrm{SL}(2,\mathbb{C}) into PSL(2,)\mathrm{PSL}(2,\mathbb{C}). In particular, if Γ\Gamma is arithmetic derived from a quaternion algebra, then Γ~<𝒪1\tilde{\Gamma}<\mathcal{O}^{1} for some order 𝒪\mathcal{O}. In this way, for any ideal II we define the principal congruence subgroup of Γ\Gamma of level II as the projection of Γ~𝒪1(I)\tilde{\Gamma}\cap\mathcal{O}^{1}(I) onto Γ\Gamma. If II is a principal ideal, generated by α𝒪k\alpha\in\mathcal{O}_{k}, we denote Γ(I)\Gamma(I) by Γ(α)\Gamma(\alpha) instead of Γ(α)\Gamma(\langle\alpha\rangle). The key fact about congruence subgroups is the following lemma (Compare with  Lemma (4.1)).

Lemma 7.2.

For any γ𝒪1(I)\gamma\in\mathcal{O}^{1}(I) we have tr(γ)2modI2.\mathrm{tr}(\gamma)\equiv 2\ \mathrm{mod}\ I^{2}.

Proof.

Let γ𝒪1(I)\gamma\in\mathcal{O}^{1}(I), by definition we can write γ=1+η\gamma=1+\eta, with ηI𝒪\eta\in I\mathcal{O}. Since tr(η)=η+ηI\mathrm{tr}(\eta)=\eta+\eta^{*}\in I and Nred(γ)=ηηI2\mathrm{N_{red}}(\gamma)=\eta\eta^{*}\in I^{2} (see [KSV07, Lemma 3.3]), we have

1=γγ=1+tr(η)+Nred(γ).1=\gamma\gamma^{*}=1+\mathrm{tr}(\eta)+\mathrm{N_{red}}(\gamma).

Therefore tr(η)I2\mathrm{tr}(\eta)\in I^{2}, and then tr(γ)2modI2\mathrm{tr}(\gamma)\equiv 2\ \mathrm{mod}\ I^{2}. ∎

7.3. Displacement estimates for congruence subgroups

We will now construct hyperbolic 3-orbifolds for which we can determine their set of systoles. Before that, we introduce some notation that will be convenient for this purpose.

Let T:T:\mathbb{C}^{*}\to\mathbb{C}^{*} be the surjective holomorphic map given by T(z)=z+z1T(z)=z+z^{-1}, and let V=(,0]V=\mathbb{C}\setminus(-\infty,0]. We observe that, if γPSL(2,)\gamma\in\mathrm{PSL}(2,\mathbb{C}) is loxodromic, and t=tr(γ)t=\mathrm{tr}(\gamma) with largest eigenvalue λ\lambda, then T(λ)=tT(\lambda)=t. On the other hand, Arg(T(z))\mathrm{Arg}(T(z)) is a continuous map from T1(V)T^{-1}(V) to (π,π)(-\pi,\pi). If we write zz in its polar coordinates with |Arg(z)|<π|\mathrm{Arg}(z)|<\pi, we have that

T(z)=(|z|+|z|1)cos(Arg(z))+i(|z||z|1)sin(Arg(z)).T(z)=(|z|+|z|^{-1})\cos(\mathrm{Arg}(z))+i(|z|-|z|^{-1})\sin(\mathrm{Arg}(z)).

Thus, |Arg(T(z))|=π2|\mathrm{Arg}(T(z))|=\frac{\pi}{2} if, and only if, |Arg(z)|=π2|\mathrm{Arg}(z)|=\frac{\pi}{2}. Moreover, if |z|>1|z|>1 and |Arg(z)|π2|\mathrm{Arg}(z)|\neq\frac{\pi}{2} we obtain

(7.3) tan(Arg(T(z)))=|z||z|1|z|+|z|1tan(Arg(z)).\tan(\mathrm{Arg}(T(z)))=\frac{|z|-|z|^{-1}}{|z|+|z|^{-1}}\tan(\mathrm{Arg}(z)).
Proposition 7.3.

