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

Fourier decay, Green’s functions and Schottky groups

Ilyas Bayramov
(August 2023)
Abstract

In this note, I would like to discuss an approach to the construction of Green’s function of laplacians on algebraic surfaces, indicated by Manin in [9], towards the computation of the Green’s function on surfaces using Schottky uniformization. We shall see that the exact geometric interpretation of the formula mentioned there is obscure, and try to remedy the situation by investigating convergence of deformations of that formula.

1 Acknowledgments

I thank Peter Kosenko and Dennis Sullivan for many useful suggestions.

2 Motivation for Manin’s formula

There is a certain generalization due to Arakelov of the usual intersection pairing between the divisors that is useful, for example, in the proof of Mordell’s theorem. Here by divisor we mean a Weil divisor, i.e., a complex linear combination of points on the Riemann surface SS.

To each divisor aa of degree 0 on a (smooth) algebraic curve SS, by Riemann-Roch [5], corresponds a meromorphic differential form ωa\omega_{a} such that

a=xSresx(ωa)x.a=\operatornamewithlimits{\sum}\limits_{x\in S}res_{x}(\omega_{a})\cdot x.

This differential form is unique up to an addition of a holomorphic differential form to it. Since the real parts of periods of a holomorphic differential determine it uniquely (by, say, Riemann period relations) and we have the freedom to choose any holomorphic differential, we can insist that ωa\omega_{a} has all of its periods purely imaginary.

Then dga:=ωaω¯a\text{d}g_{a}:=\omega_{a}-\bar{\omega}_{a} is exact since all of its periods are 0.0. gag_{a} is then called the Green-Arakelov function.

Definition[5]: Let b=ymy(y),b=\operatornamewithlimits{\sum}\limits_{y}m_{y}(y), a divisor of degree 0; let aa be another such divisor, |a||b|=|a|\cap|b|=\emptyset, (y)S,(y)\in S, SS an algebraic curve. Then the height pairing between aa and bb is g(a,b)=ymyga(y).g(a,b)=\operatornamewithlimits{\sum}\limits_{y}m_{y}g_{a}(y).

The goal is to compute ga.g_{a}. To do that, we need to take a detour into low-dimensional topology.

3 Schottky uniformization: existence

Schottky uniformization of a Riemann surface SS is a representation ρ:π1(S)PSL(2,),\rho:\pi_{1}(S)\to PSL(2,\mathbb{C}), satiasfying the following conditions. Its image, called a Schottky group, is a free group Γ\Gamma that is purely hyperbolic, i.e., the absolute values of traces of its elements are strictly bounded by 22 from below. PSL(2,)PSL(2,\mathbb{C}) acts on 3\mathbb{H}^{3} by isometries, and on S2S^{2} by linear fractional transformations. S2S^{2} can be identified with the boundary of 3\mathbb{H}^{3} at infinity (aka ”the conformal infinity”).

Now, any infinite discrete group acting on a compact space, in this case 3S2,\mathbb{H}^{3}\cup S^{2}, must have accumulation points. The set of such accumulation points must then be contained in S2S^{2} (because PSL(2,)PSL(2,\mathbb{C}) acts on 3\mathbb{H}^{3} by isometries). For a given Γ\Gamma we denote it by Λ(Γ)\Lambda(\Gamma) and call it the limit set. Its complement in §2\S^{2} is denoted by Ω(Γ)\Omega(\Gamma) and is called the domain of discontinuity. By Sullivan’s dictionary, these correspond to Julia and Fatou sets, respectively.

”Uniformization” then comes from the application of Ahlfors’ finiteness theorem.

It states that for any finitely generated discrete subgroup of PSL(2,)PSL(2,\mathbb{C}),

Ω(Γ)/Γ\Omega(\Gamma)/\Gamma

is a Riemann surface of finite area. Hence, in this particular case it is, as well, and to each Riemann surface we can associate a corresponding Schottky uniformization by Koebe’s retrosection theorem.

