这是indexloc提供的服务,不要输入任何密码

Domination between non-Fuchsian representations and anti-de Sitter geometry

Farid Diaf and Abderrahim Mesbah and Nathaniel Sagman Farid Diaf: Institut de Recherche Mathématique Avancée, UMR 7501, Université de Strasbourg et CNRS, 7 rue René Descartes, 67000 Strasbourg, France. f.diaf@unistra.fr Abderrahim Mesbah: Beijing Institute of Mathematical Sciences and Applications, Beijing, China. abderrahimmesbah@bimsa.cn Nathaniel Sagman: Department of Mathematics, University of Toronto, 40 St George Street, Toronto, ON M5S 2E4, Canada. nathaniel.sagman@utoronto.ca
Abstract.

Motivated by work of various authors on domination between surface group representations, harmonic maps, and 33-dimensional anti-de Sitter geometry, we study a new domination problem between non-Fuchsian representations of closed surface groups. We solve the problem for representations that admit branched harmonic immersions, and we show that, outside of this case, the problem cannot always be solved. We then show that a dominating pair gives rise to an anti-de Sitter 33-manifold with singularities, and we construct large families of branched anti-de Sitter 33-manifolds.

1. Introduction

Let Σ\Sigma be an oriented surface with fundamental group Γ\Gamma and let (2,σ)(\mathbb{H}^{2},\sigma) be the hyperbolic space of constant curvature 1-1. As is standard, the group PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) acts on (2,σ)(\mathbb{H}^{2},\sigma) by Möbius transformations. We say that a representation j:ΓPSL(2,)j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) strictly dominates another representation ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) if there exists a map F:22F:\mathbb{H}^{2}\to\mathbb{H}^{2} with equivariance Fj(γ)=ρ(γ)FF\circ j(\gamma)=\rho(\gamma)\circ F, γΓ\gamma\in\Gamma, that contracts hyperbolic distance. Formally, the latter condition means that there exists λ<1\lambda<1 such that for all p,q2,p,q\in\mathbb{H}^{2},

dσ(F(p),F(q))λdσ(p,q).d_{\sigma}(F(p),F(q))\leq\lambda d_{\sigma}(p,q).

Domination has attracted considerable attention and has been studied for maps between spaces other than 2\mathbb{H}^{2}. When Σ\Sigma is closed, it is equivalent to strict domination of the simple translation length spectrum [GK17, Theorem 1.8] (see also [AY19] for a related dynamical application). The original motivation for domination comes from 33-dimensional anti-de Sitter geometry. After Thurston’s geometrization program, it is natural to study Lorentzian structures on 3-manifolds, and the study of the 33-dimensional anti-de Sitter space 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} fits into the broader study of Clifford-Klein forms. Following a long line of research (see [KR85], [Sal00], [Kas10], [GK17], etc.), it was proved that when Σ\Sigma is closed, a strictly dominating pair is equivalent to a closed 33-manifold locally modeled on 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} (more on this below).

A representation j:ΓPSL(2,)j:\Gamma\to\mathrm{PSL}(2,\mathbb{R}) is called Fuchsian if it is discrete and faithful; equivalently, it is the holonomy of a hyperbolic structure on Σ\Sigma. In [DT16], the authors used harmonic maps to show that any non-Fuchsian representation of a closed surface group is dominated by a Fuchsian one. The same domination result was proved in [GKW15] using folded hyperbolic surfaces. The results of [DT16] and [GKW15] have been extended in many directions, including for surfaces with punctures (see [Sag23], [Sag24], [GS22]), for representations to Lie groups of higher rank (see, for instance, [Tho15, Theorem 3], [CTT19, Theorem 4], [DL20, Theorem 1.8 and Conjecture 1.11], [DL22], [BG25]), and for actions on CAT(1)\textrm{CAT}(-1) metric spaces (see [MB23]).

In this paper, we study notions of domination between non-Fuchsian representations of closed surface groups. We refer to a map f:Σ~2f:\widetilde{\Sigma}\to\mathbb{H}^{2} as just ρ\rho-equivariant (or, if ρ\rho is unspecified, equivariant) if for all γΓ\gamma\in\Gamma, fγ=ρ(γ)ff\circ\gamma=\rho(\gamma)\circ f, where the first Γ\Gamma-action is via deck transformations. Toward the definition below, we note that, given an equivariant map f:Σ~2f:\widetilde{\Sigma}\to\mathbb{H}^{2}, the pullback metric descends from Σ~\widetilde{\Sigma} to Σ\Sigma. For C1C^{1} equivariant maps ff and hh, we write fσ<hσf^{*}\sigma<h^{*}\sigma to mean that fσ(v,v)hσ(v,v)f^{*}\sigma(v,v)\leq h^{*}\sigma(v,v) for every unit tangent vector vv, and that the inequality is strict when hσh^{*}\sigma is positive definite. We also recall that when Σ\Sigma is closed and equipped with a Riemann surface structure XX, and ρ\rho is reductive, there exists a ρ\rho-equivariant harmonic map f:X~2f:\widetilde{X}\to\mathbb{H}^{2}; the map is not always unique, but the pullback metric fσf^{*}\sigma is (see Section 2.1).

Definition 1.1.

Let j,ρ:ΓPSL(2,)j,\rho:\Gamma\to\mathrm{PSL}(2,\mathbb{R}) be representations.

  1. (1)

    We say that jj dominates ρ\rho in the pullback sense if there exist C1C^{1} maps h:Σ~2h:\widetilde{\Sigma}\to\mathbb{H}^{2} and f:Σ~2,f:\widetilde{\Sigma}\to\mathbb{H}^{2}, which are respectively jj- and ρ\rho-equivariant, such that fσ<hσ.f^{*}\sigma<h^{*}\sigma.

  2. (2)

    When Σ\Sigma is closed and both jj and ρ\rho are reductive, we say that jj dominates ρ\rho in the harmonic maps sense if there exists a Riemann surface structure XX on Σ\Sigma such that the ρ\rho-equivariant and jj-equivariant harmonic maps f,h:X~2f,h:\widetilde{X}\to\mathbb{H}^{2} satisfy fσ<hσ.f^{*}\sigma<h^{*}\sigma.

The conditions are motivated by [DT16] (see Section 1.2 below). Strict domination implies domination in the pullback sense, and it is equivalent to both domination in the pullback sense and in the harmonic maps sense when jj is a Fuchsian representation of a closed surface group. The first condition is a little flimsy–one has a lot of freedom with ff and hh–while the second one is more rigid.

1.1. Main results

In this paper, the two main questions we address are the following.

  1. (1)

    Given ρ\rho, can we dominate it in the pullback or harmonic maps sense by a non-Fuchsian representation jj?

  2. (2)

    What geometry do dominating pairs encode?

Concerning (1), we provide a complete answer for certain classes of representations (Theorem C and Corollary C), and we also find that there is a menagerie of interesting examples that are worthy of further study (Theorem D). Our results toward (1), together with further context relating to harmonic maps, are contained in Section 1.2 below. At this stage, we can state one positive result, which extends [DT16, Theorem A]. Representations from a closed surface group to PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) are organized into connected components according to an integer invariant called the Euler number (see Section 2.2). The set of possible Euler numbers is [2g+2,2g2][-2g+2,2g-2]\cap\mathbb{Z} and a representation ρ\rho is Fuchsian precisely when eu(ρ)=±(2g2)\textrm{eu}(\rho)=\pm(2g-2), see [Gol80]. Note that when eu(ρ)0\textrm{eu}(\rho)\neq 0, the representation ρ\rho is reductive and the ρ\rho-equivariant harmonic map is unique. Our first domination result concerns the case where this harmonic map is a branched immersion. Recall that a map ff between surfaces has a branch point at a point pp if one can find local coordinates zz and ww centered at pp and f(p)f(p) respectively on which ff takes the form zw=znz\mapsto w=z^{n}. We say that ff is a branched immersion if it is an immersion apart from at a (discrete) set of branch points.

Theorem A.

Let XX be a closed Riemann surface of genus 2\geq 2 with fundamental group Γ\Gamma and let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be a representation with 0<eu(ρ)<2g20<\textrm{eu}(\rho)<2g-2 such that the unique equivariant harmonic map f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma) is a branched immersion. Then, for every kk between 11 and 2g2eu(ρ)2g-2-\textrm{eu}(\rho), we can choose a representation j:ΓPSL(2,)j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) with eu(j)=eu(ρ)+k\textrm{eu}(j)=\textrm{eu}(\rho)+k and equivariant harmonic map hh such that fσ<hσf^{*}\sigma<h^{*}\sigma.

The assumption that the Euler class is positive is not restrictive, since its sign can be flipped by applying an outer automorphism of the fundamental group. Theorem A is essentially a consequence of Theorem C below. Pairs (ρ,f)(\rho,f) as above are constructed and parametrized using Propositions 1.4 and 1.6 below. Equivariant branched immersions are notable because they are equivalent to hyperbolic cone structures on Σ\Sigma (see [Tan94] for explanation). The equivariant harmonic map hh associated with jj is also a branched immersion, so Theorem A can be alternatively cast as a domination result between hyperbolic cone structures. The choices of jj of Euler number eu(ρ)+k\textrm{eu}(\rho)+k are parametrized by effective divisors of degree kk that dominate the branching divisor of ff. For k=2g2eu(ρ),k=2g-2-\textrm{eu}(\rho), there is just one choice, and this is the Fuchsian representation from [DT16, Theorem A].

Theorem C in fact produces many examples of domination data (ρ,j,f,h)(\rho,j,f,h) where ff is not a branched immersion, but there is no immediate geometric description for these maps, so we don’t include them in Theorem A. Nevertheless, these other examples are important for our applications to anti-de Sitter geometry.

For the second question (2), we state our main result here, but we first need to recall the basics on 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}. The three-dimensional anti-de Sitter space 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} is a Lorentzian space form of constant negative sectional curvature, and it can be modeled on PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) with the Killing metric. In this model, the group of space and time-orientation preserving isometries is PSL(2,)2\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})^{2}, acting by right and left multiplication

(A,B)X=AXB1.(A,B)\cdot X=AXB^{-1}.

An anti-de Sitter (AdS) manifold is a Lorentzian manifold of constant negative sectional curvature; equivalently, up to scaling the metric, it is one that is locally isometric to 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}.

For surface group representations j,ρ:ΓPSL(2,)j,\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), we define Γj,ρ<PSL(2,)2\Gamma_{j,\rho}<\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})^{2} by

Γj,ρ={(j(γ),ρ(γ)):γΓ}.\Gamma_{j,\rho}=\{(j(\gamma),\rho(\gamma)):\gamma\in\Gamma\}. (1)

Kulkarni and Raymond proved that every group acting properly discontinuously on 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} identifies with some Γj,ρ\Gamma_{j,\rho}, and that (up to switching factors) jj must be Fuchsian [KR85]. The quotient is a circle bundle over 2/j(Γ)\mathbb{H}^{2}/j(\Gamma), and the circle fibers are timelike geodesics. It was proved by Salein in [Sal00] (if direction) and Kassel in [Kas10] (only if direction) that, when Σ\Sigma is a closed surface of genus at least 22, Γj,ρ\Gamma_{j,\rho} acts properly and discontinuously on 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} if and only if jj strictly dominates ρ\rho (see also [GK17, Theorem 1.8]). For surveys on these results, see [Tho17, Sections 0.1-0.2] and [Sag24, Sections 1.1-1.2].

In her thesis [Jan22], Janigro studied singular anti-de Sitter 33-manifolds that fiber via timelike geodesics over hyperbolic cone surfaces. Janigro defined the notion of a spin-cone AdS 33-manifold, in which the singularities occur along timelike geodesics, generalizing a construction due to Barbot and Meusburger in the context of flat Lorentzian spacetimes, where they modeled massive particles with spin [BM12]. Janigro introduced a weak notion of domination for hyperbolic cone surfaces, which she related to spin-cone AdS 33-manifolds. We extend Janigro’s work by showing that domination in the pullback sense gives rise to a singular AdS 33-manifold (see Theorem 5.3). This extension is obtained by adding a singular set to the AdS 33-manifold constructed by Janigro, thus making the geometric picture more precise. We will see in Section 5 that this singular set is closely related to the singular set of hh, and we note that it also allows for more types of singularities than just spin-cone singularities. We then specialize to the case where the map hh is a branched immersion and analyze the resulting singularities in detail. We arrive at the following theorem.

Theorem B.

Let Σ\Sigma be a hyperbolic surface with fundamental group Γ\Gamma and let ρ,j:ΓPSL(2,)\rho,j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be two representations. Assume that there exist a ρ\rho–equivariant map ff and a jj–equivariant branched immersion hh such that fσ<hσf^{*}\sigma<h^{*}\sigma. Denote by CC the singular set of hh and C0=C/ΓC_{0}=C/\Gamma. Then there exists a three–manifold 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}, topologically a solid torus, with the following properties.

  1. (1)

    The group Γj,ρ\Gamma_{j,\rho} admits a properly discontinuous action on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}.

  2. (2)

    There exists a continuous map :𝔸d𝕊f,hΣ~\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\widetilde{\Sigma} that is equivariant with respect to the action of Γj,ρ\Gamma_{j,\rho} on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} and the action of Γ\Gamma on Σ~\widetilde{\Sigma}. Moreover, the fibers of \mathcal{F} are topological circles.

  3. (3)

    Let L=1(C)L=\mathcal{F}^{-1}(C) and set 𝔸d𝕊f,h:=𝔸d𝕊f,hL\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}:=\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\setminus L. Then 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is an anti-de Sitter manifold, and the restriction of \mathcal{F} to 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is an 𝕊1\mathbb{S}^{1}-principal bundle over Σ~\C\tilde{\Sigma}\backslash C with timelike geodesic fibers.

