517
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Confirming Brennan’s Conjecture Numerically on a Counterexample to Thurston’s K = 2 Conjecture

Abstract

It was shown by Bishop that if Thurston’s K = 2 conjecture holds for some planar domain, then Brennan’s conjecture holds for the Riemann map of that domain as well. In this paper we show numerically that the original counterexample to Thurston’s K = 2 conjecture given by Epstein, Marden and Marković is not a counterexample to Brennan’s conjecture.

1 Introduction

1.1 Thurston’s K = 2 conjecture

Let Ω be a simply connected proper subdomain of CS2=H3 where H3 is the 3-dimensional hyperbolic space. Let Ω˜ be the union of all hyperbolic half-spaces H such that HH3Ω. Then we define the dome of Ω to be Dome(Ω)=Ω˜H3. It is known (see the book by Epstein and Marden [Citation7] for a detailed account) that Dome(Ω) with the path metric induced from H3 is isometric to the unit disk D with its hyperbolic metric. Using this isometry, we give Dome(Ω) a conformal structure.

Note that Ω and Dome(Ω) share the common boundary Ω, so we setF={f:ΩDome(Ω):f(x)=x for all xΩ and f is continuous}.

Let Möb(Ω) be the set of all Möbius transformations that preserve Ω. Each such transformation extends to an isometry of H3 preserving Dome(Ω), so we defineFeq={fF:f°γ=γ°f for all γMöb(Ω)}.

Definition 1.

Given a simply connected proper subdomain Ω of the plane, we define K(Ω)=inf{K>0:there exists a Kquasiconformal map fF}, and Keq(Ω)=inf{K>0:there exists a Kquasiconformal map fFeq}.

It was shown by Sullivan that supΩK(Ω)< in [Citation13]. A more detailed proof of this result and a proof that supΩKeq(Ω)< was given by Epstein and Marden in [Citation7]. Thurston conjectured the following result.

Conjecture (Thurston’s K = 2 conjecture). For any simply connected proper subdomain ΩC, we have supΩK(Ω)=supΩKeq(Ω)=2.

Epstein, Marden and Marković constructed counterexamples for this conjecture in [Citation8] for the equivariant case and in [Citation10] for the general case. Komori and Matthews then constructed in [Citation12] a more explicit counterexample for the equivariant case building on the ideas of [Citation8].

1.2 Brennan’s conjecture

We remind the reader that given a simply connected proper subdomain ΩC, by the Riemann mapping theorem there exists a conformal map F:ΩD. Brennan conjectured the following regarding the growth of F near Ω.

Conjecture (Brennan’s conjecture). Let Ω be a simply connected proper subdomain of C, and let F:ΩD be a conformal map. Then FLp(Ω) for all 43<p<4.

Brennan’s conjecture in full generality is still open.

1.3 Bishop’s theorem

Bishop showed in [Citation2] that if Thurston’s K = 2 conjecture holds for some domain, then Brennan’s conjecture holds for that domain as well.

Theorem 1.

Let Ω be a simply connected proper subdomain of the complex plane, and let F:ΩD be a conformal map. Then FLp(Ω) for all p<2K(Ω)K(Ω)1. In particular if K(Ω)=2, then Brennan’s conjecture holds for Ω.

It was shown by Epstein, Marden and Marković in [Citation9] that Thurston’s K = 2 conjecture holds for domains in C that are convex (in the Euclidean sense), by Theorem 1 therefore proving Brennan’s conjecture in this case. Bishop showed in [Citation3] that K(Ω)7.82 for all simply connected planar domains Ω, therefore giving a new proof that FL2.29(Ω) for conformal maps F:ΩD.

1.4 Our results

By Theorem 1, any counterexample to Brennan’s conjecture would also be a counterexample to Thurston’s K = 2 conjecture. In this paper we investigate if Brennan’s conjecture holds for a particular counterexample to Thurston’s K = 2 conjecture.

In this section we denote the counterexample to Thurston’s K = 2 conjecture from [Citation12] by Ω. In [Citation12] it was shown that K(Ω)=Keq(Ω)>2. We show numerically that Brennan’s conjecture holds for Ω.

Main result. Let F:ΩD be a conformal map and set p=sup{p>0:FLp(Ω)}. We show numerically that 5.52<p<5.54. In particular p>4 so Brennan’s conjecture holds for Ω.

The domain Ω is a connected component of the domain of discontinuity of an explicit Kleinian once-punctured torus group. Our results strongly suggest that Brennan’s conjecture holds for all domains constructed in this way. Equivalently, our results suggest that Brennan’s conjecture holds for all quasidisks invariant under a group of Möbius transformations, such that the quotient is a once-punctured torus.

1.5 General strategy

The domain Ω has an action by a Kleinian once-punctured torus group Γ. Write F:ΩD for its Riemann mapping and f=F1 for its inverse. Let Γ0=F°Γ°F1 be the once-punctured torus Fuchsian group obtained by conjugating Γ by F, and let ρ:Γ0Γ be the induced isomorphism. Let Φ be a fundamental domain for the action of Γ0 on D. Then f(Φ) is a fundamental domain for the action of Γ on Ω.

Since Ω and Γ are known, we can numerically compute the Riemann map F:ΩD using Schwarz-Christoffel mappings and a polygonal approximation to Ω. The quality of this estimate deteriorates near the boundary Ω, making it imprecise to check if FLp(Ω) directly. However for γΓ, the behavior of F on γf(Φ) is controlled by γ and ρ1(γ). Therefore FLp(Ω) if and only if a certain series over Γ0 depending on p,Γ0 and ρ converges.