4 Schottky uniformization: uniqueness

This uniformization is not unique; to see where exactly the non-uniqueness arises from, we must consider the action of a Schottky group Γ\Gamma on all of 3Ω(Γ).\mathbb{H}^{3}\cup\Omega(\Gamma). Then the quotient K(Γ)K(\Gamma) of such an action admits a hyperbolic metric in its interior, and it is actually evident that K(Γ)K(\Gamma) is a handlebody. Then S:=Ω(Γ)/ΓS:=\Omega(\Gamma)/\Gamma determines a point in the Teichmüller space up to the action of the elements of the mapping class group that have representatives extending to homeomorphisms of K(Γ)K(\Gamma) that are isotopic to the identity. This subgroup of Mod(S)\text{Mod}(S) is denoted by Mod0(S)\text{Mod}_{0}(S) and is known as the handlebody group of a fixed genus g.g.

5 Compactification 𝔗¯\bar{\mathfrak{T}} of the Teichmuller space

Thurston defined a compactification of the Teichmuller space as follows: consider the set 𝒮\mathcal{S} of isotopy classes of simple closed curves on a surface. Then for two such classes a,b𝒮a,b\in\mathcal{S} the geometric intersection number i(a,b)i(a,b) is given by the smallest number of intersections between the representatives of aa and bb, respectively. Denote by +\mathbb{R}_{+} the set of nonegative reals with the usual induced topology. Then there exists a map

i:𝒮+𝒮\0,i_{*}:\mathcal{S}\to\mathbb{R}_{+}^{\mathcal{S}}\backslash 0,
i(a)(b)=i(a,b).i_{*}(a)(b)=i(a,b).

Now, there also exists a map π:+𝒮\0P(+𝒮).\pi:\mathbb{R}_{+}^{\mathcal{S}}\backslash 0\to P(\mathbb{R}_{+}{\mathcal{S}}). Here arises the question of topology on these spaces: we take +𝒮\mathbb{R}_{+}^{\mathcal{S}} with the weak (product) topology, and P(+𝒮)P(\mathbb{R}_{+}^{\mathcal{S}}) with the quotient topology. Then the closure 𝒮¯\bar{\mathcal{S}} of πi(𝒮)\pi\circ i_{*}(\mathcal{S}) in quotient topology is homeomorphic to a sphere S6g62b,S^{6g-6-2b}, where bb is the number of punctures.

On the other hand, there also exists an embedding of Teichmuller space into P(𝒮).P(\mathbb{R}^{\mathcal{S}}). It is given by the length function

l:𝔗×𝒮+𝒮,l:\mathfrak{T}\times\mathcal{S}\to\mathbb{R}_{+}^{\mathcal{S}},
l(θ,α)=infγαθ(γ),l(\theta,\alpha)=\inf\limits_{\gamma\in\alpha}\theta(\gamma),

where γ\gamma is a simple closed curve in the isotopy class of α.\alpha. Since this is well-defined, because both θ\theta and α\alpha are defined up to isotopy, this gives an embedding l:𝔗+𝒮\0l_{*}:\mathfrak{T}\to\mathbb{R}_{+}^{\mathcal{S}}\backslash 0 which is a homeomorphism onto the image.