Let kk be an imaginary quadratic field, and Γ\Gamma an arithmetic Kleinian group derived from a quaternion algebra over kk. Then, there exist L,ϵ>0L,\epsilon>0 such that if γΓ\gamma\in\Gamma is a loxodromic element with

(γ)>L and 0hol(γ)<ϵ,\ell(\gamma)>L\mbox{ and }0\leq\mathrm{hol}(\gamma)<\epsilon,

then γ2\gamma^{2} realizes the systole of Γ(tr(γ))\3\Gamma(\mathrm{tr}(\gamma))\backslash\mathbb{H}^{3}.

Proof.

By hypothesis, there exist a quaternion algebra AA over kk, an order 𝒪\mathcal{O} of AA and a kk-monomorphism of algebras ρ:ASL(2,)\rho:A\to\mathrm{SL}(2,\mathbb{C}) such that Γ~<ρ(𝒪1)\tilde{\Gamma}<\rho(\mathcal{O}^{1}). Hence, we can assume that Γ~<𝒪1.\tilde{\Gamma}<\mathcal{O}^{1}.

Suppose that tr(γ)=t𝒪k\mathrm{tr}(\gamma)=t\in\mathcal{O}_{k}, and let γ~\tilde{\gamma} be a representative of γ\gamma in Γ~\tilde{\Gamma}. By definition

Γ~(t)=Γ~𝒪1(t),1,\tilde{\Gamma}(t)=\langle\tilde{\Gamma}\cap\mathcal{O}^{1}(t),-1\rangle,

and Γ(t)=Γ~(t)/{±1}\Gamma(t)=\tilde{\Gamma}(t)/\{\pm 1\}. Since γ~2tγ~+1=0\tilde{\gamma}^{2}-t\tilde{\gamma}+1=0 in 𝒪\mathcal{O} we get that γ~2Γ~(t)-\tilde{\gamma}^{2}\in\tilde{\Gamma}(t), and then γ2Γ(t)\gamma^{2}\in\Gamma(t).

Now, let λ\lambda\in\mathbb{C} be the eigenvalue of γ\gamma with |λ|>1|\lambda|>1. By  (7.1) we have (γ2)=2(γ)=4log(|λ|)\ell(\gamma^{2})=2\ell(\gamma)=4\log(|\lambda|). Since γ2Γ(t)\gamma^{2}\in\Gamma(t) we get that

sys(Γ(tr(γ))\3)4log(|λ|)=(γ2).\mathrm{sys}(\Gamma(\mathrm{tr}(\gamma))\backslash\mathbb{H}^{3})\leq 4\log(|\lambda|)=\ell(\gamma^{2}).

Our purpose is to give conditions on tt such that this inequality becomes an equality. Let ηΓ(t)\eta\in\Gamma(t) be a loxodromic element. There exists a representative, say η~\tilde{\eta}, of η\eta with η~Γ~𝒪1(t)\tilde{\eta}\in\tilde{\Gamma}\cap\mathcal{O}^{1}(t). It follows from Lemma 7.2 that, if τ=tr(η)\tau=\mathrm{tr}(\eta), then τ=2+t2ζ\tau=2+t^{2}\zeta for some ζ𝒪k\zeta\in\mathcal{O}_{k} and ζ0\zeta\neq 0. If τ=T(μ)\tau=T(\mu) with |μ|>1|\mu|>1, since t2=λ2+λ2+2t^{2}=\lambda^{2}+\lambda^{-2}+2, then τ=2+t2ζ\tau=2+t^{2}\zeta can be rewritten as

(7.4) T(μ)=2(ζ+1)+ζT(λ2).T(\mu)=2(\zeta+1)+\zeta T(\lambda^{2}).