We note that the estimates of F have higher accuracy away from Ω. We can hence still use them to reliably estimate ρ and Γ0. Then we use these estimates to check numerically if the series mentioned in the previous paragraph converges.

1.6 A more detailed outline of the argument and the computation

The paper is divided into a theoretical Section 2, a brief section where we describe Ω and Γ in more detail Section 3, and a numerical methods and results Section 4.

In the theory part, we show that Ω|F|pdxdy=D|f|2pdxdy is equal to a certain infinite series depending on Γ0 and ρ, up to a bounded multiplicative error. We do this by decomposing D into γΦ, where Φ is a fundamental domain for the action of Γ0 on D. We express the integral γΦ|f|2pdxdy in terms of γ and ρ(γ), up to a bounded multiplicative error. We work in greater generality, considering Riemann maps f:DΩ that conjugate a Fuchsian group Γ to a Kleinian group. We do this in Section 2, where the main result is Theorem 2.

In the proof of Theorem 2, issues arise since Φ is not assumed to be compact. This is handled by showing that on a horoball H, the derivative |f| achieves its minimum at the closest point of H to the origin, up to a bounded multiplicative error. This is Lemma 1 and is the most involved part of the proof of Theorem 2.

We now briefly describe how Ω is constructed in [Citation12]. They start with a hyperplane in H3 along with a Fuchsian once-punctured torus group that preserves it. Fix a hyperbolic element in this group, and consider the orbit of its axis. This is a discrete set of geodesics, along which they bend the hyperplane. The resulting pleated plane intersects the boundary of H3 in a curve that bounds Ω. We give an equivalent form of their construction in Section 3 that does the bending in the complex plane, without mentioning H3.

This construction makes it easy to identify a point on the boundary of Ω, and to see that the orbit of any point on Ω is dense in Ω. We use this observation to construct finite polygons that approximate Ω. These approximations are described in Section 4.1.

We use Schwarz-Christoffel mappings to numerically compute an approximation to the Riemann map F:ΩD. This approximation is used to (numerically) compute the generators of Γ0. This is explained in Section 4.2.

We use estimates of Γ0 and ρ to compute initial terms of the sum from Theorem 2. The terms appear to decay exponentially for p < 5.52 and to increase exponentially for p > 5.54. From this we conclude the Main result. This is done in Section 4.4.

2 Theoretical results

Throughout this section, we let Γ be a Fuchsian group such that D/Γ is a finite area hyperbolic surface. Let f:DΩ be a conformal map, normalized so that f(0)=0 and f(0)=1. Suppose that f conjugates Γ to a Kleinian group, and let ρ:ΓPSL(2,C) be the induced homomorphism. Our main theoretical result is the following estimate.

Theorem 2.

Given q > 0, there exists a constant C=C(Γ,q) that depends only on Γ and q such that 1CγΓ|γ(0)|q+2|ρ(γ)(0)|qD1|f(z)|qdxdyCγΓ|γ(0)|q+2|ρ(γ)(0)|q.

We now outline the proof of Theorem 2. The idea is to estimate the integral separately over H and DH, where H is the union of a certain Γ-invariant collection of horoballs based at fixed points of the parabolics in Γ. The proof consists of three steps.

  1. For any horoball H in H, we let z0 be the closest point of H to the origin. We show the inequality |f(z)|C|f(z0)| for zH, for some universal constant C. We will later use this to show that the integral from Theorem 2 over H is dominated by the integral over DH.

    We show |f(z)|C|f(z0)| using the fact that f conjugates a parabolic element preserving H to a parabolic Möbius transformation. The core of the proof of this bound are estimates on the growth and derivative of a univalent map g:HC satisfying g(z+1)=g(z)+1.

    We show the necessary bounds on g as Proposition 1 in Section 2.1. Then we derive |f(z)|C|f(z0)| in Lemma 1 in Section 2.2.

    In Corollary 1, Section 2.2, using |f(z)|C|f(z0)| we bound the integral of H|f(z)|qdxdy in terms of |f(z0)| and 1|z0|.

  2. In Section 2.3 we define H and construct a compact fundamental domain Φ* for the action of Γ on DH. We show an estimate on γΦ*|f(z)|qdxdy as Proposition 3 in Section 2.4, making essential use of the compactness of Φ*. This estimate is in terms of the Euclidean distance of γΦ* to the boundary D of the unit disk D, and the absolute value of the derivative |f(γ(0))|.

  3. By equivariance of f, the derivative f(γ(z)) is related to ρ(γ)(f(z)) and γ(z). This allows us to replace |f| from the estimates in the first two steps with the derivatives |ρ(γ)| and |γ|. For hyperbolic isometries γ of the disk, we relate |γ(0)| and |γ(0)|, so we also replace |γ(0)| with |γ(0)| in the estimates in the first two steps. This is done in Section 2.5, and concludes the proof of Theorem 2.

Notation and conventions

We write XZ,W,Y if there exists a constant C=C(Z,W,)>0 that depends only on the variables in the subscript, so that XCY. We analogously write XZ,W,Y if YZ,W,X and XZ,W,Y if XZ,W,Y and YZ,W,X.

We will use hyperbolic geometry in several places in this paper. We always use the metric of constant curvature –1. In particular on the unit disk D we have the metric 4|dz|2(1|z|2)2, and on the upper half-plane H we have the metric |dz|2Im(z)2. Whenever we write dist(·,·), we are referring to the distance coming from the hyperbolic metric on either D or H.

2.1 Univalent maps and parabolic isometries

The main result of this subsection is concerned with the growth of univalent maps g defined on the upper half-plane H that commute with the parabolic γ given by γ(z)=z+1. By taking quotients H/γ and C/γ, g descends to a univalent map h:D{0}C. The estimates we show on g come from the inequalities on h and its derivatives near 0, and follow from the general theory of univalent maps.