Now it is obvious that the geometric intersection number of a𝒮a\in\mathcal{S} with itself is 0,0, and by [2], the geometric intersection number can be extended to 𝔗\mathfrak{T} and 𝒮¯\bar{\mathcal{S}} in such a way that the self-intersection number is respectively a non-zero constant on all of l(𝔗)l_{*}(\mathfrak{T}) and 0 on 𝒮¯.\bar{\mathcal{S}}. Since we consider the weak topology on +𝒮\0,\mathbb{R}_{+}^{\mathcal{S}}\backslash 0, we see that if a sequence θn\theta_{n} of elements of l(𝔗)l_{*}(\mathfrak{T}) leaves any compact set of +𝒮\0\mathbb{R}_{+}^{\mathcal{S}}\backslash 0 in such topology, then we must have that for at least one of the elements α\alpha of 𝒮\mathcal{S} l(θn,α).l(\theta_{n},\alpha)\to\infty. Then if a sequence π(θn)\pi(\theta_{n}) converges to some point bb in P(𝒮),P(\mathbb{R}^{\mathcal{S}}), a sequence of their preimages in +𝒮\mathbb{R}_{+}^{\mathcal{S}} must converges to a function aπ1(b);a\in\pi^{-1}(b); that means that there exists numbers λn>0\lambda_{n}>0 such that λnθn\lambda_{n}\theta_{n} converges to a;a; as per the above, λn0,\lambda_{n}\to 0, as at least one of the coordinates (θn(s))s𝒮(\theta_{n}(s))_{s\in\mathcal{S}} goes to .\infty. Then by the bilinearity (!) of the intersection number i(λnθn,λnθn)=|λn|2i(θn,θn)=C|λn|20=i(a,a).i(\lambda_{n}\theta_{n},\lambda_{n}\theta_{n})=|\lambda_{n}|^{2}i(\theta_{n},\theta_{n})=C|\lambda_{n}|^{2}\to 0=i(a,a). This implies that a𝒮¯,a\in\bar{\mathcal{S}}, by the converse of the above statement about 𝒮¯,\bar{\mathcal{S}}, which is less obvious. Notice that this argument is very similar to a measure concentration phenomenon. Thus, 𝔗¯=πl(𝔗)𝒮\bar{\mathfrak{T}}=\pi\circ l_{*}(\mathfrak{T})\cup{\mathcal{S}} is a compactification of 𝔗.\mathfrak{T}.

6 The action of Mod0(S)\text{Mod}_{0}(S) on 𝔗¯\bar{\mathfrak{T}}

Profoundly, the situation in this case is reminiscent of the Kleinian group action on 3:\mathbb{H}^{3}: for the subgroup Mod(H)\text{Mod}(H) of Mod(S)\text{Mod}(S) that consists of elements that have representatives extending to diffeomorphisms of the handlebody H,H=SH,\partial H=S, there is a domain of discontinuity on 𝔗¯,\bar{\mathfrak{T}}, and a limit set that is contained in 𝒮¯.\bar{\mathcal{S}}. Now, following Masur [10], lemma 1.1, we have that if a curve on SS bounds a disk in the corresponding handlebody, and α1,,αg\alpha_{1},\dots,\alpha_{g} are simple closed curves that bound disks in HH and cut SS into a 2g2g holed sphere, then β𝒮\beta\in\mathcal{S} bounds a disk iff i(β,αi)=0i(\beta,\alpha_{i})=0 or we have the following situation. Suppose i(β,αi)0i(\beta,\alpha_{i})\neq 0 for some ii. Choose (βl)l{1,,g}𝒮(\beta_{l})_{l\in\{1,\dots,g\}}\in\mathcal{S} with i(βl,αj)=δlji(\beta_{l},\alpha_{j})=\delta_{lj} and βl\beta_{l}-free generators of π1(H)=Γ.\pi_{1}(H)=\Gamma. As a word in βl\beta_{l} and βl1,β\beta_{l}^{-1},\beta is trivial. Up to a cyclic permutation, there must be either βjβj1\beta_{j}\beta_{j}^{-1} or βj1βj\beta_{j}^{-1}\beta_{j} somewhere in the word. But then β\beta crosses αj\alpha_{j} and then crosses it in the opposite direction.

Now, consider 𝒮c𝒮,\mathcal{S}_{c}\subset\mathcal{S}, the set of simple closed curves bounding disks in the handlebody. Then by Theorem 2.2 in Masur 𝒮c¯\bar{\mathcal{S}_{c}} is the limit of the actions of both Mod(H)\text{Mod}(H) and Mod0(S).\text{Mod}_{0}(S).