  4. (4)

    The quotient f,h:=𝔸d𝕊f,h/Γj,ρ\mathcal{M}^{f,h}_{*}:=\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}/\Gamma_{j,\rho} is a branched anti-de Sitter manifold with singular locus L/Γj,ρL/\Gamma_{j,\rho}. The holonomy representation π1(f,h)Isom0(𝔸d𝕊3)\pi_{1}(\mathcal{M}^{f,h}_{*})\to\mathrm{Isom}_{0}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}) factors

    π1(f,h)π1(Σ\C0)π1(Σ)Isom0(𝔸d𝕊3),\pi_{1}(\mathcal{M}^{f,h}_{*})\to\pi_{1}(\Sigma\backslash C_{0})\to\pi_{1}(\Sigma)\to\mathrm{Isom}_{0}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}),

    where the first map is induced by \mathcal{F}, the second is induced by inclusion, and the third map is (j,ρ)(j,\rho).

Branched AdS manifolds are a special case of spin-cone AdS manifolds (see Definition 4.5). For clarity, in Sections 5 and 6, we provide further exposition on Janigro’s work in [Jan22] and explain how our work builds on it. It is worth noting that, in the case where hh is a global diffeomorphism, both 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} and 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} coincide with the anti–de Sitter space 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}, and we thus recover previously known results about closed AdS 33-manifolds.

Combining Theorems A and B, we obtain the following result, which provides a large class of examples of branched anti-de Sitter 33-manifolds.

Corollary B.

Let XX be a closed Riemann surface of genus g2g\geq 2 with fundamental group Γ\Gamma and let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be a representation with 0<eu(ρ)<2g20<\textrm{eu}(\rho)<2g-2 such that the unique equivariant harmonic map f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma) is a branched immersion. Then, for every kk between 11 and 2g2eu(ρ)2g-2-\textrm{eu}(\rho), we can choose a representation j:ΓPSL(2,)j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), eu(j)=eu(ρ)+k\textrm{eu}(j)=\textrm{eu}(\rho)+k, with equivariant map hh such that the data (ρ,j,f,h)(\rho,j,f,h) determines a branched anti-de Sitter 33-manifold as in Theorem B.

A current conjecture, attributed to Goldman, is that every faithful representation of non-zero and non-extremal Euler class uniformizes a branched hyperbolic structure, or, equivalently, satisfies the hypothesis of the corollary. This conjecture is linked in a circle of ideas around a question of Bowditch as well as Goldman’s conjecture about mapping class group dynamics on the PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})-character variety (see, for instance, [Far21]).

As indicated above, we will see in Theorem C that more examples of domination arise than in Theorem A (in particular, where ff is not a branched immersion but hh is), so Corollary B can also be expanded.

Remark 1.2.

If one really wanted to construct a lot of spin-cone AdS 33-manifolds, then one could use harmonic maps from (universal covers of) punctured surfaces and parabolic Higgs bundles. In this paper, we’re interested in the geometry that one can get out of actions of closed surface groups.

1.2. Harmonic maps and domination

Here we give a detailed overview of our main results on harmonic maps and domination. See Sections 2.1-2.2 for preliminaries on harmonic maps.

For motivation, we recall the work of Deroin-Tholozan in [DT16]. Let XX be a closed Riemann surface of genus at least 22 with fundamental group Γ\Gamma and let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be a non-Fuchsian reductive representation with ρ\rho-equivariant harmonic map f:X~2f:\widetilde{X}\to\mathbb{H}^{2}. Every equivariant harmonic map from X~\widetilde{X} determines a holomorphic quadratic differential on XX called the Hopf differential. By [Wol89], there exists a unique Fuchsian representation jj together with an equivariant harmonic diffeomorphism h:X~2h:\widetilde{X}\to\mathbb{H}^{2} with the same Hopf differential as ff. In [DT16], it is shown that

fσ<hσ.f^{*}\sigma<h^{*}\sigma.

The map F=fh1:22F=f\circ h^{-1}:\mathbb{H}^{2}\to\mathbb{H}^{2} is (j,ρ)(j,\rho)-equivariant and strictly contracting, and hence shows that jj strictly dominates ρ\rho. Tholozan went on to prove that every strictly dominating pair arises in this fashion [Tho17]. Moreover, the combined works [DT16] and [Tho17] show the following.

Theorem 1.3 (Deroin-Tholozan [DT16] and Tholozan [Tho17] combined).

Let j,ρ:ΓPSL(2,)j,\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be reductive representations with jj Fuchsian. Then jj strictly dominates ρ\rho if and only if there exists a unique Riemann surface structure XX together with equivariant harmonic maps hh and ff for jj and ρ\rho respectively with the same Hopf differential and such that fσ<hσf^{*}\sigma<h^{*}\sigma.

At this point, Definition 1.1 is well motivated, and we can refine the problem (1): we look for pairs of equivariant harmonic maps with the same Hopf differential such that the domination inequality holds. In order to do so, we need to understand how harmonic maps with the same Hopf differential can differ. Via the non-abelian Hodge correspondence, we establish the following.

Proposition 1.4.

(Proposition 2.7, loosely stated) An equivariant harmonic map f:X~2f:\widetilde{X}\to\mathbb{H}^{2} (up to translation) is equivalent to a pair (ϕ,D)(\phi,D) consisting of a holomorphic quadratic differential ϕ\phi on XX, the Hopf differential of ff, and an effective divisor DD on XX satisfying certain conditions. When ϕ\phi is not zero, DD is the divisor of the square root of the holomorphic energy of ff.

See Proposition 2.7 for the precise statement (which includes the case ϕ=0\phi=0). When D=0D=0, one recovers Wolf’s parametrization of Teichmüller space [Wol89], and when ϕ=0\phi=0, one is led to Troyanov’s uniformization for hyperbolic branched metrics [Tro91]. Proposition 1.4 is probably known to some experts (compare with [Hit87, Theorem 10.8])–see Section 2.3 for explanation.

Now we can make our problem (1) more precise: for which pairs (ϕ,D1)(\phi,D_{1}) and (ϕ,D2)(\phi,D_{2}) do we have domination? The problems turns out to be quite delicate, but we find a complete solution when (ϕ,D2)(\phi,D_{2}) gives rise to a branched harmonic immersion. Proposition 1.6, although not necessary for the proof, tells us when we have a branched immersion.

Definition 1.5.

Let D1D_{1} and D2D_{2} be divisors on XX. We write D2<D1D_{2}<D_{1} to mean that D2(p)D1(p)D_{2}(p)\leq D_{1}(p) for all pXp\in X and that there exists qXq\in X such that D2(q)<D1(q)D_{2}(q)<D_{1}(q).

Proposition 1.6.

An equivariant harmonic map f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma) associated with a pair (ϕ,D)(\phi,D), ϕ0\phi\neq 0, is a branched immersion if and only if 2D<(ϕ)2D<(\phi) or (ϕ)<2D(\phi)<2D.

When ϕ=0\phi=0, the harmonic map is weakly conformal and automatically a branched immersion. The conditions 2D<(ϕ)2D<(\phi) and (ϕ)<2D(\phi)<2D correspond to positive and negative Euler numbers respectfully. The key elements of Proposition 1.6 are contained in the literature, but the result has never been stated as above. The “only if” direction is observed in [BBDH21, Lemma 3.2], and the “if” direction follows from [SG24, Theorem 4.1, Proposition 6.5] (which improves [BBDH21, Theorems 3.4 and 4.1]). On route to Theorem C, we establish all the tools necessary to prove Proposition 1.6, so for the convenience of the reader we write out a proof. We will also prove a version of Proposition 1.6 for surfaces with boundary (see Lemma 3.8).

Finally, we state our solution to the domination problem for branched harmonic immersions. The theorem concerns dominating arbitrary maps by branched immersions, and the corollary is about dominating branched immersions themselves.

Theorem C.

Let f,h:X~(2,σ)f,h:\widetilde{X}\to(\mathbb{H}^{2},\sigma) be equivariant harmonic maps with holomorphic data (ϕ,D1)(\phi,D_{1}) and (ϕ,D2)(\phi,D_{2}) respectively, with degD1,degD22g2\deg D_{1},\deg D_{2}\leq 2g-2. Assuming hh is a branched immersion, then

fσ<hσif and only ifD2<D1 and D1+D2<(ϕ).f^{*}\sigma<h^{*}\sigma\quad\text{if and only if}\quad D_{2}<D_{1}\textrm{ and }D_{1}+D_{2}<(\phi).
Corollary C.

Let f,h:X~(2,σ)f,h:\widetilde{X}\to(\mathbb{H}^{2},\sigma) be equivariant harmonic maps with holomorphic data (ϕ,D1)(\phi,D_{1}) and (ϕ,D2)(\phi,D_{2}) respectively, with degD1,degD22g2\deg D_{1},\deg D_{2}\leq 2g-2. Assuming ff is a branched immersion, then

fσ<hσif and only ifD2<D1.f^{*}\sigma<h^{*}\sigma\quad\text{if and only if}\quad D_{2}<D_{1}.

The assumption on the degrees is not necessary, but just keeps the statement cleaner. If degD22g2\deg D_{2}\geq 2g-2, then we can flip back to the case degD22g2\deg D_{2}\leq 2g-2 by using an outer automorphism of the fundamental group. A harmonic map with data (ϕ,D)(\phi,D) becomes a map with data (ϕ,(ϕ)D)(\phi,(\phi)-D), and the pullback metric is unaffected, and thus the condition D2<D1D_{2}<D_{1} becomes (ϕ)D2<D1(\phi)-D_{2}<D_{1}. Similar if degD12g2.\deg D_{1}\geq 2g-2. It is worth noting that if hh is not a branched immersion, the condition D2<D1D_{2}<D_{1} does not guarantee domination; see Proposition 3.6.

The main result toward the proof of Theorem C is Proposition 3.2, which is of independent interest. It provides an inequality between functions satisfying a Bochner-type equation, generalizing the key inequality from [SY97, Section 1.8] and [DT16, Lemma 2.6]. This proposition is also a key step in establishing Proposition 1.6.

Remark 1.7.

Given harmonic maps f,h:X~(2,σ)f,h:\widetilde{X}\to(\mathbb{H}^{2},\sigma) with the same Hopf differential, the map F=(f,h):X~(2×2,σ(σ))F=(f,h):\widetilde{X}\to(\mathbb{H}^{2}\times\mathbb{H}^{2},\sigma\oplus(-\sigma)) determines a maximal surface (zero mean curvature at immersed points). The domination condition fσ<hσf^{*}\sigma<h^{*}\sigma says that FF is a spacelike immersion off the singular set of hh.

Remark 1.8.

In the works [DT16], [Tho17], [Sag23], [Sag24], the main domain results concern pairs (ρ,j),(\rho,j), where ρ\rho is an action by isometries on a CAT(1)\textrm{CAT}(-1) Hadamard manifold MM and jj is a Fuchsian representation to PSL(2,).\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}). It should be possible to prove extensions of Theorems A and C for such pairs. Toward this, the key observations are that, for maps to MM, there is a notion of holomorphic energy (see [DT16, Section 2.2]) and we have a Bochner formula (see [DT16, Lemma 2.3]).

With our application to anti-de Sitter geometry in mind, we are mainly concerned with dominating by branched immersions, which puts us in the case |eu(ρ)|<|eu(j)||\textrm{eu}(\rho)|<|\textrm{eu}(j)|. That being said, after Theorem C, it is natural to inquire about the equality case |eu(ρ)|=|eu(j)|.|\textrm{eu}(\rho)|=|\textrm{eu}(j)|. Since it’s not focused toward an application in this paper, we don’t carry out a general treatment, but we just give examples that point to future directions.

Theorem D.

For all even g2g\geq 2, there exists a Riemann surface YY with fundamental group Γ\Gamma such that the following holds.

  1. (1)

    There exist representations ρ,j:ΓPSL(2,)\rho,j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) with equivariant harmonic maps f,h:Y~(2,σ)f,h:\widetilde{Y}\to(\mathbb{H}^{2},\sigma) respectively, such that eu(ρ)=eu(j)=0\textrm{eu}(\rho)=\textrm{eu}(j)=0 and fσ<hσf^{*}\sigma<h^{*}\sigma.

  2. (2)

    There exists a representation ρ\rho with equivariant harmonic map ff with the property that there exist no pairs (j,h)(j,h) consisting of a non-Fuchsian representation jj and an equivariant harmonic map hh with the same Hopf differential such that fσ<hσf^{*}\sigma<h^{*}\sigma.

In (2), the underlying representation has Euler number zero.

1.3. Future directions

We list a few research directions that follow this work.

1.3.1. Length spectrum domination

In line with [GKW15], it is natural to study pairs ρ,j:ΓPSL(2,)\rho,j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) such that, for all γΓ,\gamma\in\Gamma, (ρ(γ))(j(γ))\ell(\rho(\gamma))\leq\ell(j(\gamma)), where ()\ell(\cdot) is the hyperbolic translation length. It would be interesting to compare this length spectrum domination with the domination considered here.

1.3.2. Higher rank

Our domination problems generalize easily for representations to a higher rank non-compact semisimple Lie group GG, with the space 2\mathbb{H}^{2} replaced with a Riemannian symmetric space of GG. For generalizations of [DT16], with Fuchsian representations replaced with Hitchin representations to G=SL(n,)G=\textrm{SL}(n,\mathbb{R}) see [DL20, Theorem 1.8 and Conjecture 1.11] and [DL22]. Extending our results in this paper seems approachable for harmonic maps arising from Coxeter cyclic GG-Higgs bundles, which are equivalent to solutions to affine Toda equations (see [ST25]). There is even a Toledo number for cyclic GG-Higgs bundles, which generalizes the Euler number (see [GPG24]).

1.3.3. More on singular AdS 33-manifolds

Our constructions in Section 5 (in particular, Theorem 5.3) allow the construction of general singular AdS 33-manifolds from dominating pairs, rather than only branched AdS 33-manifolds arising from branched immersions. By developing more exotic examples of singular AdS 33-manifolds, one may gain further insight into the nature of singularities in anti–de Sitter geometry.

1.4. Outline of the paper

In Sections 2 and 3, we study the domination problem for harmonic maps, while in Sections 4, 5, and 6, we study anti-de Sitter geometry and construct singular anti-de Sitter 33-manifolds from dominating pairs. Sections 2-3 and 4-6 can essentially be read independently.