In Section 2.1.1 we recall some general results from the theory of univalent maps. The main result there is Claim 1. Then in Section 2.1.2 we state and prove the main result of this subsection.

2.1.1 General results on univalent maps

We recall some general theorems about univalent maps, that we later use in Section 2.1.2. We then show an estimate on how closely a univalent map h follows its linear approximation h(0)+h(0)z near 0. This is Claim 1.

Theorem (Koebe quarter theorem). Let h:DC be a univalent function with h(0)=0 and h(0)=1. Then h(D) contains the disk of radius 14 around 0.

Theorem (Koebe distortion theorem). Let h:DC be a univalent function with h(0)=0 and h(0)=1. Then |z|(1+|z|)2|h(z)||z|(1|z|)2,1|z|(1+|z|)3|h(z)|1+|z|(1|z|)3,1|z|1+|z||zh(z)h(z)|1+|z|1|z|.

Moreover, the second inequality implies that for any univalent function f:DC, we have for z,wD, |f(z)|(1|z|2)|f(w)|(1|w|2)e2dist(z,w).

The following Claim is the estimate we use to derive Proposition 1, the main result of this subsection.

Claim 1. Let h:DC be a univalent map. Then whenever |z|<12, we have |h(z)h(0)h(0)z|<100|h(0)||z|2.

Remark 1.

The optimal constant in Claim 1 can easily be computed from de Branges’ theorem [4]. However since this is a fairly simple result, we choose to include a more elementary proof (with a non-optimal constant) below.

Proof

of Claim 1. Without loss of generality we can suppose h(0)=0 and h(0)=1. Writeh(z)=z+n=2anzn.

Let γ be the circle of radius 23 centered at the origin. Then|an|=|12πiγh(z)zn+1dz|12π(32)n02π|h(23eiθ)|dθ.

By the Koebe distortion theorem, we have|h(23eiθ)|6.

Therefore |an|6(32)n. Therefore for |z|<12, we have|h(z)z|6n=2(32|z|)n<6·94|z|2n=0(34)n<100|z|2.

2.1.2 Growth of univalent maps that commute with a parabolic Möbius transformation

We now state and prove the main result of this subsection, Proposition 1. As explained at the start of this subsection, starting with a map g:HC that commutes with γ(z)=z+1, we construct a map h:D{0}H/γC/γC{0}. We show that it extends to 0 with h(0)=0 and then apply Claim 1 to h. This shows how h behaves near 0, and hence how g behaves near infinity. We also obtain some information on g from the Koebe distortion theorem and the Koebe quarter theorem applied to h.

Proposition 1.

There exists a universal constant C > 0 such that the following holds. Let g:HC be a univalent holomorphic map with g(z+1)=g(z)+1. Then there exists a complex number αC (depending on g) so that the following holds.

  1. Whenever Im(z)C, we have |g(z)z+α|Ce2πIm(z).

  2. Whenever Im(z)C, we have 1C|g(z)|C.

  3. When Im(z)>log2πIm(α), we have zg(H).

Proof.

Note that exp(2πig(z)) is 1-periodic, so there exists a holomorphic map h:D{0}C defined by h(exp(2πiz))=exp(2πig(z)). Since g is univalent, so is h. We first show that h extends to 0.

Claim 2. The map h has a removable singularity at 0 and can be extended holomorphically so that h(0)=0.

Proof.

Since h is univalent, it does not have an essential singularity at 0. We extend h to 0 so that it is a meromorphic function. Let γ(t)=exp(2πi(t+iR)) for t[0,1], and R > 0 large. Then h(exp(2πiz))h(exp(2πiz))exp(2πiz)=g(z), and hence12πiγh(z)h(z)dz=01g(t+iR)dt=1.

This shows that h has a simple zero at 0. Hence h(0)=0. □

We can now apply Claim 1 to get|h(z)h(0)z|<100|h(0)||z|2,whenever |z|<12. We now set C>log22π, so that when Im(z)C, we have |exp(2πiz)|<12, and hence|1h(0)exp(2πig(z))exp(2πiz)|<100e4πIm(z).

Therefore|exp(2πi(g(z)z+α))1|<100e2πIm(z),where αC is such that exp(2πiα)=h(0). For C large enough, whenever Im(z)C, we will have |g(z)z+αm|110 for some integer mZ. We replace α with αm, so that |g(z)z+α|110. Since exp is bilipschitz on the disk centered at 0 of radius 110, we see that|g(z)z+α|e2πIm(z), and (1) follows.

To show (2), we note that for w=exp(2πiz),g(z)=wh(w)h(w).

By the Koebe distortion theorem, since h is univalent on the unit disk, we have1|w|1+|w||wh(w)h(w)|1+|w|1|w|,so for any C > 0 fixed, |w|=e2πIm(z)e2πC, and we have |g(z)|=|wh(w)h(w)|1.

We now turn to (3). Note that Im(α)=12πRe(2πiα)=12πlog|h(0)|. By the Koebe quarter theorem, the set h(D) contains the disk D centered at 0 with radius 14|h(0)|=14e2πIm(α). Note that exp(2πiz)D if and only if Im(z)Im(α)+12πlog4=log2πIm(α), so the final claim of the Proposition follows. □

2.2 Derivative bounds on a horoball

Here we give a lower bound on the derivative |f| of a univalent map f:DC on a horoball, assuming that f conjugates a parabolic isometry of D that preserves that horoball to a parabolic Möbius transformation. The main result is Lemma 1. This lemma is used to bound the integral |f|qdxdy over a horoball in Corollary 1. This corollary is the result we use in the coming sections.