By (7.1) it is sufficient to show that |μ|>|λ|2|\mu|>|\lambda|^{2}. We will divide our analysis in two cases:

Case 1. ζ𝒪k\zeta\notin\mathcal{O}^{*}_{k}: Since kk is a quadratic field, we have |ζ|22|\zeta|^{2}\geq 2, and then |ζ|2|\zeta|\geq\sqrt{2}. Firstly, we can rewrite (7.4) as

(7.5) T(μ)\displaystyle T(\mu) =ζλ2(1+(λ2)2+2(λ2)1+2(λ2ζ1))\displaystyle=\zeta\lambda^{2}\left(1+(\lambda^{2})^{-2}+2(\lambda^{2})^{-1}+2(\lambda^{-2}\zeta^{-1})\right)
(7.6) =ζλ2R(λ2,ζ),\displaystyle=\zeta\lambda^{2}R(\lambda^{2},\zeta),

where R(z,θ)=1+z2+2z1+2z1θ1R(z,\theta)=1+z^{-2}+2z^{-1}+2z^{-1}\theta^{-1} is defined on ×(𝒪{0})\mathbb{C}^{*}\times(\mathcal{O}-\{0\}). Since |θ|1|\theta|\geq 1 for any θ𝒪{0}\theta\in\mathcal{O}-\{0\}, it follows that for any δ>0\delta>0 there exists N>0N>0 such that for (z,θ)×(𝒪{0})(z,\theta)\in\mathbb{C}^{*}\times(\mathcal{O}-\{0\}) with |z|>N|z|>N, it holds |R(z,θ)|>1δ.|R(z,\theta)|>1-\delta. In particular, we can choose N0>2N_{0}>2 such that |z|>N0|z|>N_{0} implies |R(z,θ)|>324.|R(z,\theta)|>\dfrac{3\sqrt{2}}{4}.

Case 2. ζ𝒪k\zeta\in\mathcal{O}^{*}_{k}: Since kk is a quadratic imaginary field, ζ\zeta is a nthn\mathrm{th} root of unity. In particular, ζ\zeta is an algebraic integer of degree ϕ(n)\phi(n), where ϕ\phi denotes the Euler’s totient function (see [Lan94, Chapter IV, Theorem 2]). Since ζ\zeta has degree 22, we conclude that n{1,2,3,4,6}n\in\{1,2,3,4,6\}, i.e. ζJ={±1,±i,±ω,±ω2}\zeta\in J=\{\pm 1,\pm i,\pm\omega,\pm\omega^{2}\}, where ω=12i32\omega=\frac{1}{2}-i\frac{\sqrt{3}}{2} is a primitive sixth root of unity.

By Proposition 7.1 we have that

4cosh((η)2)=|T(μ)2|+|T(μ)+2|.4\cosh\left(\frac{\ell(\eta)}{2}\right)=|T(\mu)-2|+|T(\mu)+2|.

However, |ζ|=1|\zeta|=1 and then Equation (7.4) implies that

|T(μ)2|+|T(μ)+2|=|T(λ2)+2|+|T(λ2)+2+4ζ1|.|T(\mu)-2|+|T(\mu)+2|=|T(\lambda^{2})+2|+|T(\lambda^{2})+2+4\zeta^{-1}|.

Hence, we can guarantee that (η)2(γ)\ell(\eta)\geq 2\ell(\gamma) whenever

(7.7) |T(λ2)+2+4ζ1||T(λ2)2||T(\lambda^{2})+2+4\zeta^{-1}|\geq|T(\lambda^{2})-2|

for any ζJ\zeta\in J. In order to compute the difference |T(λ2)+2+4ζ1||T(λ2)2||T(\lambda^{2})+2+4\zeta^{-1}|-|T(\lambda^{2})-2|, we consider the map defined on (π/2,π/2)(-\pi/2,\pi/2), by

hP,ζ(ϕ)=|Peiϕ+2+4ζ|2|Peiϕ2|2h_{P,\zeta}(\phi)=|Pe^{i\phi}+2+4\zeta|^{2}-|Pe^{i\phi}-2|^{2}

where P>1P>1 and ζ\zeta\in\mathbb{C} are fixed. In this way Inequality (7.7) is equivalent to