In more detail, Section 2 provides preliminaries on harmonic maps and carefully states and proves Proposition 2.7, which characterizes harmonic maps via holomorphic data (ϕ,D)(\phi,D). In Section 3, we address the domination problem and prove Proposition 1.6 as well as Theorems A, C, and D. In Section 4, we introduce spin-cone AdS 33-manifolds, and in Section 5, we construct singular AdS 33-manifolds from general dominating pairs (Theorem 5.3). Finally, in Section 6, we specialize to branched immersions and show that, in this case, the examples from Section 5 are branched AdS 33-manifolds, thereby proving Theorem B.

1.5. Acknowledgements

This project stems from discussions with Andrea Seppi, to whom we are very grateful, and to whom we thank further for helpful comments on the first draft. We also thank Francesco Bonsante for sharing the thesis [Jan22].

2. Preliminaries on harmonic maps

2.1. Equivariant harmonic maps

Let Σg\Sigma_{g} be a closed oriented surface of genus at least 22 and let XX be a Riemann surface structure on Σg\Sigma_{g} with universal cover X~\tilde{X}. Let (M,σ)(M,\sigma) be a Riemannian manifold and let f:X~Mf:\widetilde{X}\to M be a C2C^{2} map. The derivative dfdf defines a section of the endomorphism bundle TX~fTMT^{*}\widetilde{X}\otimes f^{*}TM. Complexifying TX~fTMT^{*}\widetilde{X}\otimes f^{*}TM, let df=f+¯fdf=\partial f+\overline{\partial}f be the decomposition into (1,0)(1,0) and (0,1)(0,1) parts. We denote by \nabla the connection on fTMf^{*}TM\otimes\mathbb{C} induced by the Levi-Civita connection of σ\sigma, and we also use \nabla for its extension to fTMf^{*}TM\otimes\mathbb{C}-valued forms. The definition below depends on the Riemann surface structure XX and the metric σ.\sigma.

Definition 2.1.

The map ff is harmonic if 0,1f=0.\nabla^{0,1}\partial f=0.

In this paper, we are concerned with equivariant harmonic maps from X~\widetilde{X} to the space (2,σ)(\mathbb{H}^{2},\sigma), where 2\mathbb{H}^{2} is the 22-dimensional upper half-plane and σ\sigma is a hyperbolic metric of constant curvature 1-1. The starting point is the existence theorem of Donaldson [Don87] and Corlette [Cor88]. A representation ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) is irreducible if it is not contained in a parabolic subgroup, and ρ\rho is reductive if the Zariski closure of ρ(Γ)\rho(\Gamma) is a reductive group. Geometrically, and perhaps more intuitively, ρ\rho is irreducible if the induced action of Γ\Gamma on the Gromov boundary \partial_{\infty}\mathbb{H} has no global fixed point, and ρ\rho is reductive if ρ(Γ)\rho(\Gamma) preserves a geodesic in 2.\mathbb{H}^{2}.

Theorem 2.2 (Donaldson, Corlette).

Let XX be a Riemann surface structure on Σg\Sigma_{g} and let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be a representation, acting by isometries on (2,σ).(\mathbb{H}^{2},\sigma). Then there exists a ρ\rho-equivariant harmonic map f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma) if and only if ρ\rho is reductive.

By Sampson’s argument in [Sam78, Theorem 3], if ρ\rho is irreducible, then ff is unique. When ρ\rho is just reductive and not irreducible, again by [Sam78, Theorem 3], there is a 11-parameter family of harmonic maps, each of which is a parametrization onto the invariant geodesic, and all of the harmonic maps are related by an isometric translation along the geodesic. Since the harmonic maps are related by isometries, the pullback metric never depends on the choice of harmonic map.

2.2. Energies and Hopf differentials

Given a harmonic map from a Riemann surface, one can associate a number of analytic objects. For a detailed reference, see [SY97, Section 1-2]. To begin, let XX be a Riemann surface as above and fix a metric ν\nu on XX that’s compatible with the Riemann surface structure. We use ν\nu as well to denote the lift to a metric on the universal cover X~\widetilde{X}.

Keeping things general for now, let ρ:ΓIsom(M,σ)\rho:\Gamma\to\textrm{Isom}(M,\sigma) be an action by isometries and let f:X~(M,σ)f:\widetilde{X}\to(M,\sigma) be a ρ\rho-equivariant C2C^{2} map. The (possibly degenerate) pullback metric fσf^{*}\sigma extends bilinearly to the complexified tangent bundle of X~\widetilde{X} and then decomposes into types as

fσ=2σ(f,¯f)+σ(f,f)+σ(¯f,¯f).f^{*}\sigma=2\sigma(\partial f,\overline{\partial}f)+\sigma(\partial f,\partial f)+\sigma(\overline{\partial}f,\overline{\partial}f). (2)

We write 2σ(f,¯f)=e(f)ν,2\sigma(\partial f,\overline{\partial}f)=e(f)\nu, where e(f)e(f) is called the energy density function. As well, the quadratic differential ϕ(f):=σ(f,f)\phi(f):=\sigma(\partial f,\partial f) is called the Hopf differential of ff (note also that ϕ¯(f)=σ(¯f,¯f)\overline{\phi}(f)=\sigma(\overline{\partial}f,\overline{\partial}f)). Since ff is ρ\rho-equivariant, e(f),ϕ(f),e(f),\phi(f), ϕ¯(f)\overline{\phi}(f), and fσf^{*}\sigma are invariant under the action of Γ\Gamma and hence descend to XX. The formula (2) is rewritten as

fσ=e(f)ν+ϕ(f)+ϕ¯(f).f^{*}\sigma=e(f)\nu+\phi(f)+\overline{\phi}(f). (3)

The significance of e(f)e(f) stems from the fact that the equation 0,1f=0\nabla^{0,1}\partial f=0 can be seen as the Euler-Lagrange equation for the Dirichlet energy

(X,f)=Σge(f)𝑑Aν,\mathcal{E}(X,f)=\int_{\Sigma_{g}}e(f)dA_{\nu},

where dAνdA_{\nu} is the area form of ν\nu (note (X,f)\mathcal{E}(X,f) does not depend on the choice of ν\nu). That is, harmonic maps are equivalently critical points of (X,)\mathcal{E}(X,\cdot). As for the Hopf differential ϕ(f)\phi(f), it is holomorphic when ff is harmonic, and when the target has dimension at most 2,2, so for example when M=2,M=\mathbb{H}^{2}, the converse holds as well.

We now set M=2M=\mathbb{H}^{2}. In this special case, there are more analytic quantities to probe the harmonic map. Working in local coordinates, where σ=σ(w)|dw|2\sigma=\sigma(w)|dw|^{2}, ν=ν(z)|dz|2\nu=\nu(z)|dz|^{2}, we set

H(f)=σ(f(z))ν(z)|fz|2,L(f)=σ(f(z))ν(z)|fz¯|2.H(f)=\frac{\sigma(f(z))}{\nu(z)}\Big|\frac{\partial f}{\partial z}\Big|^{2},\hskip 2.84526ptL(f)=\frac{\sigma(f(z))}{\nu(z)}\Big|\frac{\partial f}{\partial\overline{z}}\Big|^{2}.

The function H(f)H(f) is called the holomorphic energy and L(f)L(f) is called the anti-holomorphic energy. They satisfy

e(f)=H(f)+L(f),|ϕ(f)|2ν2=H(f)L(f),e(f)=H(f)+L(f),\hskip 2.84526pt|\phi(f)|^{2}\nu^{-2}=H(f)L(f),

and the vanishing of the Jacobian of ff is equivalent to the vanishing of the function

J(f):=H(f)L(f).J(f):=H(f)-L(f).

When ff is harmonic, the functions H(f)H(f) and L(f)L(f) are either identically zero or have isolated zeros and satisfy the Bochner formulae: away from the zeros of H(f),H(f),

12ΔνlogH(f)=κσ(H(f)L(f))+κν,\frac{1}{2}\Delta_{\nu}\log H(f)=-\kappa_{\sigma}(H(f)-L(f))+\kappa_{\nu}, (4)

where κσ\kappa_{\sigma} and κν\kappa_{\nu} are the sectional curvatures of σ\sigma and ν\nu, respectively (so κσ=1\kappa_{\sigma}=-1), and away from the zeros of L(f),L(f),

12ΔνlogL(f)=κσ(L(f)H(f))+κν\frac{1}{2}\Delta_{\nu}\log L(f)=-\kappa_{\sigma}(L(f)-H(f))+\kappa_{\nu}

(see [SY97, Section 1.7]). Note that the Bochner formula for L(f)L(f) can be seen as a consequence of the Bochner formula for H(f)H(f) and the equation |ϕ(f)|2ν2=H(f)L(f)|\phi(f)|^{2}\nu^{-2}=H(f)L(f). Note that for a non-irreducible representation, again using that harmonic maps are related by isometries, H(f)H(f) is independent of the choice of harmonic map.

We end this subsection with a discussion on Euler numbers. The Euler number of a representation ρ,\rho, denoted eu(ρ),\textrm{eu}(\rho), can be defined in many ways; for instance, using characteristic classes, or as an obstruction to lifting ρ\rho to PSL(2,)~\widetilde{\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})}. Here, we give the most naive definition:

eu(ρ):=12πΣg𝑑Afσ,\textrm{eu}(\rho):=\frac{1}{2\pi}\int_{\Sigma_{g}}dA_{f^{*}\sigma}, (5)

where f:Σ~g2f:\widetilde{\Sigma}_{g}\to\mathbb{H}^{2} is any ρ\rho-equivariant C2C^{2} map, and dAfσdA_{f^{*}\sigma} is the area form of fσf^{*}\sigma (which might be degenerate). By an application of Stoke’s theorem, the integral above indeed depends only on ρ\rho and not on ff. It is also useful to recall the equality of 22-forms

dAfσ=J(f)dAν.dA_{f^{*}\sigma}=J(f)dA_{\nu}. (6)

The following lemma allows us to make Definition 2.4 below.

Lemma 2.3.

Let XX be a Riemann surface structure on Σg\Sigma_{g} and let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be a reductive representation with equivariant harmonic map f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma). If eu(ρ)0\textrm{eu}(\rho)\geq 0, then H(f)H(f) cannot be identically zero.

Proof.

By (5), (6), and J(f)=H(f)L(f)J(f)=H(f)-L(f), we have

12πΣgH(f)𝑑Aν=eu(ρ)+12πΣgL(f)𝑑Aν.\frac{1}{2\pi}\int_{\Sigma_{g}}H(f)dA_{\nu}=\textrm{eu}(\rho)+\frac{1}{2\pi}\int_{\Sigma_{g}}L(f)dA_{\nu}. (7)

If eu(ρ)>0\textrm{eu}(\rho)>0 then the right hand side above is strict, so H(f)H(f) has to be positive somewhere. If eu(ρ)=0\textrm{eu}(\rho)=0 and H(f)=0H(f)=0 identically, then the right hand side shows that ff is a constant map, which is impossible. ∎

Definition 2.4.

Let XX be a Riemann surface structure on Σg\Sigma_{g} and let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be a representation with eu(ρ)0\textrm{eu}(\rho)\geq 0 and equivariant harmonic map f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma). The divisor of (X,ρ)(X,\rho), denoted DX(ρ),D_{X}(\rho), is the divisor of the square root of the holomorphic energy of ff.

Starting from the characterization (5), a classical argument, which involves the Bochner formula, (6), and the Gauss-Bonnet theorem, can be applied nearly verbatim to prove the following. The argument can be found in [SY97, pp. 11-12].

Proposition 2.5.

Let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be reductive with eu(ρ)0\textrm{eu}(\rho)\geq 0. Then eu(ρ)=2g2degDX(ρ).\textrm{eu}(\rho)=2g-2-\deg D_{X}(\rho).

If we reverse the orientation of Σg\Sigma_{g} or precompose ρ\rho with an outer automorphism of PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), then the Euler numbers flip sign. For this reason, we often restrict to the case eu(ρ)0\textrm{eu}(\rho)\geq 0. For eu(ρ)0\textrm{eu}(\rho)\leq 0, one has results analogous to above, with H(f)H(f) replaced with L(f).L(f).

2.3. Divisors and Higgs bundles

The purpose of this section is to prove Proposition 2.7 below (or, Proposition 1.4). We use Higgs bundles and the non-abelian Hodge correspondence; Higgs bundles will not come up again, so the unfamiliar reader might benefit from skipping the proofs on a first reading.

For a non-zero holomorphic section α\alpha of a holomorphic line bundle on XX, we denote the divisor by (α).(\alpha). Let 𝒦\mathcal{K} be the canonical bundle of X,X, so that H0(X,𝒦2)H^{0}(X,\mathcal{K}^{2}) is the space of holomorphic quadratic differentials on XX.

Definition 2.6.

Let 𝒟\mathcal{D} be the set of pairs (ϕ,D),(\phi,D), where DD is an effective divisor on XX with degD2g2\deg D\leq 2g-2 and ϕH0(X,𝒦2)\phi\in H^{0}(X,\mathcal{K}^{2}), such that, if ϕ0\phi\neq 0, then D(ϕ)D\leq(\phi) (as functions). We add the further condition that if degD=2g2,\deg D=2g-2, then ϕ0\phi\neq 0.

Proposition 2.7.

The set 𝒟\mathcal{D} is in bijection with the space of conjugacy classes of reductive representations from Γ\Gamma to PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) with non-negative Euler class. If (ϕ,D)(\phi,D) is associated with ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), then D=DX(ρ)D=D_{X}(\rho) and ϕ\phi is the Hopf differential of any ρ\rho-equivariant harmonic map.

To parametrize representations with non-positive Euler numbers, one takes pairs (ϕ,D)(\phi,D) such that degD2g2\deg D\geq 2g-2 and D(ϕ)D\geq(\phi); simply use an outer automorphism. About the case degD=2g2\deg D=2g-2, the equation (7) shows that if degD=2g2\deg D=2g-2, i.e, eu(ρ)=0\textrm{eu}(\rho)=0, then no reductive representation can carry a harmonic map ff with ϕ(f)=0\phi(f)=0, for then we would have H(f)=L(f)=0H(f)=L(f)=0.