We first define the horoballs we will consider. All horoballs will be open. Given a horoball H, the horocycle H admits a natural orientation as follows. The vector v along H at zH is positive if (v, w) is a positively-oriented basis of TzD, where w is the vector at z tangent to the geodesic ray from z to the point at infinity of H.

Definition 2.

A horoball HD is l-adapted to a parabolic isometry γAut(D) if the following conditions hold,

  1. the parabolic γ preserves H,

  2. the distance supzHdist(z,γ(z))=l,

  3. the parabolic γ moves points on H in the positive direction.

When 0H, we define the anchor of H to be the point zH closest to the origin in the hyperbolic metric.

Definition 3.

We say that a parabolic γAut(D) is positive if γ moves points on H in the positive direction, for some horoball H that is preserved by γ. We say that γ is negative if it is not positive.

We note that γAut(D) is a positive parabolic if and only if γ1 is a negative parabolic. It is clear from the definitions that only positive parabolics in Aut(D) can be adapted to horoballs.

Remark 2.

When γ(z)=z+1 in the upper half-plane model, the horoballs {z:Im(z)C} are dist(iC,iC+1)-adapted. The function Cdist(iC,iC+1) is decreasing, converges to infinity as C0 and to 0 as C, so by continuity l-adapted horoballs exist and are unique for all l>0. Any positive parabolic is conjugate to γ, so the analogous picture holds for an arbitrary positive parabolic.

Our main result is the following.

Lemma 1.

There exist universal constants L,C>0 such that the following holds. Let f:DC be a univalent map, and let H be a horoball not containing 0 that is l-adapted to a parabolic γAut(D), with l<L. Let the anchor of H be z0. We assume that f conjugates γ to a parabolic Möbius transformation. Then for all zH, |f(z)|C|f(z0)|.

Remark 3.

It is crucial that f(D) does not contain infinity in Lemma 1. Without this assumption, the Lemma does not hold. Note that we allow f to be unbounded, or equivalently f(D).

Corollary 1.

Let L,C,l be as in Lemma 1. Let H be a horoball not containing 0 that is l-adapted to a parabolic γΓ, with anchor z0H. Then for any q > 0,H1|f(z)|qdxdyπ4Cq(1|z0|2)2|f(z0)|q.

Proof.

It is standard that f conjugates γ to another parabolic ρ(γ). From Lemma 1 we see thatH1|f(z)|qdxdy1Cq|f(z0)|qHdxdy=1Cq|f(z0)|qπ(1|z0|)24π4Cq(1|z0|2)2|f(z0)|q.

Proof of Lemma 1.

The desired inequality will follow from the estimates in Proposition 1. We first need to conjugate f to a map HC that commutes with zz+1.

Suppose without loss of generality that γ fixes 1. We will construct a commuting square

where A and B are Möbius transformations, and g(z+1)=g(z)+1.

We let A(z)=zλiz+λi for λ>0 chosen such that A1°γ°A(z)=z+1. We choose B depending on finiteness of the fixed point of f°γ°f1.

We first give some preliminary observations that will be useful in both cases. The following Claim relates f and g.

Claim 3. We have for all z, (f°A)(z)=(B°g)(z)g(z)(z+iλ)22λi.

Proof.

We have f=B°g°A1, and A1(z)=λi1+z1z. Thereforef(A(z))=B(g(z))g(z)(A1)(A(z)).

We have (A1)(A(z))=2iλ(1zλiz+λi)2=(z+λi)22λi, and the result follows. □

For zH, we write z=A1(z). In particular z0=A1(z0).

Claim 4. We have Re(z0)=0 and Im(z)Im(z0)λ for all zH.

Proof.

We have A()=1 and A(λi)=0. Since the geodesic ray [0,1) contains z0, the geodesic ray [λi,) contains z0. In particular Re(z0)=0 and Im(z0)λ.

Since H is a horoball containing 1, the horoball A1(H) contains infinity. Since z0A1(H), the second claim follows. □

Case 1 We first assume that f°γ°f1 fixes infinity. We let B(z) = az for aC{0} chosen so that B1°(f°γ°f1)°B(z)=z+1. Since A and B are invertible, the commuting square exists by setting g=B1°f°A. Then f°A conjugates zz+1 to f°γ°f1, and hence B1°f°A=g conjugates zz+1 to itself. In particular g(z+1)=g(z)+1.

It follows from Claim 3 that|f(z)|=|a2λ||g(z)||z+λi|2.

By Claim 4, we have|f(z)||f(z0)|=|g(z)||g(z0)||z+λi|2|z0+λi|2|g(z)||g(z0)|1,where we used Proposition 1 in the second estimate.

Case 2Assume now that f°γ°f1 fixes a point aC. Set B(z)=a+bz, where bC{0} is chosen so that B1°(f°γ°f1)°B(z)=z+1. As in Case 1, both A and B are invertible so the commuting square exists, and g(z+1)=g(z)+1.

By Claim 3,|f(A(z))|=|b2λ|1|g(z)|2|g(z)||z+λi|2.

We have|f(z)||f(z0)|=|g(z0)g(z)|2|g(z)g(z0)||z+λiz0+λi|2.

By Proposition 1, |g(z)g(z0)|1 and there exists αC such that|g(z)z+α|e2πIm(z).

Since infinity is not in the image of f, 0 is not in the image of g. In particular by Proposition 1, log2πIm(α)0. For l small enough, which corresponds to Im(z0) large enough, we have Im(zα)Im(z0)log2πe2πIm(z0), and hence |g(z)||zα|. Therefore|f(z)||f(z0)|=|z0αz0+λi|2/|zαz+λi|2.