(7.8) h|T(λ2)|,ζ1(Arg(T(λ2)))0h_{|T(\lambda^{2})|,\zeta^{-1}}(\mathrm{Arg}(T(\lambda^{2})))\geq 0

for any ζJ\zeta\in J. We then look for conditions on Arg(λ)\mathrm{Arg}(\lambda) such that (7.8) holds for any ζJ\zeta\in J. It is clear that hP,1(ϕ)=0h_{P,-1}(\phi)=0 for any PP and ϕ\phi, so we can assume that ζJ{1}\zeta\in J\setminus\{-1\}.

It is straightforward to check that

hP,ζ(ϕ)=16(1+(ζ))+8Pcos(ϕ)[1+(ζ)+(ζ)tan(ϕ)]h_{P,\zeta}(\phi)=16(1+\Re(\zeta))+8P\cos(\phi)[1+\Re(\zeta)+\Im(\zeta)\tan(\phi)]

.

Hence, if ζ=1\zeta=1, since cos(ϕ)>0\cos(\phi)>0 we have hP,1(ϕ)>0h_{P,1}(\phi)>0 for all ϕ(π/2,π/2)\phi\in(-\pi/2,\pi/2). It follows from ζJ{±1}\zeta\in J\setminus\{\pm 1\} that

(7.9) 1+(ζ)121+\Re(\zeta)\geq\frac{1}{2}

and

(7.10) (ζ)1 , (ζ)0.\Im(\zeta)\geq-1\mbox{ , }\Im(\zeta)\neq 0.

Therefore hP,ζ>0h_{P,\zeta}>0 whenever

(7.11) 1+(ζ)+(ζ)tanϕ>0.1+\Re(\zeta)+\Im(\zeta)\tan\phi>0.

Suppose now that 0<tan(ϕ)<120<\tan(\phi)<\frac{1}{2}. If (ζ)0\Im(\zeta)\geq 0 then (7.11) follows from (7.9). On the other hand, if (ζ)<0\Im(\zeta)<0, by (7.10) we get 0<(ζ)10<-\Im(\zeta)\leq 1, and together with (7.9) we obtain that

tan(ϕ)<12<1+(ζ)(ζ),\tan(\phi)<\frac{1}{2}<\frac{1+\Re(\zeta)}{-\Im(\zeta)},

from which (7.11) follows. Then, if 0<Arg(λ)<12arctan(12)0<\mathrm{Arg}(\lambda)<\frac{1}{2}\arctan\left(\frac{1}{2}\right), then 0<tan(Arg(T(λ2)))<120<\tan(\mathrm{Arg}(T(\lambda^{2})))<\frac{1}{2} by (7.3) and the fact that Arg(λ2)=2Arg(λ)\mathrm{Arg}(\lambda^{2})=2\mathrm{Arg}(\lambda). Therefore (7.8) follows as desired.

Since (γ2)=2(γ)\ell(\gamma^{2})=2\ell(\gamma) and hol(γ2)=2hol(γ)=2Arg(λ)\mathrm{hol}(\gamma^{2})=2\mathrm{hol}(\gamma)=2\mathrm{Arg}(\lambda), we conclude from the analysis of the two cases for L=4log(N0)>0L=4\log(N_{0})>0 (with N0N_{0} given in Case 1), and ϵ=12arctan12\epsilon=\frac{1}{2}\arctan\frac{1}{2}, that if (γ)>L\ell(\gamma)>L and 0hol(γ)<ϵ0\leq\mathrm{hol}(\gamma)<\epsilon, then γ2\gamma^{2} minimizes the set of displacements of Γ(tr(γ))\Gamma(\mathrm{tr}(\gamma)), and therefore

sys(Γ(t)\3)=2(γ).\mathrm{sys}(\Gamma(t)\backslash\mathbb{H}^{3})=2\ell(\gamma).