Proposition 2.7 should be known to experts. The parametrization of the character variety by pairs (ϕ,D)(\phi,D) is proved in Hitchin’s original paper [Hit87, Theorem 10.8] (outside of Euler number 0, although comments are made on that case), but the characterization in terms of zeros of the holomorphic energy does not appear to be recorded. We essentially redo Hitchin’s proof in a different language, and explain, from our point of view, how DD comes from the holomorphic energy.

2.3.1. GG-Higgs bundles

Since we’re considering the adjoint group PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), we work with GG-Higgs bundles. Representations with even Euler class lift to SL(2,),\textrm{SL}(2,\mathbb{R}), and so for those representations one could use linear Higgs bundles. In general, one could transfer to the linear setting using the isomorphism PSL(2,)SO0(1,2)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\simeq\textrm{SO}_{0}(1,2), but the proofs are a bit faster and more natural in the principal bundle setting.

There are many excellent sources on GG-Higgs bundles, and we don’t need to recall everything here. We mostly draw on [ST25], which we refer to for more details. Let GG be a non-compact simple complex Lie group with maximal compact subgroup KK. The space G/K,G/K, equipped with the metric induced by the Killing form on the Lie algebra of G,G, is a Riemannian symmetric space of non-compact type. For G=PSL(2,)G=\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), this symmetric space is 2\mathbb{H}^{2}. Let 𝔤\mathfrak{g} and 𝔨\mathfrak{k} be the Lie algebras of GG and KK respectively, and let 𝔭\mathfrak{p} be the Killing orthogonal complement of 𝔨\mathfrak{k} in 𝔤\mathfrak{g}. We write G,K,𝔤,G^{\mathbb{C}},K^{\mathbb{C}},\mathfrak{g}^{\mathbb{C}}, etc., for complexifications.

Definition 2.8.

A GG-Higgs bundle over a Riemann surface XX is a pair (P,Φ),(P,\Phi), where PP is a principal KK^{\mathbb{C}}-bundle and Φ\Phi is a holomorphic 11-form valued in the associated bundle P×Ad|K×𝔭P\times_{\textrm{Ad}|_{K^{\mathbb{C}}}}\times\mathfrak{p}^{\mathbb{C}} called the Higgs field.

See [ST25, Section 2.2] for more details on the discussion below. As is well known, an equivariant harmonic map ff to G/KG/K gives rise to a GG-Higgs bundle. Very briefly, GG/KG\to G/K is a principal KK-bundle that carries a principal connection induced from the Maurer-Cartan form on GG. The map ff pulls back a KK-bundle QQ over XX with a principal connection. Using the Maurer-Cartan isomorphism, the derivative of ff identifies as a 11-form valued in Q×Ad|K𝔭Q\times_{\textrm{Ad}|_{K}}\mathfrak{p}, say ψ\psi. Upon complexifying QQ to a principal KK^{\mathbb{C}}-bundle PP, we can split ψ\psi into types as ψ=ψ1,0+ψ0,1\psi=\psi^{1,0}+\psi^{0,1}, and the harmonicity of ff is equivalent to the assertion that, with respect to the Koszul-Malgrange holomorphic structure on PXP\to X associated with the principal connection on QQ, ψ1,0\psi^{1,0} is a Higgs field. Given a GG-Higgs bundle, it arises from a harmonic map, in a way that undoes the procedure above, if and only if one can solve Hitchin’s self-duality equations (see [ST25, Definition 2.3]).

2.3.2. Proposition 2.7

As explained in [ST25, Section 3.5], a PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})-Higgs bundle is equivalent to a PSL(2,)\textrm{PSL}(2,\mathbb{C})-Higgs bundle equipped with a holomorphic gauge transformation ss of PP satisfying certain properties and such that sϕ=ϕs^{*}\phi=-\phi. In the language of [ST25], (P,ϕ)(P,\phi) is Coxeter cyclic. By [ST25, Proposition 1.2], the Higgs bundle is equivalent to the data of a line bundle LL and sections α\alpha and β\beta of L𝒦L\otimes\mathcal{K} and L1𝒦L^{-1}\otimes\mathcal{K} respectively. An argument of Hitchin from [Hit87, Section 10] shows that we can always specify things so that, when the GG-Higgs bundle arises from an equivariant harmonic map, then degL\deg L is the Euler number of the underlying representation. In this case, the Hopf differential, up to dividing by a positive scalar cc, is the product αβ:=αβ\alpha\beta:=\alpha\otimes\beta. From the proof of [ST25, Proposition 1.2], under this constraint, two GG-Higgs bundles giving triples (L,α,β)(L,\alpha,\beta) and (L,α,β)(L^{\prime},\alpha^{\prime},\beta^{\prime}) as above are isomorphic if and only if there exists an isomorphism from LLL\to L^{\prime} that intertwines α\alpha with α\alpha^{\prime} and β\beta with β\beta^{\prime}.

From [ST25, Theorems A and 4.3], once we’ve fixed a conformal metric ν\nu on XX, a solution to Hitchin’s self-duality equations is equivalent to a Hermitian metric μ\mu on LL (with dual metric μ1\mu^{-1} on L1L^{-1}) solving the equation, for functions eα=μ(α,α)ν2e_{\alpha}=\mu(\alpha,\alpha)\nu^{-2} and eβ=μ1(β,β)ν2e_{\beta}=\mu^{-1}(\beta,\beta)\nu^{-2} on XX, away from their zeros,

12Δνlogeα=eαeβ+κν,12Δνlogeβ=eβeα+κν.\frac{1}{2}\Delta_{\nu}\log e_{\alpha}=e_{\alpha}-e_{\beta}+\kappa_{\nu},\hskip 2.84526pt\frac{1}{2}\Delta_{\nu}\log e_{\beta}=e_{\beta}-e_{\alpha}+\kappa_{\nu}. (8)

It is also shown in [ST25] that eαeβ=|ϕ|2ν2e_{\alpha}e_{\beta}=|\phi|^{2}\nu^{-2}, where ϕ\phi is the Hopf differential of the harmonic map, and that eα+eβe_{\alpha}+e_{\beta} is the energy density of the harmonic map. There is one subtlety: distinct solutions to Hitchin’s equations could produce the same solution to (8). Also, note that [ST25, Theorems A] concerns solutions to Hitchin’s equations for stable GG-Higgs bundles, but the general case follows using [ST25, Remark 3.17] (it is exactly this latter case in which we can have multiple solutions). From the relation eαeβ=|ϕ|2ν2e_{\alpha}e_{\beta}=|\phi|^{2}\nu^{-2}, solving (8) requires solving only for eαe_{\alpha}, and the equations reduce to the Bochner formula (4).

The equations (8) are a basic example of affine Toda equations. Existence and uniqueness results for affine Toda equations were recently established in [McI25, Theorem 1.3], which shows in our context (the very simplest case of [McI25, Theorem 1.3]) that, given (L,α,β)(L,\alpha,\beta), if α,β0\alpha,\beta\neq 0, then (8) has a unique solution for eαe_{\alpha} with a prescribed vanishing divisor. By virtue of L𝒦L\otimes\mathcal{K} and L1𝒦L^{-1}\otimes\mathcal{K} having non-vanishing sections, it’s implicit in this case that 22gdegL2g22-2g\leq\deg L\leq 2g-2. If either of α\alpha or β\beta is equal to zero, [McI25, Theorem 1.3] shows that one has a solution if and only if 22gdegL2g22-2g\leq\deg L\leq 2g-2 and degLdegL1\deg L\neq\deg L^{-1} (degL=degL1\deg L=\deg L^{-1} can occur only if the common degree is 0) and moreover the solution is unique. We prove the following.

Proposition 2.9.

Assume that degL>0\deg L>0 and that (8) admits a solution with associated representation ρ\rho and harmonic map ff. Then H(f)=eβH(f)=e_{\beta} and L(f)=eαL(f)=e_{\alpha}.

Proof.

Let e(f)e(f) and ϕ(f)\phi(f) be the energy and Hopf differential of ff respectively. Consider the function, on X×X\times\mathbb{C},

p(z,λ)=λ2e(f)(z)λ+(|ϕ(f)|2ν2)(z).p(z,\lambda)=\lambda^{2}-e(f)(z)\lambda+(|\phi(f)|^{2}\nu^{-2})(z).

Over each point zz of XX, {eα(z),eβ(z)}\{e_{\alpha}(z),e_{\beta}(z)\} and {H(f)(z),L(f)(z)}\{H(f)(z),L(f)(z)\} are the zeros of p(z,)p(z,\cdot). Since all of the functions in question are real analytic (for they solve a semi-linear elliptic PDE with real analytic coefficients), it follows that, as sets of functions on XX, {eα,eβ}={H(f),L(f)}.\{e_{\alpha},e_{\beta}\}=\{H(f),L(f)\}. Since degL>0,\deg L>0, so that degLdegL1\deg L\neq\deg L^{-1}, it follows that deg(eα)deg(eβ)\deg(e_{\alpha})\neq\deg(e_{\beta}), for their divisors agree with that of α\alpha and β\beta respectively. By Proposition 2.5, deg(eβ)\deg(e_{\beta}) captures the vanishing divisor of H(f)H(f), and hence H(f)=eβH(f)=e_{\beta} and L(f)=eα.L(f)=e_{\alpha}.

Note that the proof used degL0\deg L\neq 0 only in the last line. If degL=0\deg L=0, then, as unordered sets of functions, {eα,eβ}={H(f),L(f)}\{e_{\alpha},e_{\beta}\}=\{H(f),L(f)\}.

Remark 2.10.

We did not strictly need to introduce H(f)H(f) and L(f)L(f), nor their Bochner formulae. Indeed, we could have just defined DX(ρ)D_{X}(\rho) using eαe_{\alpha} and eβe_{\beta}. We prefer to use H(f)H(f) and L(f)L(f) and to show the equality with eαe_{\alpha} and eβe_{\beta} respectively because these are classical functions and carry geometric meaning.

Lemma 2.11.

Let ρ\rho be a reductive representation with eu(ρ)0\textrm{eu}(\rho)\geq 0 and carrying equivariant harmonic map ff, giving rise to data (cαβ,(β))(c\alpha\beta,(\beta)). Then the conjugacy class of ρ\rho is determined by (αβ,(β))(\alpha\beta,(\beta)).

Recall that the cc above is the constant chosen so that ϕ=cαβ\phi=c\alpha\beta. In making the statement we used that for eu(ρ)0\textrm{eu}(\rho)\geq 0, β\beta does not vanish identically. In the proof below, when unspecified, a product of sections is the tensor product.

Proof.

We say that (L,α,β)(L,\alpha,\beta) and (L,α,β)(L^{\prime},\alpha^{\prime},\beta^{\prime}) are isomorphic if the corresponding GG-Higgs bundles are isomorphic. When we can solve the self-duality equations, this is equivalent to the associated representations being conjugate. Thus, we only need to show that the isomorphism class of (L,α,β)(L,\alpha,\beta) is determined by (αβ,(β))(\alpha\beta,(\beta)).

Setting φ(L,α,β)=(ϕ,D)\varphi(L,\alpha,\beta)=(\phi,D), with ϕ=cαβ\phi=c\alpha\beta and D=(β)D=(\beta), the section β\beta defines a \mathbb{C}^{*}-family of isomorphisms from L1𝒦𝒪(D)L^{-1}\otimes\mathcal{K}\otimes\mathcal{O}(-D) to the trivial bundle 𝒪\mathcal{O}; each isomorphism takes β\beta to a constant in \mathbb{C}^{*}, and specifying that constant determines the isomorphism. We choose the isomorphism so that β\beta is sent to 11. There is an induced isomorphism from LL to 𝒦𝒪(D)\mathcal{K}\otimes\mathcal{O}(D). Under the dual isomorphism from L1L^{-1} to 𝒦1𝒪(D)\mathcal{K}^{-1}\otimes\mathcal{O}(D), since cαβ=ϕ,c\alpha\beta=\phi, α\alpha becomes c1ϕβ1.c^{-1}\phi\beta^{-1}.

With preliminaries established, we can now prove Proposition 2.7.

Proof of Proposition 2.7.

Let \mathcal{L} be the set of isomorphism classes of triples (L,α,β)(L,\alpha,\beta) such that we can solve (8) (uniquely) and let k\mathcal{L}_{k} be the subset of \mathcal{L} such that LL has degree kk. For k0k\geq 0, we define φk:k𝒟\varphi_{k}:\mathcal{L}_{k}\to\mathcal{D} by φk([L,α,β])=(ϕ,D),\varphi_{k}([L,\alpha,\beta])=(\phi,D), where ϕ=cαβ\phi=c\alpha\beta and DD is the divisor of the square root of the holomorphic energy of the harmonic map. For k>0k>0, this is just (β)(\beta), but for k=0k=0 we’re not quite able to distinguish. Each φk\varphi_{k} indeed lands in 𝒟\mathcal{D}: for k>0k>0, Proposition 2.9 gives H(f)=eβH(f)=e_{\beta}, and hence degD=degL1+2g2=2g2k\deg D=\deg L^{-1}+2g-2=2g-2-k. Still in the case k>0k>0, the condition D(ϕ)D\leq(\phi) is obvious. For k=0k=0, we can’t pick out whether H(f)=eβH(f)=e_{\beta} or eαe_{\alpha}, but the existence and uniqueness theory in this case shows that neither function vanishes identically, and as well using that degL=degL1=0,\deg L=\deg L^{-1}=0, we get that degD=2g2.\deg D=2g-2. We prove the proposition by showing that, for every k0k\geq 0, φk\varphi_{k} is a bijection onto the set {(ϕ,D)𝒟:degD=2g2k}.\{(\phi,D)\in\mathcal{D}:\deg D=2g-2-k\}.