Let L be small enough so that for all zA1(H), we have Im(z)1>10log2π>10Im(α). Hence|zαz+λi|2(Re(z)Re(α))2+Im(z)2|z|2+λ2|z|2+Re(α)2|z|2.

Similarly we have|z0αz0+λi|2Re(α)2+Im(z0)2Im(z0)2+λ2Re(α)2+Im(z0)2Im(z0)21+Re(α)2|z|2|zαz+λi|2,where we used Im(z0)λ>0 in the second estimate and |z|2Im(z)2Im(z0)2 in the third estimate. Therefore |f(z)||f(z0)|1, as desired.□

2.3 Dividing the disk into adapted horoballs and a cocompact subset

Recall that to show Theorem 2, we split the domain D into horoballs on which we use Lemma 1, and into the orbit of a compact set. We describe this splitting here, and how to estimate the integral from Theorem 2 over translates of a compact set in the next subsection.

Let L > 0 be the constant from Lemma 1. Choose l arbitrarily with 0<l<min(L,infγΓdist(0,γ(0))). Let H be the union over all positive parabolic γΓ of the (open) horoball Hγ that is l-adapted to γ. We let Φ be the closed Dirichlet fundamental region for Γ centered at 0, and write Φ*=ΦH.

Proposition 2.

The set Φ* is compact with 0 in its interior and is a fundamental domain for the action of Γ on DH.

Proof.

Implicit in the statement of the proposition is the claim that H is Γ-invariant. This follows from uniqueness of l-adapted horoballs that was explained in Remark 2. The final claim follows from this and the fact that Φ is a fundamental domain for the action of Γ on D.

Since D/Γ has finite area, by Siegel’s theorem the Dirichlet fundamental domain is a convex hyperbolic polygon [11, Theorem 4.1.1]. It is well known that any vertex at infinity vΦ¯S1 of this polygon is a fixed point of a parabolic γvΓ [1, Theorem 9.4.5 (4)]. By replacing γv with γv1 if necessary, we may assume that γv is positive. Therefore H contains horoballs Hv:=Hγv that are l-adapted to γv , and hence based at v, for each vΦ¯S1. ThereforeΦ*=ΦHΦvΦ¯S1Hvis bounded, and thus compact.

For any positive parabolic γΓ, since dist(0,γ(0))>l, we have 0Hγ. Moreover since infγΓdist(0,γ(0))>l, we haveinfγΓdist(0,Hγ)>0.

Therefore 0 lies in the interior of Φ*=ΦγHγ. □

2.4 Integral estimates on the lift of a compact part

Recall that one of the steps in the proof of Theorem 2 is bounding DH|f|qdxdy. We do this by splittingDH1|f(z)|qdxdy=γΓγΦ*1|f(z)|qdxdy.

Our goal in this subsection is to show how γΦ*|f|qdxdy depends on γΓ. This follows from the Koebe distortion theorem that guarantees that |f| changes at most by a constant factor over γΦ*.

Proposition 3.

For any q > 0 there exists a constant C=C(Γ,q) such that 1C(1|γ(0)|2)2|f(γ(0))|qγΦ*1|f(z)|qdxdyC(1|γ(0)|2)2|f(γ(0))|qfor all γΓ.

Proof.

By Proposition 2, Φ* is a compact set whose interior contains 0, so we can pick radii r=r(Γ)<R=R(Γ) that depend only on Γ such that B(0,r)Φ*B(0,R). Here we denote by B(z, C) the hyperbolic disk centered at z of radius C.

By the Koebe distortion theorem, we have for zγΦ*,e2R(Γ)e2dist(γ(0),z)|f(γ(0))(1|γ(0)|2)f(z)(1|z|2)|e2dist(γ(0),z)e2R(Γ).

In particular, we have |f(z)|(1|z|2)Γ|f(γ(0))|(1|γ(0)|2).

Since γΦ*B(γ(0),R), and the Euclidean diameter of the hyperbolic disk B(γ(0),R) is R(1|γ(0)|2) up to a bounded multiplicative error, on γΦ*, we have 1|z|21|γ(0)|2. In particular, |f(z)|Γ|f(γ(0))| for zγΦ*.

ThereforeΦ*1|f(z)|qdxdyΓ,q1|f(γ(0))|qγΦ*dxdy.

The region γΦ* contains a hyperbolic disk of radius r, and hence a Euclidean disk of radius rr(1|γ(0)|2). Thereforer2(1|γ(0)|2)2γΦ*dxdyR2(1|γ(0)|2)2,and hence γΦ*dxdyΓ(1|γ(0)|2)2. HenceΦ*1|f(z)|qdxdyΓ,q(1|γ(0)|2)2|f(γ(0))|q.

2.5 Proof of Theorem 2

We note that f(γ(z))=ρ(γ)(f(z)). Differentiating at 0, we see that f(γ(0))γ(0)=ρ(γ)(0)f(0). Therefore(1) f(γ(0))=f(0)ρ(γ)(0)γ(0).(1)

Since γ is a Möbius transformation of the disk, we can write it as γ(z)=λza1a¯z, where aD and |λ|=1. Therefore(2) |γ(0)|=1|a|2=1|γ(0)|2.(2)

Using (1), (2) and Proposition 3, we find that(3) DH1|f(z)|qdxdy=γΓγΦ*1|f(z)|qdxdyΓ,qγΓ|γ(0)|q+2|ρ(γ)(0)|q.(3)

For any positive parabolic γΓ, by Proposition 2 we have 0Hγ. Fix such a γΓ, and let z0 be the anchor of Hγ. Then z0γΦ* for some γΓ. Hence by Corollary 1 we haveHγ1|f(z)|qdxdyq(1|z0|2)2|f(z0)|q.