Finally, there is the canonical measure mm on 𝒮¯,\bar{\mathcal{S}}, see Masur’s theorem B,B, such that m(N:={a𝒮¯𝒮c¯|i(a,b)=0 for some b𝒮c¯})=0.m(N:=\{a\in\bar{\mathcal{S}}-\bar{\mathcal{S}_{c}}|i(a,b)=0\text{ for some }b\in\bar{\mathcal{S}_{c}}\})=0. Taking 𝒪:=(𝒮¯𝒮c¯)N,\mathcal{O}:=(\bar{\mathcal{S}}-\bar{\mathcal{S}_{c}})-N, we have that 𝒪\mathcal{O} is the domain of discontinuity of the action of both Mod(H)\text{Mod}(H) and Mod0(S)\text{Mod}_{0}(S) on 𝒮\mathcal{S} by theorem 2.1 in Masur. Moreover, in genus 2,2, according to Masur, the limit set 𝒮c¯\bar{\mathcal{S}_{c}} has measure 0.0.

7 Deforming the Schottky group

We consider here AH(H)AH(H) to be the set of all discrete faithful representations of Γ\Gamma into PSL(2,)PSL(2,\mathbb{C}) up to conjugation. Now, according to Li-Ohshika-Lecuire [12], we have that if ρnAH(H)\rho_{n}\in AH(H) corresponds to a sequence of elements of 𝔗\mathfrak{T} that converges to an element of Masur domain, then ρn\rho_{n} converges to an element of AH(H).AH(H). Finally, we have, according to Anderson-Canary-Culler-Shalen [1], that the set of Kleinian groups with the limit set the whole sphere is dense in the boundary of the domain of convex-cocompact representations, i.e., in this case, the quasi-conformal deformations of a Schottky group. Thus, if a sequence of elements ana_{n} of 𝔗\mathfrak{T} leaves any compact subset of 𝔗\mathfrak{T} in the quotient of weak topology described above, then its limit a𝔗a\in\mathfrak{T} almost always corresponds to a group Γa\Gamma_{a} with limit set Λ(Γa)=S2.\Lambda(\Gamma_{a})=S^{2}. This is why we can almost quasiconformally deform a Schottky group to another one that has S2S^{2} as its limit set.

8 Manin’s formula

In what follows, a,b,c,d\langle a,b,c,d\rangle denotes the cross-ratio of a,b,c,d.a,b,c,d.

Using Schottky uniformization, Manin [9] defined ωa\omega_{a} for any divisor of degree 0, as in section 11; for a=(x)(y)a=(x)-(y) it is

ν(x)(y)lX(x,y)ωl,\nu_{(x)-(y)}-\operatornamewithlimits{\sum}\limits_{l}X(x,y)\omega_{l},

where

ν(x)(y)=dlogW(x)(y),z0(z)\nu_{(x)-(y)}=\text{dlog}W_{(x)-(y),z_{0}}(z)

,

Xl(a,b)=1hS(gk)log|a,b,z+(h),z(h)|hC(gk|gl)log|z+(gk)z(gk),hz+(gl),hz(gl)|,X_{l}(a,b)=-\sqrt{-1}\frac{\operatornamewithlimits{\sum}\limits_{h\in S(g_{k})}\text{log}|\langle a,b,z^{+}(h),z^{-}(h)\rangle|}{\operatornamewithlimits{\sum}\limits_{h\in C(g_{k}|g_{l})}\text{log}|\langle z^{+}(g_{k})z^{-}(g_{k}),hz^{+}(g_{l}),hz^{-}(g_{l})\rangle|},
ωl=112πdlogW(gkz1)(z1),z0(z).\omega_{l}=-\sqrt{-1}\frac{1}{2\pi}\text{dlog}W_{(g_{k}z_{1})-(z_{1}),z_{0}}(z).

Here,

Wd,q(r)=hΓwd(hr)wd(hq),W_{d,q}(r)=\operatornamewithlimits{\prod}\limits_{h\in\Gamma}\frac{w_{d}(hr)}{w_{d}(hq)},

for dd a divisor in P1,\mathbb{C}P^{1}, wdw_{d} is the meromorphic function with this divisor. Also, S(gk)S(g_{k}) is the conjugacy class of g,g, and C(gk|gl)C(g_{k}|g_{l})-double coset representatives.