By Lemma 2.11, each φk\varphi_{k} is injective. There is no issue for k=0k=0: none of α\alpha, β\beta, or ϕ\phi are zero sections, so we can express D=(β)D=(\beta) or D=(αβ)(β).D=(\alpha\beta)-(\beta). For surjectivity, fix a pair (ϕ,D)(\phi,D) with degD=2g2k\deg D=2g-2-k. Looking for a φk\varphi_{k}-preimage (L,α,β),(L,\alpha,\beta), we take L=𝒦1𝒪(D)L=\mathcal{K}^{-1}\otimes\mathcal{O}(D). Note that, via the inclusion 𝒪𝒪(D)\mathcal{O}\to\mathcal{O}(D), the constant section 11 of 𝒪\mathcal{O} determines a canonical section of 𝒪(D).\mathcal{O}(D). We take α\alpha to be this section and define β\beta by β=c1ϕα1.\beta=c^{-1}\phi\alpha^{-1}. Then φk([L,α,β])=(ϕ,D),\varphi_{k}([L,\alpha,\beta])=(\phi,D), as desired. ∎

3. Harmonic maps and domination

In this section, we prove Theorems A, C, and D, and Corollary C. Throughout, let XX be a closed Riemann surface of genus gg at least 22 and let ν\nu be a conformal metric on XX.

3.1. Domination inequality

Here we establish the key analytic input toward our main results, Proposition 3.2 below. Proposition 3.2, interesting in its own right, generalizes the most important inequality from [DT16], namely, [DT16, Lemma 2.6] (whose proof generalizes a classical argument as found in [SY97, Section 1.8]). See also [SG24, Lemma 4.3].

Let ϕ\phi be a holomorphic quadratic differential on XX.

Definition 3.1.

We say that a function uu on the complement of a discrete subset of XX is a Bochner solution for ϕ\phi if u=logH(f)u=\log H(f) or u=logL(f)u=\log L(f) for some equivariant harmonic map ff with Hopf differential ϕ\phi.

Equivalently, uu is a Bochner solution if there is a divisor DD satisfying certain conditions (for example, dominated by (ϕ)(\phi) if ϕ0\phi\neq 0) such that on the complement of the support of DD, uu is defined, C2C^{2}, and solves

12Δνu=eu|ϕ|ν2eu+κν,\frac{1}{2}\Delta_{\nu}u=e^{u}-|\phi|_{\nu}^{2}e^{-u}+\kappa_{\nu}, (9)

where |ϕ|ν2=|ϕ|2ν2|\phi|_{\nu}^{2}=|\phi|^{2}\nu^{-2}. Moreover, at a point pp in the support of DD, uu is asymptotic to D(p)log|z|D(p)\textrm{log}|z|.

Proposition 3.2.

Let u1u_{1} and u2u_{2} be Bochner solutions for ϕ\phi with divisors D1D_{1} and D2D_{2} respectively. If D2<D1D_{2}<D_{1}, then u1<u2u_{1}<u_{2} on the complement of the support of D2D_{2}.

Proof.

We first show u1u2u_{1}\leq u_{2}, and then we promote the result to u1<u2u_{1}<u_{2} outside of the support of D2D_{2}. Set u:=u1u2u:=u_{1}-u_{2}. By our assumptions, uu is bounded above and tends to -\infty on a non-empty discrete subset of XX. Assume for the sake of contradiction that u>0u>0 at a point. Then the open subset U={zX:u(z)>0}U=\{z\in X:u(z)>0\} is non-empty. Since uu tends to -\infty somewhere, UU is a proper open subset of XX. Taking the Laplacian of uu, (9) yields

12Δνu=(eu1eu2)|ϕ|ν2(eu1eu2)=eu2(eu1)eu2|ϕ|ν2(eu1).\frac{1}{2}\Delta_{\nu}u=(e^{u_{1}}-e^{u_{2}})-|\phi|_{\nu}^{2}(e^{-u_{1}}-e^{-u_{2}})=e^{u_{2}}(e^{u}-1)-e^{-u_{2}}|\phi|_{\nu}^{2}(e^{-u}-1). (10)

Since U\partial U is just the zero set of uu, uu is continuous on U¯\overline{U}. By (10), u|U¯u|_{\overline{U}} is subharmonic on UU. By the weak maximum principle, u|U¯u|_{\overline{U}} is maximized on U\partial U. But this contradicts u|U=0u|_{\partial U}=0. We deduce that u0u\leq 0.

We now prove that the inequality is strict. The strict inequality clearly holds near the support of D1D2D_{1}-D_{2}. On the complement of the support of D1D_{1}, which we will call VV, we have

12Δνu=(1+eu1u2|ϕ|ν2)(eu1eu2)=(eu2+eu1|ϕ|ν2)(eu1).\frac{1}{2}\Delta_{\nu}u=\Big(1+e^{-u_{1}-u_{2}}|\phi|_{\nu}^{2}\Big)(e^{u_{1}}-e^{u_{2}})=(e^{u_{2}}+e^{-u_{1}}|\phi|_{\nu}^{2})(e^{u}-1).

Since exx+1e^{x}\geq x+1 and u0u\leq 0,

ΔνuKu,\Delta_{\nu}u\geq Ku,

where K=2maxX(eu2+eu1|ϕ|ν2)).K=2\textrm{max}_{X}(e^{u_{2}}+e^{-u_{1}}|\phi|_{\nu}^{2})). Hence, by a consequence of the strong maximum principle [Min87], either u=0u=0 or u<0u<0 on all of VV. Since uu tends to -\infty as we approach the support of D1D2D_{1}-D_{2}, the former cannot occur. We conclude that u2<u1u_{2}<u_{1} on the set in question. ∎

Remark 3.3.

Proposition 3.2 extends easily to the case of a closed surface Σ\Sigma with compact boundary Σ\partial\Sigma. If u1u_{1} and u2u_{2} extend continuously to and agree on Σ\partial\Sigma, then the open subset UU from the proof above does not intersect Σ\partial\Sigma, and from this observation the proof goes through. This slight extension will be used in the proof of Theorem D.

3.2. Proof of Proposition 1.6

We now take a slight digression to prove Proposition 1.6. We include this proof for the sake of completeness, and because we will reference it in the proof of Lemma 3.8. As in the statement of the proposition, let f:X~(2,σ)f:\widetilde{X}\to(\mathbb{H}^{2},\sigma) be an equivariant harmonic map with holomorphic data (ϕ,D)(\phi,D), ϕ0\phi\neq 0.

Proof of Proposition 1.6.

For the “only if” direction, as in [BBDH21, Lemma 3.2], the sign of the Jacobian of a branched immersion does not flip. The divisor of the anti-holomorphic energy is (ϕ)D(\phi)-D, so if 2D(p)<(ϕ)(p)2D(p)<(\phi)(p), we have J(f)>0J(f)>0 near pp, and if 2D(p)>(ϕ)(p)2D(p)>(\phi)(p), we have J(f)<0J(f)<0 near pp. Thus, for a branched immersion, 2D(ϕ)2D-(\phi) does not flip sign.

For the main “if” direction, if 2D<(ϕ)2D<(\phi), we apply Proposition 3.2 with u1=logL(f)u_{1}=\log L(f) and u2=logH(f)u_{2}=\log H(f). Then, D1=(ϕ)DD_{1}=(\phi)-D and D2=DD_{2}=D, and hence our assumption shows that J(f)=H(f)L(f)>0J(f)=H(f)-L(f)>0 on the complement of the support of DD. By the argument from [BBDH21, pp. 12] (or using the Hartman-Wintner formula as in [Woo77]), the isolated singular points of ff are branch points. If 2D>(ϕ)2D>(\phi), we apply Proposition 3.2 with u1=logH(f)u_{1}=\log H(f) and u2=logL(f)u_{2}=\log L(f) and the argument is symmetric. ∎

3.3. Necessary and sufficient condition

We now move on to the domination problem. Proposition 3.4 below clarifies the role of HH and LL. Define

Uf={zX:H(f)(z)L(f)(z)},Vf={zX:H(f)(z)L(f)(z)}U_{f}=\{z\in X:H(f)(z)\geq L(f)(z)\},\hskip 2.84526ptV_{f}=\{z\in X:H(f)(z)\leq L(f)(z)\}

and

Uh={zX:H(h)(z)L(h)(z)}Vh={zX:H(h)(z)L(h)(z)}.U_{h}=\{z\in X:H(h)(z)\geq L(h)(z)\}\hskip 2.84526ptV_{h}=\{z\in X:H(h)(z)\leq L(h)(z)\}.

Note that UfVfU_{f}\cap V_{f} is the singular set of f,f, and UhVhU_{h}\cap V_{h} is the singular set of hh.

Proposition 3.4.

Assume that hh and ff have the same Hopf differential. For hσfσh^{*}\sigma\geq f^{*}\sigma to hold everywhere, it is necessary and sufficient that the following four conditions are satisfied.

  1. (1)

    On UfUh,U_{f}\cap U_{h}, H(h)H(f)H(h)\geq H(f).

  2. (2)

    On UfVh,U_{f}\cap V_{h}, L(h)H(f)L(h)\geq H(f).

  3. (3)

    On VfUh,V_{f}\cap U_{h}, H(h)L(f)H(h)\geq L(f).

  4. (4)

    On VfVhV_{f}\cap V_{h}, L(h)L(f)L(h)\geq L(f).

For hσ>fσh^{*}\sigma>f^{*}\sigma in a neighbourhood of a point xx, it is necessary and sufficient that (1) if xUfUhx\in U_{f}\cap U_{h}, H(h)(x)>H(f)(x),H(h)(x)>H(f)(x), (2) if xUfVh,x\in U_{f}\cap V_{h}, L(h)(x)>H(f)(x)L(h)(x)>H(f)(x), and similar for (3) and (4).

Here, fσhσf^{*}\sigma\leq h^{*}\sigma (as opposed to fσ<hσf^{*}\sigma<h^{*}\sigma) means that fσ(v,v)hσ(v,v)f^{*}\sigma(v,v)\leq h^{*}\sigma(v,v) for every unit tangent vector vv.

Proof.

From the formula (3), hσfσh^{*}\sigma\geq f^{*}\sigma if and only if e(h)e(f)e(h)\geq e(f). Rewriting e=H+Le=H+L as

e=(H1/2L1/2)2+2H1/2L1/2e=(H^{1/2}-L^{1/2})^{2}+2H^{1/2}L^{1/2} (11)

and recalling the formula |ϕ|ν2=H1/2L1/2,|\phi|_{\nu}^{2}=H^{1/2}L^{1/2}, the condition e(h)e(f)e(h)\geq e(f) is equivalent to demanding that

|H1/2(h)L1/2(h)||H1/2(f)L1/2(f)|.|H^{1/2}(h)-L^{1/2}(h)|\geq|H^{1/2}(f)-L^{1/2}(f)|. (12)

To make the notation easier on the eyes, note that (12) is equivalent to

|H(h)L(h)||H(f)L(f)|.|H(h)-L(h)|\geq|H(f)-L(f)|.

With this in mind, we check that |H(h)L(h)||H(f)L(f)||H(h)-L(h)|\geq|H(f)-L(f)| is necessary and sufficient. We explicitly write out the proof only for UfUhU_{f}\cap U_{h} and UfVhU_{f}\cap V_{h} and leave the rest to the reader, since the arguments for VfUhV_{f}\cap U_{h} and VfVhV_{f}\cap V_{h} are totally analogous and don’t add anything new.

On UfUh,U_{f}\cap U_{h}, assuming (1), we have H(h)H(f)L(f)H(h)\geq H(f)\geq L(f). From H(h)L(h)=H(f)L(f),H(h)L(h)=H(f)L(f), we get L(f)L(h).L(f)\geq L(h). Hence H(h)L(h)H(f)L(f)0H(h)-L(h)\geq H(f)-L(f)\geq 0, and taking absolute values yields the result. If (1) fails then we reverse the inequalities to see that H(f)L(f)H(h)L(h)0.H(f)-L(f)\geq H(h)-L(h)\geq 0.

On UfVh,U_{f}\cap V_{h}, assuming (2), L(h)H(f)L(f).L(h)\geq H(f)\geq L(f). Similar to above, H(h)L(h)=H(f)L(f)H(h)L(h)=H(f)L(f) implies that L(f)H(h),L(f)\geq H(h), and hence L(h)H(h)H(f)L(f)0.L(h)-H(h)\geq H(f)-L(f)\geq 0. If (2) does not hold, we get H(f)L(f)L(h)H(h)0.H(f)-L(f)\geq L(h)-H(h)\geq 0.

As we said above, we omit the arguments for VfUhV_{f}\cap U_{h} and VfVhV_{f}\cap V_{h}. The strictness statement is obtained by going back into the proof above and making the inequalities strict. ∎

3.4. Theorems A and C

Throughout this subsection, we use the sets of the form UfU_{f}, UhU_{h}, Vf,V_{f}, and VhV_{h} from Section 3.3. The main lemma is an application of Proposition 3.2.

Lemma 3.5.

Let f,h:X~(2,σ)f,h:\widetilde{X}\to(\mathbb{H}^{2},\sigma) be equivariant harmonic maps with holomorphic data (ϕ,D1)(\phi,D_{1}) and (ϕ,D2)(\phi,D_{2}) respectively. If D2<D1D_{2}<D_{1}, then H(f)H(h)H(f)\leq H(h), strictly away from the support of D2D_{2}. If D2<(ϕ)D1D_{2}<(\phi)-D_{1}, then L(f)H(h)L(f)\leq H(h), strictly away from the support of D2D_{2}.

Proof.

For the first statement, set u1=logH(f)u_{1}=\log H(f) and u2=logH(h)u_{2}=\log H(h), which are Bochner solutions with divisors D1D_{1} and D2D_{2} matching the divisors from the statement of the lemma. The result is immediate from Proposition 3.2. For the second statement, we go through the same results, but take u1=logL(f)u_{1}=\log L(f), which, as a Bochner solution, has divisor (ϕ)D.(\phi)-D.

We are now ready to prove Theorems A and C. We first prove Theorem C, then Corollary C, from which we deduce Theorem A.

Proof of Theorem C.