Since γΦ* has hyperbolic diameter bounded over γΓ, we have by the Koebe distortion theorem|f(γ(0))(1|γ(0)|2)f(z0)(1|z0|2)|Γ1.

We also have as in the proof of Proposition 3,1|z0|2Γ1|γ(0)|2.

ThereforeHγ1|f(z)|qdxdyΓ,q(1|γ(0)|2)2|f(γ(0))|qΓ,qγΦ*1|f(z)|qdxdy.

Since any fundamental region γΦ* intersects at most a bounded number of horoballs in H, we see that(4) H1|f(z)|qdxdyΓ,qDH1|f(z)|qdxdy.(4)

Combining (3) and (4), we see thatD1|f(z)|qdxdyΓ,qDH1|f(z)|qdxdyΓ,qγΓ|γ(0)|q+2|ρ(γ)(0)|q.

3 Grafting a once-punctured torus

In this section and in the rest of the paper, denote by Ω the domain from [Citation12], and by Γ the Kleinian once-punctured torus group that acts on Ω. Here we give a construction of Ω and Γ.

In [Citation12], Ω is obtained by starting with a hyperplane in H3 and bending it along a certain discrete set of geodesics. The result of this is a set in H3 that disconnects H3=S2, and Ω is one of the resulting connected components. Since here we do not work with hyperbolic space H3, we give a concrete description of Ω as the union of regions in the complex plane bounded by 4 or 2 circular arcs.

In this section we work in the upper half-plane, closely following [12, §2]. We denote the Möbius transformations and subsets of H with a bar to distinguish them from their counterparts in the disk model.

We start with a once-punctured torus group generated by Möbius transformationsA¯=(coshλ2coshλ2+1coshλ21coshλ2),B¯=(coshτ2cothλ4sinhτ2sinhτ2coshτ2tanhλ4).

When λ and τ are real, the group A¯,B¯ acts on H and HA¯,B¯ is a once-punctured torus. We fix λ=2cosh132 throughout, and we let τ=λ2+iθ with θ a real parameter, to be set later. We denote by B¯(θ) the Möbius transformation obtained by setting τ=λ2+iθ in the formula above, and set Γ¯(θ)=A¯,B¯(θ). We write Γ¯=Γ¯(0). Define a natural homomorphism f¯θ:Γ¯Γ¯(θ) by f¯θ(A¯)=A¯,f¯θ(B¯(0))=B¯(θ).

Let Φ be the ideal quadrilateral with vertices ±tanhλ4,±cothλ4, so that Φ is a fundamental domain for the action of Γ¯ on H. Let Σ be the union of axes of all elements of Γ¯. Then HΣ is a disjoint union of γInt(Φ), where Int denotes the topological interior.

Define a map ψ¯θ:HΣS2=C{} byψ¯θ(z)=f¯θ(γ)(γ1z),where γΓ¯ is the unique element with the property that zγInt(Φ).

For any hyperbolic element γΓ¯, the axis of γ is a boundary of exactly two (ideal) quadrilaterals η1Φ,η2Φ of the form ηΦ for ηΓ¯. Order η1,η2 so that η11η2{A¯,B¯}. Then ψ¯θ(η2Φ)=f¯θ(η1)f¯θ(η11η2)Φ and ψ¯θ(η1Φ)=f¯θ(η1)Φ. Since f¯θ(A¯)=A¯ we can extend ψ¯θ over all axes in Σ that separate η1Φ and η1A¯Φ. Denote the union of all other axes in Σ by ΣB . For θ<0 there is a region Gθ between Φ and f¯θ(B)Φ bounded by two circular arcs (see ).Ω¯(θ)=ψ¯θ(HΣB)γΓ¯(θ)γGθ.

Fig. 1 The gap Gθ between Φ=f¯θ(id)Φ and f¯θ(B)Φ in the image of ψ¯θ.

Fig. 1 The gap Gθ between Φ=f¯θ(id)Φ and f¯θ(B)Φ in the image of ψ¯θ.

Then Ω¯(θ) is simply connected and Γ¯(θ)-invariant.

Fix η(z)=i1+z1z to be a Möbius transformation that maps D to H. Set θ0=arccos19π.

Definition 4.

We define Ω=η1(Ω¯(θ0)), and Γ=η1°Γ¯(θ0)°η.

We also set f(γ)=η1°f¯θ0(γ)°η and ψ=η1°ψ¯θ0. A picture of Ω can be found in .

Fig. 2 Domain Ω.

Fig. 2 Domain Ω.

4 Numerical results

Let F:ΩD be the Riemann map and f=F1 be its inverse. We denote by Γ0=F°Γ°F1 and by ρ:Γ0Γ the homomorphism induced by f.

By a simple change of coordinates we see that(5) Ω|F|pdxdy=D|f|2pdxdy.(5)

By Theorem 2 and (5), for p > 2 we have FLp(Ω) if and only if(6) γΓ0|γ(0)|p|ρ(γ)(0)|p2<.(6)

In this section we describe how we estimate ρ and Γ0, and how we estimate the range of values of p where the series (6) converges.

Since Γ is a free group of rank 2 generated by A and B (see Section 3), it suffices to estimate ρ1(A) and ρ1(B) for generators A and B of Γ. As explained in Section 3, the Möbius transformations A and B are given explicitly. To estimate ρ1(A)=f1°A°f and ρ1(B)=f1°B°f, we estimate the Riemann mapping f:DΩ using Driscoll’s Schwarz-Christoffel toolbox [Citation5, Citation6], and then compute f1°A°f and f1°B°f.