7.4. Proof of Theorem D

Let γPSL(2,)\gamma\in\mathrm{PSL}(2,\mathbb{C}) be a loxodromic element. We can associate to γ\gamma the complex number z(γ)=e(γ)2eihol(γ)z(\gamma)=e^{\frac{\ell(\gamma)}{2}}e^{i\mathrm{hol}(\gamma)}. Thus, by construction in Section (7.1) we have that T(z(γ))T(z(\gamma)) is the trace of some lifting of γ\gamma in SL(2,)\mathrm{SL}(2,\mathbb{C}). Thereby, we will adopt as the trace of γ\gamma the complex number written as

T(γ):=T(z(γ)).T(\gamma):=T(z(\gamma)).

Note that this definition of trace remains invariant under conjugation and extend the definition of trace to the conjugacy class of any subgroup of PSL(2,)\mathrm{PSL}(2,\mathbb{C}). For a complex number zz we will define the norm of zz as the nonnegative real number |z|2|z|^{2}. If Γ<PSL(2,)\Gamma<\mathrm{PSL}(2,\mathbb{C}) is a Kleinian group, we define σ(N,I)\sigma(N,I) (resp. τ(N,I)\tau(N,I)) as the number of primitive conjugacy classes of Γ\Gamma with norm of trace at most NN and holonomy in I,I, counted with multiplicity (resp. counted without multipliticy). By definition, the mean multiplicity is given by

μ0(N,I)=σ(N,I)τ(N,I).\mu_{0}(N,I)=\dfrac{\sigma(N,I)}{\tau(N,I)}.

These definitions will be convenient for presenting the following proposition.

Proposition 7.4.

Let Γ\Gamma be an arithmetic Kleinian group derived from a quaternion algebra over an imaginary quadratic field kk. For any subinterval I[0,2π]I\subset[0,2\pi], let μ0(N,I)\mu_{0}(N,I) be the mean multiplicity of primitive conjugacy classes of Γ\Gamma with trace of norm at most NN and holonomy contained in II. Then there exists a constant c>0c>0 depending only on kk and II such that

μ0(N,I)cNlog(N) when N.\mu_{0}(N,I)\gtrsim c\frac{N}{\log(N)}\mbox{ when }N\to\infty.
Proof.

Let 𝒪k\mathcal{O}_{k} be the ring of integers of kk. For any conjugacy class [γ]Γ[\gamma]\subset\Gamma we have T(γ)𝒪kT(\gamma)\in\mathcal{O}_{k}. Moreover, if |z(γ)|>1|z(\gamma)|>1, then the norm of T(γ)T(\gamma) is at most |z(γ)|2+3|z(\gamma)|^{2}+3. For any L>0L>0 and I[0,2π]I\subset[0,2\pi], consider

𝒩(L,I)=#{[γ]Γγ is primitive, |z(γ)|L and hol(γ)I}.\mathcal{N}(L,I)=\#\{[\gamma]\subset\Gamma\mid\gamma\mbox{ is primitive, }|z(\gamma)|\leq L\mbox{ and }\mathrm{hol}(\gamma)\in I\}.

By [SW99, Corollary] (see also [MMO14, Thm. 1.3] for a more explicit statement), there exists a constant c1c_{1} which depends only on II such that

𝒩(L,I)c1L4log(L) when L.\mathcal{N}(L,I)\sim c_{1}\frac{L^{4}}{\log(L)}\mbox{ when }L\to\infty.

Hence, σ(N,I)\sigma(N,I) is at least 𝒩(N3,I)\mathcal{N}(\sqrt{N-3},I), implying that

(7.12) σ(N,I)c1N2log(N)\sigma(N,I)\gtrsim c_{1}^{\prime}\frac{N^{2}}{\log(N)}

for some constant c1c_{1} depending only on II and NN sufficiently large.