By our assumptions, Uh=XU_{h}=X. Assume that fσ<hσf^{*}\sigma<h^{*}\sigma. This rules out D1=D2D_{1}=D_{2} (for then we would have fσ=hσf^{*}\sigma=h^{*}\sigma everywhere). The equality D1+D2=(ϕ)D_{1}+D_{2}=(\phi) is ruled out too, since degD12g2\deg D_{1}\leq 2g-2, and hh being a branched immersion implies that degD2<2g2\deg D_{2}<2g-2. If D1(p)<D2(p)D_{1}(p)<D_{2}(p) at a point pXp\in X, then

limzpH(h)H(f)(z)=0.\lim_{z\to p}\frac{H(h)}{H(f)}(z)=0. (13)

As well, D1(p)<D2(p)<(ϕ)D2(p)<(ϕ)D1(p)D_{1}(p)<D_{2}(p)<(\phi)-D_{2}(p)<(\phi)-D_{1}(p), which implies that pUf=UfUhp\in U_{f}=U_{f}\cap U_{h}. Then, Proposition 3.4, specifically situation (1), contradicts fσhσf^{*}\sigma\leq h^{*}\sigma. Using similar reasoning, we will show that D1+D2<(ϕ)D_{1}+D_{2}<(\phi). Suppose that (D1+D2)(p)>(ϕ)(p)(D_{1}+D_{2})(p)>(\phi)(p) at a point pXp\in X. Then,

D1(p)D2(p)>((ϕ)D1)(p),D_{1}(p)\geq D_{2}(p)>((\phi)-D_{1})(p),

which implies pVf=UhVfp\in V_{f}=U_{h}\cap V_{f}. Analogous to (13),

limzpH(h)L(f)(z)=0,\lim_{z\to p}\frac{H(h)}{L(f)}(z)=0,

which via Proposition 3.4 implies that fσ>hσf^{*}\sigma>h^{*}\sigma near pp, and thus gives a contradiction. We conclude that D2<D1D_{2}<D_{1} and D1+D2<(ϕ).D_{1}+D_{2}<(\phi).

Conversely, assume that D2<D1D_{2}<D_{1} and that D1+D2<(ϕ).D_{1}+D_{2}<(\phi). We write X=UfVfX=U_{f}\cup V_{f}. The condition D2<D1D_{2}<D_{1}, together with Lemma 3.5 and Proposition 3.4, imply that fσ<hσf^{*}\sigma<h^{*}\sigma on UfU_{f}. The condition D1+D2<(ϕ)D_{1}+D_{2}<(\phi) is equivalent to D2<(ϕ)D1D_{2}<(\phi)-D_{1}, which via Lemma 3.5 and Proposition 3.4 implies that fσ<hσf^{*}\sigma<h^{*}\sigma on VfV_{f}. Hence, fσ<hσf^{*}\sigma<h^{*}\sigma everywhere. ∎

Proof of Corollary C.

For both implications, we can assume that hh is a branched immersion, since this is implied by both D2<D1D_{2}<D_{1} (from Proposition 1.6) and fσ<hσf^{*}\sigma<h^{*}\sigma. Since deg(D1),deg(D2)2g2\deg(D_{1}),\deg(D_{2})\leq 2g-2, Proposition 1.6 implies 2D1<(ϕ)2D_{1}<(\phi) and 2D2<(ϕ)2D_{2}<(\phi). Hence, the condition D1+D2<(ϕ)D_{1}+D_{2}<(\phi) is automatic. The corollary is then immediate from Theorem C. ∎

Proof of Theorem A.

Let ρ\rho, ff, and kk be as in the statement of the theorem and let (ϕ,D1)(\phi,D_{1}) be the holomorphic data of ff. If ϕ0\phi\neq 0, we pick any divisor D2D_{2} of degree degD1k\deg D_{1}-k such that D2D1D_{2}\leq D_{1}, and we apply Proposition 2.7 and Corollary C part (1) to produce the desired representation jj and equivariant harmonic map hh. If ϕ=0\phi=0, we pick D2D_{2} of degree degD1k\deg D_{1}-k and dominated by both D1D_{1} and the divisor of a non-zero quadratic differential, and we apply Proposition 2.7 and Corollary C as above. ∎

We conclude this subsection by showing that an assumption such as hh is a branched harmonic immersion” is necessary. This result shows that the domination problem can become delicate.

Proposition 3.6.

For ϕ0\phi\neq 0, consider equivariant harmonic maps hh and ff with holomorphic data (ϕ,D1)(\phi,D_{1}) and (ϕ,D2)(\phi,D_{2}) respectively. Assume that D2<D1D_{2}<D_{1}. If there exist distinct points pp and qq at which 2D1(p)<(ϕ)(p)2D_{1}(p)<(\phi)(p) and 2D2(q)>(ϕ)(q)2D_{2}(q)>(\phi)(q) respectively, then neither fσhσf^{*}\sigma\leq h^{*}\sigma nor hσfσh^{*}\sigma\leq f^{*}\sigma hold everywhere.

Proof.

By Lemma 3.5, H(f)H(h)H(f)\leq H(h) everywhere, strictly away from the support of D2D_{2}. Since H(h)L(h)=H(f)L(f)H(h)L(h)=H(f)L(f), we obtain

J(h)=H(h)L(h)H(f)L(f)=J(f)J(h)=H(h)-L(h)\geq H(f)-L(f)=J(f)

everywhere, with the same strictness condition. We deduce that

J(h)0J(f)0,J(f)0J(h)0.J(h)\leq 0\Rightarrow J(f)\leq 0,\hskip 2.84526ptJ(f)\geq 0\Rightarrow J(h)\geq 0. (14)

By assumption, the open subset UfU_{f} is non-empty (it contains the point pp). Thus, by (14), UfUhU_{f}\subset U_{h}, and by Proposition 3.4 case (1), hσfσh^{*}\sigma\leq f^{*}\sigma fails around pp. On the other hand, the presence of the point qq shows that VhV_{h} is non-empty, and (14) shows that VhVfV_{h}\subset V_{f}. By Proposition 3.4 case (4), hσfσh^{*}\sigma\leq f^{*}\sigma does not hold around qq. ∎

3.5. Theorem D

Both types of examples from Theorem D come from variations on the same construction. Let X1X_{1} be a closed Riemann surface with one boundary component γ\gamma and form the Riemann surface double YY, which comes with an antiholomorphic involution τ\tau whose set of fixed points is γ\gamma. Necessarily, YY has even genus, and of course any even genus smooth surface can be seen to arise from such a doubling. We attach a conformal metric to YY so that we can define functions such as H()H(\cdot) and L()L(\cdot). Let ϕ\phi be a non-zero holomorphic quadratic differential with no zeros on γ\gamma and that is symmetric with respect to τ\tau, i.e, such that τϕ=ϕ¯\tau^{*}\phi=\overline{\phi}. We view X1X_{1} and X2:=τ(X1)X_{2}:=\tau(X_{1}) as subsets of YY. We choose a first divisor DD^{\prime} on X1X_{1} dominated by (ϕ)|X1(\phi)|_{X_{1}}. If D′′=(ϕ)|X2D^{\prime\prime}=(\phi)|_{X_{2}}, we define D=D+(D′′τ(D))D=D^{\prime}+(D^{\prime\prime}-\tau(D^{\prime})). We point out that for such pairs, degD=deg(ϕ)2=2g2\deg D=\frac{\deg(\phi)}{2}=2g-2, and hence the representation corresponding to (ϕ,D)(\phi,D) has Euler number 0 (recall Proposition 2.5). Let τ~\widetilde{\tau} be the lift of τ\tau to the universal cover Y~\widetilde{Y}.

For this YY and (ϕ,D)(\phi,D) as above, let ρ:ΓPSL(2,)\rho:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) be the corresponding representation and let f:Y~(2,σ)f:\widetilde{Y}\to(\mathbb{H}^{2},\sigma) be a ρ\rho-equivariant harmonic map.

Lemma 3.7.

fτ~=ff\circ\widetilde{\tau}=f and J(f)=0J(f)=0 on γ.\gamma.

Proof.

In a local coordinate zz, since τ\tau is anti-holomorphic, (fτ~)z=(τ~¯)zfz¯(τ(z)).(f\circ\widetilde{\tau})_{z}=(\overline{\widetilde{\tau}})_{z}f_{\overline{z}}(\tau(z)). It follows using the definitions stemming from (3) that ϕ(fτ~)=τ~ϕ¯(f)=ϕ,\phi(f\circ\widetilde{\tau})=\widetilde{\tau}^{*}\overline{\phi}(f)=\phi, and H(fτ~)=L(f)τ.H(f\circ\widetilde{\tau})=L(f)\circ\tau. The second equality implies that the divisor associated with fτ~f\circ\widetilde{\tau} is (ϕ)Dτ=D(\phi)-D\circ\tau=D. Thus, ff and fτ~f\circ\widetilde{\tau} have the same holomorphic data. By Proposition 2.7, if the representation associated with (ϕ,D)(\phi,D) is irreducible, then f=fτ~f=f\circ\widetilde{\tau}, and if the representation is reductive but not irreducible, then ff and fτ~f\circ\widetilde{\tau} are related by translation along a geodesic. Even in the latter case, since τ\tau fixes γ\gamma, we can conclude that the two maps agree. The fact that J(f)=0J(f)=0 on γ\gamma follows from the equality H(fτ~)=L(f)τ.H(f\circ\widetilde{\tau})=L(f)\circ\tau.

Let X~1Y~\widetilde{X}_{1}\subset\widetilde{Y} be the preimage of X1X_{1} under the universal covering map. We record an analog of Proposition 1.6 for maps restricted to X~1\widetilde{X}_{1}.

Lemma 3.8.

f|int(X~1)f|_{\textrm{int}(\widetilde{X}_{1})} is a branched immersion if and only if 2D<(ϕ|X1)2D^{\prime}<(\phi|_{X_{1}}) or 2D>(ϕ|X1)2D^{\prime}>(\phi|_{X_{1}}).

Proof.

The same argument from the proof of Proposition 1.6 implies that 2D<(ϕ|X1)2D^{\prime}<(\phi|_{X_{1}}) or 2D>(ϕ|X1)2D^{\prime}>(\phi|_{X_{1}}) are necessary. For the converse, as in the proof of Proposition 1.6, we apply Proposition 3.2 and Remark 3.3 with u1=logL(f)|X1u_{1}=\log L(f)|_{X_{1}} and u2=logH(f)|X1u_{2}=\log H(f)|_{X_{1}} if 2D<(ϕ|X1)2D^{\prime}<(\phi|_{X_{1}}), and u1=logH(f)|X1u_{1}=\log H(f)|_{X_{1}} and u2=logL(f)|X1u_{2}=\log L(f)|_{X_{1}} if 2D>(ϕ|X1)2D^{\prime}>(\phi|_{X_{1}}). By Lemma 3.7, u1=u2u_{1}=u_{2} extend to and agree on X1=γ\partial X_{1}=\gamma, so Remark 3.3 indeed applies. ∎

Similar to above, we provide an analog for Corollary C. Let (ϕ,D1)(\phi,D_{1}) and (ϕ,D2)(\phi,D_{2}) be constructed as above with equivariant harmonic maps ff and hh respectively. For i=1,2i=1,2, we let DiD_{i}^{\prime} and Di′′D_{i}^{\prime\prime} be the divisors corresponding to DD^{\prime} and D′′D^{\prime\prime} respectively.

Lemma 3.9.

Assume that 2D1<(ϕ|X1)2D_{1}^{\prime}<(\phi|_{X_{1}}). Then fσ<hσf^{*}\sigma<h^{*}\sigma if and only if D2<D1D_{2}^{\prime}<D_{1}^{\prime} or D2>(ϕ|X1)D1D_{2}^{\prime}>(\phi|_{X_{1}})-D_{1}^{\prime}.

Proof.

By using an outer automorphism of the fundamental group, we can assume that degD2|X1g1\deg D_{2}|_{X_{1}}\leq g-1, and under this assumption the case D2>(ϕ|X1)D1D_{2}^{\prime}>(\phi|_{X_{1}})-D_{1} is removed. By Lemma 3.7, fσ|X2=τ(fσ|X1)f^{*}\sigma|_{X_{2}}=\tau^{*}(f^{*}\sigma|_{X_{1}}) and τ(hσ|X1)=hσ|X2\tau^{*}(h^{*}\sigma|_{X_{1}})=h^{*}\sigma|_{X_{2}}, so fσhσf^{*}\sigma\leq h^{*}\sigma on X1X_{1} if and only if fσhσf^{*}\sigma\leq h^{*}\sigma everywhere. By Lemma 3.8, f|int(X~1)f|_{\textrm{int}(\widetilde{X}_{1})} is a branched immersion.

The exact same local analysis from the proof of Theorem C shows that D2<D1D_{2}^{\prime}<D_{1}^{\prime} is necessary for the domination. Assume that D2<D1D_{2}^{\prime}<D_{1}^{\prime}. By Lemma 3.7, both J(f)J(f) and J(h)J(h) vanish on γ\gamma, and from H(f)L(f)=H(h)L(h),H(f)L(f)=H(h)L(h), we obtain that H(f)|γ=H(h)|γ.H(f)|_{\gamma}=H(h)|_{\gamma}. Setting u1=logH(f)|X1u_{1}=\log H(f)|_{X_{1}} and u2=logH(h)|X1u_{2}=\log H(h)|_{X_{1}}, we apply Proposition 3.2 and Remark 3.3 to obtain H(f)H(h)H(f)\leq H(h) on X1X_{1} (analogous to Lemma 3.5), and the inequality is strict when J(h)0J(h)\neq 0. By the argument from Theorem C, fσ<hσf^{*}\sigma<h^{*}\sigma on X1X_{1}. By symmetry, the same is true on X2X_{2}. ∎

With preparations complete, we now prove Theorem D.

Proof of Theorem D.

We work with the construction and notations as above. For (1), select pairs (ρ,f)(\rho,f) and (j,h)(j,h) giving rise to data (ϕ,D2)(\phi,D_{2}) and (ϕ,D1)(\phi,D_{1}) respectively with 2D1<(ϕ|X1)2D_{1}^{\prime}<(\phi|_{X_{1}}) and D2<D1D_{2}^{\prime}<D_{1}^{\prime}, exactly as in Lemma 3.9.