In Section 4.1 we describe how we get a polygon Ω̂ that is an approximation to Ω. Then in Section 4.2 we explain how we get estimates of ρ1(A) and ρ1(B) using this polygon and the Schwarz-Christoffel toolbox. Finally in Section 4.4 we describe how we show numerically that the sum γΓ0|γ(0)|p|ρ(γ)(0)|2p converges for p = 4, and we check for which values of p it diverges.

4.1 Finite polygonal approximations

We first remind the reader of some notation from Section 3. Recall that we have a once-punctured torus group of Möbius transformations Γ¯ acting on H, a map ψ:HΩ, and a group isomorphism f:Γ¯Γ, so that ψ is f-equivariant. Recall that η=i1+z1z:HD is the Möbius transformation we use to move from the upper half-plane model to the disk model.

Since D/Γ¯ is a once-punctured torus, the limit set Λ(Γ¯) of Γ¯ is S1. Let x0=η(cothλ4)=i1+cothλ41cothλ4=ieλ/2 be the η-image of an ideal vertex of the fundamental domain Φ for the action of Γ¯ on H. It is easy to see that Γ¯x0 is dense in S1. From the construction of ψ, we see that it extends continuously to a surjection ψ:S1Ω. Therefore ψ(Γ¯x0) is dense in Ω and by equivariance ψ(Γ¯x0)=Γψ(x0)=Γx0.

To obtain an approximation to Ω, we generate some number of random points v1,v2,,vn of the form γx0 where γ is a word in A and B of length at most some parameter l>0. These points determine the boundary of a polygon Ω̂ which is an approximation to Ω.

Our goal to use the Schwarz-Christoffel toolbox to compute the Riemann map of this polygon. For this we need the vertices vi to be sorted, which will almost never happen. To sort them, we note that it suffices to sort ψ1(vi). If vi=γix0, then ψ1(vi)=f1(γi)(cothλ4)R by equivariance.This formula is used to compute ψ1(vi)R.

The content of this subsection is summarized in Procedure Citation1. In this procedure, we use the following additional piece of notation. For a word wa,b in the free group a,b of rank 2, and elements x,yG in some group G, denote by w(a, b) the element of G obtained by substituting x for a and y for b in w.

Procedure 1: Constructing finite polygonal approximations.

Input: The number of vertices n, and the maximum word length l.

Output: The vertices v1,v2,,vn of a polygon.

1 Construct a list W1,W2,,W2(3l1) of all words of length at most l in the free group a,b of rank

2. Pick integers 1i1<i2<<in2(3l1) uniformly at random.

3 Set γk=Wik(A,B) and γ¯k=Wik(A¯,B¯).

4 Reorder i1,i2,,in so that γ¯1(cothλ4)<γ¯2(cothλ4)<<γ¯n(cothλ4).

5 Let vk=γk(iexp(λ/2)).

4.2 Procedure for estimating conjugated generators

Here we describe the method we use to estimate ρ1(A)=f1°A°f and ρ1(B)=f1°B°f, assuming that we have an estimate of f:DΩ. We assume that we can compute f(z) and f1(w) for any zD,wΩ, which is the case for Riemann mappings computed using the Schwarz-Christoffel toolbox.

More generally, we describe how to estimate f1°X°f for any Möbius transformation X preserving Ω. Since X preserves Ω, its conjugate f1°X°f preserves the disk. It therefore takes the formf1°X°f=λza1a¯z.

Our methods estimate λS1 and aD. They are summarized in Procedure Citation2. Note that we can compute f and f1 at specific points. We generate some number n of points z1,z2,,znD and then compute wi=f1(X(f(zi))). We estimate λ,a so that λzia1a¯zi is as close as possible to wi . Specifically, we use the sum-of-squares errori=1n|wiλzia1a¯zi|2.

In our implementation, we use MATLAB’s lsqcurvefit to do the optimization.

Procedure 2: Estimating f1°X°f given the Riemann mapping f:DΩ and a Möbius transformation X preserving Ω.

Input: The Riemann mapping f:DΩ, a Möbius transformation X preserving Ω, and the sample size n.

Output: Estimate X̂ of the Möbius transformation f1°X°f.

1 Generate n points z1,z2,,znD uniformly at random.

2 Compute wi=f1(X(f(zi))).

3 Look for the minimizers (λ̂,â) ofi=1n|wiλzia1a¯zi|2over (λ,a)S1×D.

4 Set X̂=λ̂zâ1â¯z.

4.3 Estimates of the conjugated generators

To actually estimate generators of Γ0=f1°Γ°f, we combine the procedures described in Section 4.1 and Section 4.2.

We first compute polygonal approximations to Ω using Procedure Citation1. We set the maximum word length l=12. We construct the following approximations:

  1. the main polygon, Ω̂ that has n = 600 vertices, and

  2. several smaller polygons, Ω̂is for 1i20, each with n = 100 vertices.

We will use Ω̂ to obtain estimates of f1°A°f and f1°B°f, and the smaller polygons Ω̂is to validate these estimates.

Using the Schwarz-Christoffel toolbox, we compute the Riemann mappings f:DΩ̂ and fis:DΩ̂is. We use Procedure Citation2 with f, and X = A and X = B. We denote the resulting estimates λA,aA and λB,aB, and call them the main estimates in this subsection. These give the estimates of f1°A°f and f1°B°f that we will use in the rest of the paper.

For validation, we also run Procedure Citation2 with Ω̂is for i=1,2,,20, and X = A and X = B. We call the resulting estimates λiA,aiA and λiB,aiB. We plot these and the main estimates in and . We see that for each parameter, there is a clear cluster in the estimates obtained from Ω̂is,i=1,2,,20 centered approximately at the estimate obtained from Ω̂. This suggests that λA,aA and λB,aB are accurate.