9 Simplification of Manin’s formula

It is known as a consequence of Fay’s trisecant theorem [14] that

exp(a1a2ω(x)(y))=θ(α+a1xω)θ(α+a2yω)θ(α+a2xω)θ(α+a1yω),\text{exp}(\operatornamewithlimits{\int}\limits_{a_{1}}^{a_{2}}\omega_{(x)-(y)})=\frac{\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{1}}^{x}\overrightarrow{\omega})\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{2}}^{y}\overrightarrow{\omega})}{\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{2}}^{x}\overrightarrow{\omega})\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{1}}^{y}\overrightarrow{\omega})},
Rea1a2ω(x)(y)=log(θ(α+a1xω))+log(θ(α+a2yω))Re\operatornamewithlimits{\int}\limits_{a_{1}}^{a_{2}}\omega_{(x)-(y)}=\log(\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{1}}^{x}\overrightarrow{\omega}))+\log(\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{2}}^{y}\overrightarrow{\omega}))
log(θ(α+a2xω))log(θ(α+a1yω));-\log(\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{2}}^{x}\overrightarrow{\omega}))-\log(\theta(\alpha+\operatornamewithlimits{\int}\limits_{a_{1}}^{y}\overrightarrow{\omega}));

then by Roelcke and Fay [7] we have

log(θ(α+xyω))=𝑑S(p)𝑑S(q)Rexpω(y)(q),\log(\theta(\alpha+\operatornamewithlimits{\int}\limits_{x}^{y}\overrightarrow{\omega}))=\int{dS(p)}\int{dS(q)}Re\operatornamewithlimits{\int}\limits_{x}^{p}\omega_{(y)-(q)},

where

Rexpω(y)(q)=hΓlog|y,q,hx,hp|;Re\operatornamewithlimits{\int}\limits_{x}^{p}\omega_{(y)-(q)}=\operatornamewithlimits{\sum}\limits_{h\in\Gamma}\log|\langle y,q,hx,hp\rangle|;

so we have as the Green’s function

Ω(Γ)Ω(Γ)Rexpω(y)(q)dS(p)dS(q)\operatornamewithlimits{\int}\limits_{\Omega(\Gamma)}\operatornamewithlimits{\int}\limits_{\Omega(\Gamma)}Re\operatornamewithlimits{\int}\limits_{x}^{p}\omega_{(y)-(q)}dS(p)dS(q)

where dSdS is the Γ\Gamma-invariant measure on Ω(Γ).\Omega(\Gamma).

10 Fixed-point theorems

We would like to rephrase the existence of the solution to Laplace’s equation on Λ(Γ)\Lambda(\Gamma) with potential (which is equivalent to being able to find Green’s function) through a fixed point problem. We follow Evans [4] in this. Consider the following setup: assume that the smooth functions {wk}k\{w_{k}\}_{k\in\mathbb{N}} form an orthonormal basis of W01,2W^{1,2}_{0}, where W01,2W^{1,2}_{0} is the closure of the space of compactly supported smooth functions in W1,2,W^{1,2}, the Sobolev space of square-integrable functions with square-integrable weak derivative. For example, we can take wkw_{k} to be harmonics so that the corresponding coefficients of wkw_{k} in a series converging to the function fW01,2f\in W^{1,2}_{0} are Foruier coefficients. Consider the system of equations

U𝐚(Dum)Dwkdx=Ufwkdx,\operatornamewithlimits{\int}\limits_{U}\mathbf{a}(Du_{m})\cdot Dw_{k}dx=\operatornamewithlimits{\int}\limits_{U}fw_{k}dx,

where um=k=1mdmkwk,u_{m}=\operatornamewithlimits{\sum}\limits_{k=1}^{m}d^{k}_{m}w_{k}, but for each umu_{m} we only consider those equations up to k=m.k=m. Here 𝐚:nn,𝐚={a1,,an}\mathbf{a}:\mathbb{R}^{n}\to\mathbb{R}^{n},\mathbf{a}=\{a^{1},\dots,a^{n}\} is a vector field. Also, we must impose the following on 𝐚:\mathbf{a}:

(𝐚(p)𝐚(q))(pq)0,(\mathbf{a}(p)-\mathbf{a}(q))(p-q)\geq 0,
𝐚C(1+|p|),\mathbf{a}\leq C(1+|p|),
𝐚(p)pα|p|2β\mathbf{a}(p)\cdot p\geq\alpha|p|^{2}-\beta

for some C,α>0C,\alpha>0 and β0.\beta\geq 0. Then we can use the fixed point theorem to show that, if we consider

vk(d):=U𝐚(j=1mdjDwj)Dwkfwkdx(k=1,,m),v^{k}(d):=\operatornamewithlimits{\int}\limits_{U}\mathbf{a}(\operatornamewithlimits{\sum}\limits_{j=1}^{m}d_{j}Dw_{j})\cdot Dw_{k}-fw_{k}dx(k=1,\dots,m),

over all d=(d1,,dm)m,d=(d_{1},\dots,d_{m})\in\mathbb{R}^{m}, then vk(d)v^{k}(d) has a zero.

This equation is equivalent to Laplace’s equation Δu+cu=0,-\Delta u+cu=0, where c=2aΔa(a)24a2c=\frac{2a\Delta a-(\nabla a)^{2}}{4a^{2}} (see [6], for example).

We can interpret the space of dd^{\prime}s in this case as a space of Fourier coefficients. Naud et al. [8] have established (non-uniform) boundedness of Fourier coefficients of functions on the limit set of Schottky groups. Thus, we have that the Green’s function exists on Λ(Γ)\Lambda(\Gamma) for any Γ\Gamma free convex-cocompact.

Now, by [11], we have that the limit sets of algebraically convergent sequence of Schottky groups, defined in section 6, converge to the limit set of the of the group Γa\Gamma_{a} in Gromov-Hausdorff sense. On the other hand, following Quint [15], we have that the Patterson-Sullivan measure is in the Lebesgue class of measures on S2,S^{2}, hence it is doubling with respect to the Fubini-Study metric on S2,S^{2}, which we identify with P1\mathbb{C}P^{1}. By the results of [3] adapted to the case of sets, rather than manifolds, we have that if we have a sequence of metric spaces equipped with measures that are doubling with respect to those metrics, then the Green’s functions on those sets converge to a Green’s function on the Gromov-Hausdorff limit of those sets.

Finally, considering both

hΓalog|y,q,hx,hp|\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a}}\log|\langle y,q,hx,hp\rangle|

and

hΓanlog|y,q,hx,hp|,\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a_{n}}}\log|\langle y,q,hx,hp\rangle|,

Γa\Gamma_{a} being the algebraic limit of Γan\Gamma_{a_{n}} as above, we have

hΓalog|y,q,hx,hp|=\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a}}\log|\langle y,q,hx,hp\rangle|=
hΓaχΩ(Γan)(y,q,hx,hp)log|y,q,hx,hp|+\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a}}\chi_{\Omega(\Gamma_{a_{n}})}(\langle y,q,hx,hp\rangle)\log|\langle y,q,hx,hp\rangle|+
+hΓaχΛ(Γan)(y,q,hx,hp)log|y,q,hx,hp|,+\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a}}\chi_{\Lambda(\Gamma_{a_{n}})}(\langle y,q,hx,hp\rangle)\log|\langle y,q,hx,hp\rangle|,

where χs\chi^{\prime}\text{s} are indicator functions. Since the action of Γa\Gamma_{a} is topologically dense on S2,S^{2}, that is, the closure of any orbit is dense, we have that, in particular, any orbit of Γan\Gamma_{a_{n}} that lies in Ω(Γan)\Omega(\Gamma_{a_{n}}) can be approximated by the orbit of the same point under Γa.\Gamma_{a}. Since the action of Γan\Gamma_{a_{n}} on Λ(Γan)\Lambda(\Gamma_{a_{n}}) is also topologically dense, we can estimate

hΓaχΛ(Γan)(y,q,hx,hp)log|y,q,hx,hp|\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a}}\chi_{\Lambda(\Gamma_{a_{n}})}(\langle y,q,hx,hp\rangle)\log|\langle y,q,hx,hp\rangle|

by

hΓanχΛ(Γan)(y,q,hx,hp)log|y,q,hx,hp|,\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a_{n}}}\chi_{\Lambda(\Gamma_{a_{n}})}(\langle y,q,hx,hp\rangle)\log|\langle y,q,hx,hp\rangle|,

which converges by [8].