For (2), we consider (ϕ,D0)(\phi,D_{0}) as above and, for D0=D0+(D0′′τ(D0))D_{0}=D_{0}^{\prime}+(D_{0}^{\prime\prime}-\tau(D_{0}^{\prime})) as above, we specify that D0=0D_{0}^{\prime}=0. Let (ρ,f)(\rho,f) be the corresponding representation and equivariant harmonic map. By Lemmas 3.7 and 3.8, f|int(X~1)f|_{\textrm{int}(\widetilde{X}_{1})} and f|int(X~2)f|_{\textrm{int}(\widetilde{X}_{2})} are branched immersions, and ff is singular on the shared frontier of X~1\widetilde{X}_{1} and X~2\widetilde{X}_{2}. In fact, since D0=0D_{0}^{\prime}=0 (and D0′′=(ϕ)|X2)D_{0}^{\prime\prime}=(\phi)|_{X_{2}}), f|int(X~1)f|_{\textrm{int}(\widetilde{X}_{1})} and f|int(X~2)f|_{\textrm{int}(\widetilde{X}_{2})} are immersions.

Let (j,h)(j,h) be any other pair with holomorphic data (ϕ,D)(\phi,D) and assume that fσ<hσf^{*}\sigma<h^{*}\sigma. Since fσf^{*}\sigma is non-degenerate on int(X1)\textrm{int}(X_{1}) and int(X2)\textrm{int}(X_{2}), the singular set of hh is contained in γ\gamma, and h|int(X~1)h|_{\textrm{int}(\widetilde{X}_{1})} and h|int(X~2)h|_{\textrm{int}(\widetilde{X}_{2})} are immersions. By applying an outer automorphism to jj (which will not alter hσh^{*}\sigma), we can assume that J(h)|X10J(h)|_{X_{1}}\geq 0. Then D|X1D|_{X_{1}} is contained in the singular set of h|int(X~1)h|_{\textrm{int}(\widetilde{X}_{1})}, and the latter being empty forces D|X1=0D|_{X_{1}}=0. Now, we have that either J(h)|X20J(h)|_{X_{2}}\geq 0 or J(h)|X20J(h)|_{X_{2}}\leq 0. If J(h)|X20J(h)|_{X_{2}}\geq 0, the non-degeneracy of hσh^{*}\sigma on int(X2)\textrm{int}(X_{2}) shows that D|X2=0D|_{X_{2}}=0, which forces jj to be Fuchsian. Similarly, if J(h)0J(h)\leq 0, then the divisor of the square root of the anti-holomorphic energy on X2X_{2}, which is by definition (ϕ|X2)D|X2(\phi|_{X_{2}})-D|_{X_{2}}, is contained in the singular set of h|int(X~2)h|_{\mathrm{int}(\widetilde{X}_{2})}, and thus (ϕ|X2)=D|X2(\phi|_{X_{2}})=D|_{X_{2}}. We therefore deduce by Proposition 2.7 that ρ\rho and jj are conjugate. But then fσ=hσf^{*}\sigma=h^{*}\sigma, and we have a contradiction. ∎

4. Anti-de Sitter 33-manifolds

In this section, we introduce the geometry of the anti-de Sitter space and the local model of a spin-cone singular anti-de Sitter manifold.

4.1. Anti-de Sitter space

Let 𝔰𝔩(2,)\mathfrak{sl}(2,\mathbb{R}) be the Lie algebra of PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}). On 𝔰𝔩(2,)\mathfrak{sl}(2,\mathbb{R}), we consider the bilinear form det-\mathrm{det}, which is invariant under the adjoint representation. It can be shown that det-\mathrm{det} coincides with 18\frac{1}{8} times the Killing form of PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}). This form has signature (2,1)(2,1) and induces a bi-invariant Lorentzian metric on PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), which we denote by g𝔸d𝕊3g^{\mathbb{A}\mathrm{d}\mathbb{S}^{3}}. We define the 33-dimensional anti-de Sitter space to be:

𝔸d𝕊3:=(PSL(2,),g𝔸d𝕊3).\mathbb{A}\mathrm{d}\mathbb{S}^{3}:=\left(\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}),g^{\mathbb{A}\mathrm{d}\mathbb{S}^{3}}\right).

The group of orientation- and time-orientation–preserving isometries is PSL(2,)×PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\times\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), acting on 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} by left and right multiplication. That is,

(A,B)X=AXB1.(A,B)\cdot X=AXB^{-1}.

A tangent vector vTx𝔸d𝕊3v\in\mathrm{T}_{x}\mathbb{A}\mathrm{d}\mathbb{S}^{3} is timelike if g𝔸d𝕊3(v,v)<0g^{\mathbb{A}\mathrm{d}\mathbb{S}^{3}}(v,v)<0, lightlike if g𝔸d𝕊3(v,v)=0g^{\mathbb{A}\mathrm{d}\mathbb{S}^{3}}(v,v)=0, and spacelike if g𝔸d𝕊3(v,v)>0g^{\mathbb{A}\mathrm{d}\mathbb{S}^{3}}(v,v)>0. We call a geodesic γ\gamma timelike if every tangent vector is timelike, lightlike if every tangent vector is lightlike, and spacelike if every tangent vector is spacelike.

It turns out that every timelike geodesic is of the form

p,q={APSL(2,)Aq=p},\ell_{p,q}=\left\{A\in\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\mid A\cdot q=p\right\}, (15)

for (p,q)2×2(p,q)\in\mathbb{H}^{2}\times\mathbb{H}^{2}. These are topological circles and have Lorentzian length π\pi. Note that under this identification, it can be checked that for any isometry (A,B)(A,B) of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}, we have

(A,B)p,q=Ap,Bq.(A,B)\cdot\ell_{p,q}=\ell_{Ap,Bq}.

Hence, the 1-to-1 correspondence (p,q)p,q(p,q)\to\ell_{p,q} is equivariant with respect to the action of PSL(2,)×PSL(2,)\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\times\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) on 2×2\mathbb{H}^{2}\times\mathbb{H}^{2} and on the set of timelike geodesics.

We end this subsection by briefly recalling the notion of a geometric structure on a 33-manifold MM. Let XX be a manifold and GG a Lie group acting transitively on XX by analytic diffeomorphisms. Then a (G,X)(G,X)-structure on MM is a maximal atlas of coordinate charts on MM with values in XX such that the transition maps are given by elements of GG. An important result from the theory is that MM is equipped with a holonomy representation ρ:π1(M)G\rho:\pi_{1}(M)\to G and a ρ\rho-equivariant local diffeomorphism dev:M~X\mathrm{dev}:\widetilde{M}\to X, called the developing map. The pair (hol,dev)(\mathrm{hol},\mathrm{dev}) is defined up to the action of GG, where GG acts by conjugation on the holonomy representation and post-composition on the developing map. In this paper, we focus on the case where XX is the three-dimensional anti-de Sitter space 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} and G=Isom0(𝔸d𝕊3)G=\mathrm{Isom}_{0}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}). The corresponding (G,X)(G,X)-structures on MM are known as anti-de Sitter structures, and a manifold endowed with such a structure is called an anti-de Sitter manifold. For further details on three-dimensional anti-de Sitter geometry, we refer the reader to [BS20].

4.2. AdS manifolds with spin-cone structure

Before getting to spin-cone singularities, we recall the model of cone singularities in 22-dimensional hyperbolic geometry.

4.2.1. Hyperbolic cone singularities

Let 2:=2{i}\mathbb{H}^{2}_{*}:=\mathbb{H}^{2}\setminus\{i\}, and let c:[0,+)2c:[0,+\infty)\to\mathbb{H}^{2} be the geodesic parametrized by hyperbolic arc length given by

c(t)=iet,c(t)=ie^{t},

so that c(0)=ic(0)=i and limt+c(t)=\lim_{t\to+\infty}c(t)=\infty in the boundary at infinity of 2\mathbb{H}^{2}. The universal cover 2~\widetilde{\mathbb{H}^{2}_{*}} of 2\mathbb{H}^{2}_{*} is given by

π:(0,+)×2(r,θ)Rθc(r).\begin{array}[]{rcl}\pi&:&(0,+\infty)\times\mathbb{R}\to\mathbb{H}^{2}_{*}\\ &&(r,\theta)\mapsto R^{\theta}c(r).\end{array} (16)

The group of deck transformations of π\pi, which is isomorphic to the fundamental group of 2\mathbb{H}^{2}_{*}, is given by

π1(2)={(r,θ)(r,θ+2kπ)k}.\pi_{1}(\mathbb{H}^{2}_{*})=\left\{(r,\theta)\mapsto(r,\theta+2k\pi)\mid k\in\mathbb{Z}\right\}.

We endow 2~\widetilde{\mathbb{H}^{2}_{*}} with the Riemannian metric obtained by pulling back the hyperbolic metric σ\sigma via π\pi. In this model, the isometry group satisfies Isom(2~)\mathrm{Isom}(\widetilde{\mathbb{H}^{2}_{*}})\cong\mathbb{R}. More precisely,

Isom(2~)={(r,θ)(r,θ+τ)τ}.\mathrm{Isom}(\widetilde{\mathbb{H}^{2}_{*}})=\left\{(r,\theta)\mapsto(r,\theta+\tau)\mid\tau\in\mathbb{R}\right\}.

For θ0\theta_{0}\in\mathbb{R}, we define the local model of a hyperbolic cone structure by

θ02:=2~/(r,θ)(r,θ+θ0).\mathbb{H}^{2}_{\theta_{0}}:=\widetilde{\mathbb{H}^{2}_{*}}/(r,\theta)\sim(r,\theta+\theta_{0}). (17)

We call the point ii the branched point of 2\mathbb{H}^{2}_{*}. To justify this terminology, we define the universal branched cover of 2\mathbb{H}^{2}_{*} as the quotient:

([0,+)×){0}:=([0,+)×)/,\left([0,+\infty)\times\mathbb{R}\right)\sqcup\{0\}:=\left([0,+\infty)\times\mathbb{R}\right)/\sim, (18)

where (0,θ)(0,θ)(0,\theta)\sim(0,\theta^{\prime}) for all θ,θ\theta,\theta^{\prime}\in\mathbb{R}, identifying all points at r=0r=0 to a single point, denoted {0}\{0\}. The universal covering map π:(0,+)×2\pi:(0,+\infty)\times\mathbb{R}\to\mathbb{H}^{2}_{*} extends to this space by setting π([0,θ])=i\pi([0,\theta])=i for all θ\theta. This is well defined since c(0)=ic(0)=i and Rθi=iR^{\theta}i=i. Thus, π(0)=i\pi(0)=i, making ii the branch point of the covering π\pi.

As a particular case, it is worth observing that when θ0=2nπ2π\theta_{0}=2n\pi\in 2\pi\mathbb{Z}, the surface 2nπ2\mathbb{H}^{2}_{2n\pi} is a degree-nn cover of 2\mathbb{H}^{2}_{*}. Indeed, the universal covering map π:2~2\pi:\widetilde{\mathbb{H}^{2}_{*}}\to\mathbb{H}^{2}_{*} induces the degree-nn cover

π¯:2nπ22,[r,θ]nπ(r,θ),\overline{\pi}:\mathbb{H}^{2}_{2n\pi}\to\mathbb{H}^{2}_{*},\quad[r,\theta]_{n}\mapsto\pi(r,\theta),

where [r,θ]n[r,\theta]_{n} denotes the class of (r,θ)(r,\theta) in 2nπ2\mathbb{H}^{2}_{2n\pi}. We also note that 2nπ2\mathbb{H}^{2}_{2n\pi} can be identified with 2\mathbb{H}^{2}_{*} via the map

2nπ22[R,θ]nRθnc(r).\begin{array}[]{ccccc}&&\mathbb{H}^{2}_{2n\pi}&\to&\mathbb{H}^{2}_{*}\\ &&[R,\theta]_{n}&\mapsto&R^{\frac{\theta}{n}}c(r).\end{array} (19)

Therefore, it will be useful to view 2nπ2\mathbb{H}^{2}_{2n\pi} as the punctured hyperbolic plane 2\mathbb{H}^{2}_{*} equipped with the degree-nn covering map

2nπ222Rθnc(r)Rθc(r).\begin{array}[]{ccccc}&&\mathbb{H}^{2}_{2n\pi}\cong\mathbb{H}^{2}_{*}&\to&\mathbb{H}^{2}_{*}\\ &&R^{\frac{\theta}{n}}c(r)&\mapsto&R^{\theta}c(r).\end{array}

In the disc model, this map corresponds to zznz\mapsto z^{n}.

4.2.2. Spin-cone structure

In what follows, for each r,θr,\theta\in\mathbb{R}, we consider the following isometries of the hyperbolic Poincaré half plane:

A(r)=(er200er2),Rθ=(cos(θ2)sin(θ2)sin(θ2)cos(θ2)).A(r)=\begin{pmatrix}e^{\frac{r}{2}}&0\\ 0&e^{-\frac{r}{2}}\end{pmatrix},\quad R^{\theta}=\begin{pmatrix}\cos\left(\frac{\theta}{2}\right)&\sin\left(\frac{\theta}{2}\right)\\ -\sin\left(\frac{\theta}{2}\right)&\cos\left(\frac{\theta}{2}\right)\end{pmatrix}.

The matrix A(r)A(r) acts as a translation of length rr along the geodesic in 2\mathbb{H}^{2} with endpoints 0 and \infty, whereas RθR^{\theta} acts as a rotation of angle θ\theta fixing the point i2i\in\mathbb{H}^{2} (alternatively, one may view it as a rotation fixing the origin in the Poincaré disc model). Let i,i\ell_{i,i} denote the set of elliptic isometries fixing i2i\in\mathbb{H}^{2}, see (15). Observe that the curve [0,π]𝔸d𝕊3[0,\pi]\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}, given by θR2θ\theta\mapsto R^{2\theta}, is the arc-length parametrization of i,i\ell_{i,i}. Hence, the Lorentzian length of i,i\ell_{i,i} is equal to π\pi, that is,

0πdet((R2θ))𝑑θ=π.\int_{0}^{\pi}\sqrt{-\det((R^{2\theta})^{\prime})}\,d\theta=\pi. (20)

Now, consider the space 𝔸d𝕊3:=PSL(2,)i,i\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}:=\mathrm{PSL}(2,\mathbb{R})\setminus\ell_{i,i}. In [Jan22, Proposition 3.5.1], Janigro defines the universal cover of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} by the map

T:(0,+)××𝔸d𝕊3(r,θ,t)RθA(r)Rt.\begin{array}[]{cccc}T&:&(0,+\infty)\times\mathbb{R}\times\mathbb{R}&\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}\\ &&(r,\theta,t)&\mapsto R^{\theta}A(r)R^{-t}.\end{array} (21)

The group of deck transformations of TT, which is isomorphic to the fundamental group of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}, is given by

π1(𝔸d𝕊3)={(r,θ,η)(r,θ+2k1π,η+2k2π)k1,k2}.\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*})=\{(r,\theta,\eta)\mapsto(r,\theta+2k_{1}\pi,\eta+2k_{2}\pi)\mid k_{1},k_{2}\in\mathbb{Z}\}.