On the other hand, since 𝒪k\mathcal{O}_{k} is a lattice in \mathbb{C}, there exists a constant c2>0c_{2}>0 depending only on kk such that

#(𝒪kB(0,R))c2R2\#(\mathcal{O}_{k}\cap B(0,R))\sim c_{2}R^{2}

(see [Lan94, Ch. V, Thm. 2]). Hence, when NN is big enough we have

(7.13) τ(N,I)c2N\tau(N,I)\lesssim c_{2}N

Thus, if we combine the asymptotic bounds (7.12) and (7.13), by the definition of mean multiplicity there exists a constant c>0c>0 which depends only in II and kk such that

μ0(N,I)cNlog(N) when N.\mu_{0}(N,I)\gtrsim c\frac{N}{\log(N)}\text{ when }N\to\infty.

We are now ready to finish the proof of Theorem D. In fact, we will state a more precise result which implies that every commensurability class of arithmetic hyperbolic 33-manifolds with imaginary quadratic invariant trace field contains a sequence of manifolds with kissing number as stated. It is well known by the Classification Theorem of Quaternions Algebras over number fields (see [MR03, Theorem 7.3.6]) that there exist compact and non compact arithmetic hyperbolic 33-manifolds with this property.

Theorem 7.5.

Let Γ<PSL(2,)\Gamma<\mathrm{PSL}(2,\mathbb{C}) be an arithmetic Kleinian group with an imaginary quadratic invariant trace field. There exists a sequence {Γj}\{\Gamma_{j}\} of torsion-free subgroups of Γ\Gamma with arbitrarily large index such that the corresponding sequence of finite volume hyperbolic 3-manifolds Mj=Γj\3M_{j}=\Gamma_{j}\backslash\mathbb{H}^{3} satisfies

Kiss(Mj)cvol(Mj)43log(vol(Mj)\mathrm{Kiss}(M_{j})\geq c\frac{\mathrm{vol}(M_{j})^{\frac{4}{3}}}{\log(\mathrm{vol}(M_{j})}

for some constant cc which does not depend on jj.

Proof.

We can suppose that Γ\Gamma is derived from a quaternion algebra since Γ(2)=γ2γΓ\Gamma^{(2)}=\langle\gamma^{2}\mid\gamma\in\Gamma\rangle has finite index in Γ\Gamma and is derived from a quaternion algebra ([MR03, Corollary 8.3.5]).

Consider the constants LL and ε\varepsilon given in Proposition 7.3, and let kk be the invariant trace field of Γ\Gamma. If we set I=[0,ε]I=[0,\varepsilon], then by Proposition 7.4 there exists a sequence of traces tj𝒪kt_{j}\in\mathcal{O}_{k} with |tj|2=Nj|t_{j}|^{2}=N_{j}\to\infty such that the number njn_{j} of primitive conjugacy classes of Γ\Gamma with trace tjt_{j} satisfy

njcNjlog(Nj),n_{j}\geq c\dfrac{N_{j}}{\log(N_{j})},

where c=ck,ε>0c=c_{k,\varepsilon}>0.

By Lemma (7.2), if NjN_{j} is large enough, tr(γ)=±2\mathrm{tr}(\gamma)=\pm 2 or |tr(γ)|>2|\mathrm{tr}(\gamma)|>2. Thus we can assume that for jj sufficiently large Γ(tj)\Gamma(t_{j}) is torsion-free and Nj>LN_{j}>L for all jj. Now we can argue as in theorem B.