Thus, we have the following:

Theorem: For almost any Schottky representation Γ0\Gamma_{0} of a fixed free group, for almost any point at the boundary of the Schottky space, there exists an unbounded sequence of representations Γn\Gamma_{n} converging to it and quasiconformal maps on S2,S^{2}_{\infty}, conjugating the action of Γ0\Gamma_{0} to Γn,\Gamma_{n}, such that for any nn, the expression hΓanlog|y,q,hx,hp|\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a_{n}}}\log|\langle y,q,hx,hp\rangle|, where Γ\Gamma is the limit of Γn\Gamma_{n} in AH(S),AH(S), for any quadruple of points y,q,x,py,q,x,p, converges.

Proof: If y,q,x,pΛ(Γan),\langle y,q,x,p\rangle\in\Lambda(\Gamma_{a_{n}}), this follows by [8] and invariance of Λ(Γan)\Lambda(\Gamma_{a_{n}}). If y,q,x,pΩ(Γan),\rangle y,q,x,p\langle\in\Omega(\Gamma_{a_{n}}), then we can estimate it from above by

hΓaχΩ(Γan)(y,q,hx,hp)log|y,q,hx,hp|,\operatornamewithlimits{\sum}\limits_{h\in\Gamma_{a}}\chi_{\Omega(\Gamma_{a_{n}})}(\langle y,q,hx,hp\rangle)\log|\langle y,q,hx,hp\rangle|,

which converges by the above. This concludes the proof.

Remark: The result above generalizes [13].

References

  • [1] J. W. Anderson, R. D. Canary, M. Culler, and P. B. Shalen. Free kleinian groups and volumes of hyperbolic 3-manifolds. Journal of differential geometry, 1996.
  • [2] F. Bonahon. Bouts des varietes hyperboliques de dimension 3. Annals of Mathematics, 124(1):71–158, 1986.
  • [3] Y. Ding. Heat kernels and green’s functions on limit space. Communications in analysis and geometry, 2002.
  • [4] L. C. Evans. Partial Differential Equations: Second Edition. American Mathematical Society, 2010.
  • [5] B. H. Gross. Local heights on curves. In G. Cornell and J. Silverman, editors, Arithmetic Geometry, pages 327––339. Springer, New York, NY, 1986.
  • [6] https://math.stackexchange.com/users/111793/cookie. Divergence structure equation. https://math.stackexchange.com/questions/1119593/divergence-structure-equation.
  • [7] A. Kokotov and K. Lagota. Green function and self-adjoint laplacians on polyhedral surfaces. Canadian Journal of Mathematics, 72(5):1324 – 1351, 2020.
  • [8] J. Li, F. Naud, and W. Pan. Kleinian schottky groups, patterson–sullivan measures, and fourier decay. Duke Mathematical Journal, 170(4):775 – 825, 2021.
  • [9] Y. I. Manin. Three-dimensional hyperbolic geometry as infinity-adic arakelov geometry. Inventiones mathematicae, 104:223–243, 1991.
  • [10] H. Masur. Measured foliations and handlebodies. Ergodic Theory and Dynamical Systems, 6:99–116, 1986.
  • [11] M. Mj and C. Series. Limits of limit sets i. Geometriae dedicata, 2013.
  • [12] K. Ohshika. Hyperbolic 3-manifolds and kleinian groups at the crossroads. preprint, 2004.
  • [13] M. Pollicott. The schottky–klein prime function and counting functions for fenchel double crosses. Monatshefte für Mathematik, 2021.
  • [14] C. Poor. Fay’s trisecant formula and cross-ratios. Proceedings of the American Mathematical Society, 114:667–671, 1992.
  • [15] J. Quint. An overview of patterson-sullivan theory. https://www.math.u-bordeaux.fr/ jquint/publications/courszurich.pdf.