Fig. 3 The crosses represent values of parameters λiA,aiA of a disk automorphism estimated using smaller 100-point polygonal approximations Ω̂is to Ω. The circles are the estimates of λA and aA obtained from Ω̂.

Fig. 3 The crosses represent values of parameters λiA,aiA of a disk automorphism estimated using smaller 100-point polygonal approximations Ω̂is to Ω. The circles are the estimates of λA and aA obtained from Ω̂.

Fig. 4 The crosses represent values of parameters λiB,aiB of a disk automorphism estimated using smaller 100-point polygonal approximations Ω̂is to Ω. The circles are the estimates of λB and aB obtained from Ω̂.

Fig. 4 The crosses represent values of parameters λiB,aiB of a disk automorphism estimated using smaller 100-point polygonal approximations Ω̂is to Ω. The circles are the estimates of λB and aB obtained from Ω̂.

Definition 5.

We set Γ̂0=Â,B̂, where Â(z)=λAzaA1aA¯z,and B̂(z)=λBzaB1aB¯z.

We also define the homomorphism ρ̂:Γ̂0Γ by setting ρ̂(Â)=A and ρ̂(B̂)=B.

4.4 Values of p for which FLp(Ω)

By Theorem 2 and (5), FLp(Ω) if and only ifγΓ0|γ(0)|p|ρ(γ)(0)|2p<.

We are hence interested in values of p for whichγΓ̂0 |γ(0)|p |ρ̂(γ)(0)|2p<.

We define Γ̂0n to be the set of all elements that have word length n in the generators Â,B̂ of Γ̂0. We then split the above sum inton0Sn(p) where Sn(p)=γΓ̂0n|γ(0)|p|ρ̂(γ)(0)|2p.

We numerically confirm that Sn(4) decays exponentially with n, see .

Fig. 5 The plot of logSn(4) as a function of n for n17. The solid line is the least-squares best linear fit to the shown data points.

Fig. 5 The plot of  log Sn(4) as a function of n for n≤17. The solid line is the least-squares best linear fit to the shown data points.

Recall the definition of p from the Main result in Section 1.4, p=sup{p>0:FLp(Ω)}. By Theorem 2 and (5), this is equal top=sup{p>0:n0Sn(p)<}.

How close p is to 4 gives an indication of how close the domain Ω is to being a counterexample to Brennan’s conjecture. Numerical results show that Sn(5.52) decays exponentially with n and that Sn(5.54) grows exponentially with n, see , meaning that 5.52<p<5.54, showing the Main result.

Fig. 6 The plots of logSn(p) as a function of n for n16 and p=5.52,5.54. The solid lines are the least-squares best linear fits to the data points (n,logSn(p)) for n6, for p=5.52,5.54.

Fig. 6 The plots of  log Sn(p) as a function of n for n≤16 and p=5.52,5.54. The solid lines are the least-squares best linear fits to the data points (n, log Sn(p)) for n≥6, for p=5.52,5.54.

Acknowledgments

I would like to thank Vladimir Marković for introducing me to this problem and for his advice while working on it, and in particular for his help on Lemma 1. I would also like to thank the reviewers for pointing out an error in an earlier draft of this paper, and for their helpful comments and suggestions.

References

  • Beardon, A. F. (1983). The Geometry of Discrete Groups. Grad. Texts in Math. 91. New York, NY: Springer.
  • Bishop, C. J. (2002). Quasiconformal Lipschitz maps, Sullivan’s convex hull theorem, and Brennan’s conjecture. Ark. Mat. 40(1): 1–26.
  • Bishop, C. J. (2007). An explicit constant for Sullivan’s convex hull theorem. In the tradition of Ahlfors and Bers III, Contemp. Math., 355. Providence, RI: American Mathematical Society, 41–69.
  • De Branges, L. (1985). A proof of the Bieberbach conjecture. Acta Math. 154: 137–152. 10.1007/BF02392821
  • Driscoll, T. A. (1996). Algorithm 756: A MATLAB toolbox for Schwarz–Christoffel mapping. ACM Trans. Math. Softw. 22(2): 168–186.
  • Driscoll, T. A. (2005). Algorithm 843: Improvements to the Schwarz-Christoffel toolbox for MATLAB. ACM Trans. Math. Softw. 31(2): 239–251.
  • Epstein, D. B. A., Marden, A. (1987). Convex hulls in hyperbolic space, a theorem of Sullivan, and measured pleated surfaces. Analytical and Geometric Aspects of Hyperbolic Space, London Math. Soc. Lecture Note Ser., 111. Cambridge, UK: Cambridge University Press, 113–253.
  • Epstein, D. B. A., Marden, A., Marković, V. (2004). Quasiconformal homeomorphisms and the convex hull boundary. Ann. Math. 159(1): 305–336.
  • Epstein, D. B. A., Marden, A., Marković, V. (2006). Convex Regions in the Plane and their Domes. Proc. London Math. Soc. 92: 624–654. 10.1017/S002461150501573X
  • Epstein, D. B. A., Marković, V. (2005). The logarithmic spiral: a counterexample to the K = 2 conjecture. Ann. Math. 161(2): 925–957.
  • Katok, S. (1992). Fuchsian groups. Chicago Lectures in Mathematics. Chicago: UCP.
  • Komori, Y., Matthews, C. A. (2006). An explicit counterexample to the equivariant K = 2 conjecture. Conform. Geom. Dyn. 10(10): 184–196.
  • Sullivan, D. (1981). Travaux de Thurston sur les groupes quasi-fuchsiens et les variétés hyperboliques de dimension 3 fibrées sur S1. Lect. Notes Math. 842. Berlin-New York: Springer, 196–214.