Let γ1,,γnjΓ\gamma_{1},\ldots,\gamma_{n_{j}}\in\Gamma be a set of representatives of all primitive conjugacy classes in Γ\Gamma of trace tjt_{j}, and let Gj=Γ/Γ(tj)G_{j}=\Gamma/\Gamma(t_{j}) be a subgroup of isometries of Mj=Γ(tj)\3M_{j}=\Gamma(t_{j})\backslash\mathbb{H}^{3}. By Proposition 7.3, each γi2Γ(tj)\gamma_{i}^{2}\in\Gamma(t_{j}), and their induced closed geodesics are systoles of MjM_{j}. Since γi2\gamma_{i}^{2} has order 22, by Lemma 6.1 we have that

Kiss(Mj)i=1nj#(Gjγi2)12nj#Gj.\mathrm{Kiss}(M_{j})\geq\sum_{i=1}^{n_{j}}\#(G_{j}\cdot\gamma_{i}^{2})\geq\frac{1}{2}n_{j}\#G_{j}.

It is a well-known fact that #Gj=[Γ:Γ(tj)]CNj3\#G_{j}=[\Gamma:\Gamma(t_{j})]\leq CN_{j}^{3} for some constant C>0C>0, which does not depend on jj (see [Mak13],[KSV07] or [Kuc15]). Moreover, vol(Mj)=μ[Γ:Γ(tj)]\mathrm{vol}(M_{j})=\mu[\Gamma:\Gamma(t_{j})] for all jj, where μ=vol(Γ\3)\mu=\mathrm{vol}(\Gamma\backslash\mathbb{H}^{3}) . Putting together the inequalities above, as in the proof of Theorem B, we obtain that

Kiss(Mj)cvol(Mj)43log(vol(Mj))\mathrm{Kiss}(M_{j})\geq c\frac{\mathrm{vol}(M_{j})^{\frac{4}{3}}}{\log(\mathrm{vol}(M_{j}))}

for any jj, where cc does not depend on jj.

References

  • [Ano83] Dmitry V Anosov, On generic properties of closed geodesics, Mathematics of the USSR-Izvestiya 21 (1983), no. 1, 1.
  • [BCP21] Thomas Budzinski, Nicolas Curien, and Bram Petri, On the minimal diameter of closed hyperbolic surfaces, Duke Mathematical Journal 170 (2021), no. 2, 365–377.
  • [BE12] Mikhail Belolipetsky and Vincent Emery, On volumes of arithmetic quotients of po(n,1)po(n,1), n odd, Proceedings of the London Mathematical Society 105 (2012), no. 3, 541–570.
  • [Bel14] Mikhail Belolipetsky, Hyperbolic orbifolds of small volume, Proceeding of the International Congress of Mathematicans, ICM 2014, 2014, pp. 837–851.
  • [BHC62] Armand Borel and Harish-Chandra, Arithmetic subgroups of algebraic groups, Annals of Mathematics 75 (1962), no. 3, 485–535.
  • [BP22] Maxime Fortier Bourque and Bram Petri, Kissing numbers of closed hyperbolic manifolds, Amer. J. Math. 144 (2022), no. 4, 1067–1085.
  • [Bus80] Peter Buser, On Cheeger’s inequality λ1>h2/4\lambda_{1}>h^{2}/4, Geometry of the Laplace operator (Proc. Sympos. Pure Math., Univ. Hawaii, Honolulu, Hawaii, 1979) (1980), 29–77.
  • [DM21] Cayo Dória and Plinio Murillo, Hyperbolic 3-manifolds with large kissing number, Proceedings of the American Mathematical Society 149 (2021), no. 11, 4595–4607.
  • [EGM87] Jürgen Elstrodt, Fritz Grunewald, and Jens Mennicke, Vahlen’s group of Clifford matrices and spin-groups, Mathematische Zeitschrift 196 (1987), no. 3, 369–390.
  • [EGM13] by same author, Groups acting on hyperbolic space: Harmonic analysis and number theory, Springer Science & Business Media, 2013.
  • [Eme09] Vincent Emery, Du volume des quotients arithmétiques de l’espace hyperbolique, Ph.D. thesis, University of Fribourg Fribourg, Switzerland, 2009.
  • [FP16] Federica Fanoni and Hugo Parlier, Systoles and kissing numbers of finite area hyperbolic surfaces, Algebraic & Geometric Topology 15 (2016), no. 6, 3409–3433.
  • [Gen15] Matthieu Gendulphe, Systole et rayon interne des variétés hyperboliques non compactes, Geometry and Topology 19 (2015), no. 4, 2039–2080.
  • [GH01] Eknath Ghate and Eriko Hironaka, The arithmetic and geometry of Salem numbers, Bulletin of the American Mathematical Society 38 (2001), no. 3, 293–314.
  • [Kee74] Linda Keen, Collars on Riemann surfaces, Discontinuous groups and Riemann surfaces (Proc. Conf., Univ. Maryland, College Park, Md., 1973), 1974, pp. 263–268.
  • [KSV07] Mikhail Katz, Mary Schaps, and Uzi Vishne, Logarithmic growth of systole of arithmetic Riemann surfaces along congruence subgroups, Journal of Differential Geometry 76 (2007), no. 3, 399–422.
  • [Kuc15] Robert Kucharczyk, Modular embeddings and rigidity for Fuchsian groups, Acta Arithmetica 169 (2015), no. 1, 77–100.
  • [Lan94] Serge Lang, Algebraic number theory, vol. 110, Springer-Verlag New York, 1994.
  • [Mak13] Shotaro Makisumi, A note on Riemann surfaces of large systole, Journal of the Ramanujan Mathematical Society 28 (2013), no. 3, 359–377.
  • [MMO14] Gregory Margulis, Amir Mohammadi, and Hee Oh, Closed geodesics and holonomies for Kleinian manifolds, Geometric and Functional Analysis 24 (2014), no. 5, 1608–1636.
  • [MR03] Colin Maclachlan and Alan W. Reid, The arithmetic of hyperbolic 3-manifolds, vol. 219, Springer, 2003.
  • [Mur17] Plinio G P Murillo, On arithmetic manifolds with large systole, Ph.D. thesis, IMPA, 2017.
  • [Mur19] by same author, Systole of congruence coverings of arithmetic hyperbolic manifolds, Groups, Geometry, and Dynamics 13 (2019), no. 3, 1083–1102.
  • [Par13] Hugo Parlier, Kissing numbers for surfaces, Journal of Topology 6 (2013), no. 3, 777–791.
  • [Pet20] Bram Petri, Extremal problems and probabilistic methods in hyperbolic geometry, Available at https://webusers.imj-prg.fr/~bram.petri/notes_eppmhg_200825.pdf.
  • [Ran74] Burton Randol, Small eigenvalues of the Laplace operator on compact Riemann surfaces, Bull. Amer. Math. Soc. 80 (1974), 996–1000. MR 400316
  • [Sch95] Paul Schmutz, Extremal Riemann surfaces with a large number of systoles., extremal Riemann surfaces (san francisco, ca, 1995), 9–19, Contemp. Math 201 (1995).
  • [Sch96a] by same author, Arithmetic Fuchsian groups and the number of systoles, Mathematische Zeitschrift 223 (1996), no. 1, 13–25.
  • [Sch96b] by same author, Compact Riemann surfaces with many systoles, Duke Mathematical Journal 84 (1996), no. 1, 191–198.
  • [SW99] Peter Sarnak and Masato Wakayama, Equidistribution of holonomy about closed geodesics, Duke mathematical journal 100 (1999), no. 1, 1–58.
  • [Wat93] Peter Waterman, Möbius transformations in several dimensions, Advances in Mathematics 101 (1993), no. 1, 87–113.
  • [Xue92] X Xue, On the Betti numbers of a hyperbolic manifold, Geometric & Functional Analysis GAFA 2 (1992), no. 1, 126–136.