We endow 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}} with the Lorentzian metric obtained by pulling back the anti-de Sitter metric g𝔸d𝕊3g^{\mathbb{A}\mathrm{d}\mathbb{S}^{3}} via TT. In this model, Isom(𝔸d𝕊3~)2\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}})\cong\mathbb{R}^{2}. More precisely, we have

Isom(𝔸d𝕊3~)={φ(θ0,η0):(r,θ,η)(r,θ+θ0,η+η0)θ0,η0}.\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}})=\{\varphi_{(\theta_{0},\eta_{0})}:(r,\theta,\eta)\mapsto(r,\theta+\theta_{0},\eta+\eta_{0})\mid\theta_{0},\eta_{0}\in\mathbb{R}\}. (22)

Note that the isometry group of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} identifies with isometries of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} that fix the timelike geodesic i,i\ell_{i,i}, that is,

Isom(𝔸d𝕊3)=PSO(2)×PSO(2)<PSL(2,)×PSL(2,).\mathrm{Isom}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*})=\mathrm{PSO}(2)\times\mathrm{PSO}(2)<\mathrm{PSL}(2,\mathbb{R})\times\mathrm{PSL}(2,\mathbb{R}).

The map TT induces a homomorphism T:Isom(𝔸d𝕊3~)Isom(𝔸d𝕊3)T_{*}:\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}})\to\mathrm{Isom}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}). If φIsom(𝔸d𝕊3~)\varphi\in\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}), then T(φ)T_{*}(\varphi) satisfies:

T(φ)T=Tφ.T_{*}(\varphi)\circ T=T\circ\varphi. (23)

The kernel of TT_{*} is the group of deck transformations of the covering TT, given by

Ker(T)={(r,θ,η)(r,θ,η)+k1(0,0,2π)+k2(0,2π,0)k1,k2}.\mathrm{Ker}(T_{*})=\{(r,\theta,\eta)\mapsto(r,\theta,\eta)+k_{1}(0,0,2\pi)+k_{2}(0,2\pi,0)\mid k_{1},k_{2}\in\mathbb{Z}\}. (24)
Remark 4.1.

If we take \ell to be another timelike geodesic in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}, then there exists an isometry of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} sending i,i\ell_{i,i} to \ell. Hence, there is an isometry between 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}} and 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}\setminus\ell}, and we still denote the isometries of 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}\setminus\ell} by φ(θ0,η0)\varphi_{(\theta_{0},\eta_{0})} as in (22).

For θ0,η0\theta_{0},\eta_{0}\in\mathbb{R}, we define Λ(θ0,η0)\Lambda(\theta_{0},\eta_{0}) as the lattice in Isom(𝔸d𝕊3~)\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}) generated by (0,θ0,η0)(0,\theta_{0},\eta_{0}) and (0,0,2π)(0,0,2\pi). That is,

Λ(θ0,η0)={(r,θ,η)(r,θ,η)+k1(0,0,2π)+k2(0,θ0,η0)k1,k2}.\Lambda(\theta_{0},\eta_{0})=\{(r,\theta,\eta)\mapsto(r,\theta,\eta)+k_{1}(0,0,2\pi)+k_{2}(0,\theta_{0},\eta_{0})\mid k_{1},k_{2}\in\mathbb{Z}\}. (25)

Observe that Λ(θ0,η0)=Λ(θ0,η0+2kπ)\Lambda(\theta_{0},\eta_{0})=\Lambda(\theta_{0},\eta_{0}+2k\pi) for any kk\in\mathbb{Z}. Hence, η0\eta_{0} may be regarded as an element of /2π\mathbb{R}/2\pi\mathbb{Z}. Without loss of generality, we may assume 0η0<2π0\leq\eta_{0}<2\pi. Following [Jan22], we give the definition below.

Definition 4.2.

Let θ0\theta_{0}\in\mathbb{R} and 0η0<2π0\leq\eta_{0}<2\pi. We define the local model of a spin-cone singularity as

𝔸d𝕊(θ0,η0)3:=𝔸d𝕊3~/Λ(θ0,η0).\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(\theta_{0},\eta_{0})}:=\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}/\Lambda(\theta_{0},\eta_{0}).

The next lemma can be viewed as a generalization, in the singular setting, of the fact that the anti-de Sitter space 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} is a circle bundle over 2\mathbb{H}^{2}, with fibers being timelike geodesics of length π\pi.

Lemma 4.3.

Let θ0\theta_{0}\in\mathbb{R} be different from zero, and 0η0<2π0\leq\eta_{0}<2\pi. Consider the projection map (r,θ,η)(r,θ)(r,\theta,\eta)\mapsto(r,\theta), which induces a map :𝔸d𝕊(θ0,η0)3θ02\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(\theta_{0},\eta_{0})}\to\mathbb{H}^{2}_{\theta_{0}}. Then, \mathcal{F} is a fibration with the property that each fiber is a timelike geodesic of length π\pi.

Proof.

Consider the projection map (r,θ,η)(r,θ)(r,\theta,\eta)\mapsto(r,\theta), which induces a map :𝔸d𝕊(θ0,η0)3θ02\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(\theta_{0},\eta_{0})}\to\mathbb{H}^{2}_{\theta_{0}}. The fiber above [r,θ][r,\theta] is given by

1([r,θ])={[r,θ,η]η}.\mathcal{F}^{-1}([r,\theta])=\{[r,\theta,\eta]\mid\eta\in\mathbb{R}\}.

We claim that 1([r,θ])\mathcal{F}^{-1}([r,\theta]) is a timelike geodesic with arc-length parametrization

c(t)=[r,θ,2t],t[0,π).c(t)=[r,\theta,2t],\quad t\in[0,\pi).

Let t1,t2[0,π)t_{1},t_{2}\in[0,\pi) be such that [r,θ,2t1]=[r,θ,2t2][r,\theta,2t_{1}]=[r,\theta,2t_{2}]. Then there exist n,mn,m\in\mathbb{Z} such that

2t1=2t2+2nπ+mη0,andmθ0=0.2t_{1}=2t_{2}+2n\pi+m\eta_{0},\quad\text{and}\quad m\theta_{0}=0.

Since θ00\theta_{0}\neq 0, it follows that m=0m=0 and hence 2t1=2t2+2nπ2t_{1}=2t_{2}+2n\pi. Since t1,t2[0,π)t_{1},t_{2}\in[0,\pi), this implies n=0n=0 and t1=t2t_{1}=t_{2}, so cc is injective.

Next, the curve

β(t)=R2tA(r0)\beta(t)=R^{2t}A(r_{0})

parametrizes a timelike geodesic of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} by arc length. Hence, the fiber 1([r,θ])\mathcal{F}^{-1}([r,\theta]) is a timelike geodesic of length π\pi. This completes the proof. ∎

We now consider the special case θ0=2nπ\theta_{0}=2n\pi for some nn\in\mathbb{N} and η0=0\eta_{0}=0, which will be our main focus. First of all, it can be shown without difficulties that the universal covering map T:(0,+)××𝔸d𝕊3T:(0,+\infty)\times\mathbb{R}\times\mathbb{R}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} induces a diffeomorphism 𝔸d𝕊(2π,0)3𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(2\pi,0)}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}, which we continue to denote by TT (see [Jan22, Proposition 3.5.1]). For each integer nn, we denote by λn\lambda_{n} the map from (0,+)××(0,+)××(0,+\infty)\times\mathbb{R}\times\mathbb{R}\to(0,+\infty)\times\mathbb{R}\times\mathbb{R} that sends (r,θ,η)(r,\theta,\eta) to (r,θn,η)(r,\frac{\theta}{n},\eta). According to the previous result, it is straightforward to check that TλnT\circ\lambda_{n} induces a diffeomorphism 𝔸d𝕊(2nπ,0)3𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(2n\pi,0)}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}, for which we keep the notation TλnT\circ\lambda_{n}. We denote by [r,θ,η]n[r,\theta,\eta]_{n} the equivalence class of (r,θ,η)(r,\theta,\eta) under the action of Λ(2nπ,0)\Lambda(2n\pi,0). We observe that

𝒞n:𝔸d𝕊(2nπ,0)3𝔸d𝕊(2π,0)3[r,θ,η]n[r,θ,η]1,\begin{array}[]{cccc}\mathcal{C}_{n}:&\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(2n\pi,0)}&\to&\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(2\pi,0)}\\ &[r,\theta,\eta]_{n}&\mapsto&[r,\theta,\eta]_{1},\end{array} (26)

is a degree-nn covering map, which is also the case for the map 𝒯n\mathcal{T}_{n} defined by

𝒯n:=TCn(Tλn)1:𝔸d𝕊3𝔸d𝕊3RθnA(r)RηRθA(r)Rη.\begin{array}[]{cccc}\mathcal{T}_{n}:=T\circ C_{n}\circ(T\circ\lambda_{n})^{-1}:&\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}&\to&\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}\\ &R^{\frac{\theta}{n}}A(r)R^{-\eta}&\mapsto&R^{\theta}A(r)R^{-\eta}.\end{array} (27)

We summarize this discussion in the following lemma.

Lemma 4.4.

The space 𝔸d𝕊(2nπ,0)3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(2n\pi,0)} is a degree-nn covering of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}.

We now move on to define anti-de Sitter manifolds with spin-cone singularities. Let MM be an oriented three-manifold, and let LL be a link in MM, i.e., a finite disjoint union of embedded circles KiMK_{i}\subset M. For each KiK_{i}, we consider TiT_{i}, a tubular neighborhood of KiK_{i}. Each such neighborhood is homeomorphic to the solid torus 𝔻2×𝕊1\mathbb{D}^{2}\times\mathbb{S}^{1}, and the complement TiKiT_{i}\setminus K_{i} is homeomorphic to 𝔻2×𝕊1\mathbb{D}^{2}_{*}\times\mathbb{S}^{1}, where 𝔻2:=𝔻2{(0,0)}\mathbb{D}^{2}_{*}:=\mathbb{D}^{2}\setminus\{(0,0)\} denotes the punctured disc.

On the universal cover of 𝔻2×𝕊1\mathbb{D}^{2}_{*}\times\mathbb{S}^{1}, we introduce cylindrical coordinates (r,x,y)(0,1)××(r,x,y)\in(0,1)\times\mathbb{R}\times\mathbb{R} such that the universal covering map is given by

𝒞:(0,1)××𝔻2×𝕊1(r,x,y)(reix,eiy).\begin{array}[]{ccccc}\mathcal{C}&:&(0,1)\times\mathbb{R}\times\mathbb{R}&\to&\mathbb{D}^{2}_{*}\times\mathbb{S}^{1}\\ &&(r,x,y)&\mapsto&(re^{ix},e^{iy}).\end{array} (28)

We define the universal branched cover of 𝔻2×𝕊1\mathbb{D}^{2}\times\mathbb{S}^{1} branched over 𝕊1\mathbb{S}^{1} as the quotient:

([0,1)××):=([0,1)××)/,\left([0,1)\times\mathbb{R}\times\mathbb{R}\right)\sqcup\mathbb{R}:=\left([0,1)\times\mathbb{R}\times\mathbb{R}\right)/\sim, (29)

where (0,x,y)(0,x,y)(0,x,y)\sim(0,x^{\prime},y) for any x,xx,x^{\prime}\in\mathbb{R} and yy\in\mathbb{R}. The real line \mathbb{R} attached to [0,1)××[0,1)\times\mathbb{R}\times\mathbb{R} can thus be identified with {[0,0,y]y}\{[0,0,y]\mid y\in\mathbb{R}\}, collapsing the 2\mathbb{R}^{2}-plane along the xx-axis. Furthermore, observe that the covering map 𝒞\mathcal{C} defined in (28) extends to ([0,1)××)/\left([0,1)\times\mathbb{R}\times\mathbb{R}\right)/\sim by taking 𝒞([0,x,y])=(0,eiy)\mathcal{C}([0,x,y])=(0,e^{iy}). Since TiKi𝔻2×𝕊1T_{i}\setminus K_{i}\cong\mathbb{D}^{2}_{*}\times\mathbb{S}^{1}, we may similarly define the universal cover of TiT_{i} branched over KiK_{i} and denote it by TiKi~Ki~\widetilde{T_{i}\setminus K_{i}}\sqcup\widetilde{K_{i}}.

Definition 4.5.

Let LL be a link in MM as defined above. We say that an anti-de Sitter structure on MLM\setminus L has spin-cone singularities along LL if the following conditions hold:

  • The restriction of the developing map

    Dev:TjK~𝔸d𝕊3\mathrm{Dev}:\widetilde{T_{j}\setminus K}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}

    extends continuously to TjKj~Kj~\widetilde{T_{j}\setminus K_{j}}\sqcup\widetilde{K_{j}}. Namely, in cylindrical coordinates (r,θ,η)(0,1)××(r,\theta,\eta)\in(0,1)\times\mathbb{R}\times\mathbb{R}, the limit

    limr0Dev(r,θ,η)=:f(η)\lim_{r\rightarrow 0}\mathrm{Dev}(r,\theta,\eta)=:f(\eta) (30)

    exists and is independent of θ\theta. See Figure 1.

  • The map ff sends Kj~\widetilde{K_{j}} onto a complete timelike geodesic j𝔸d𝕊3\ell_{j}\subset\mathbb{A}\mathrm{d}\mathbb{S}^{3}.

  • The lifted holonomy ρ:π1(TjKj)Isom(𝔸d𝕊3j~)\rho:\pi_{1}(T_{j}\setminus K_{j})\to\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}\setminus\ell_{j}}) around a meridian α\alpha encircling KjK_{j} is given by φ(θj,ηj)\varphi_{(\theta_{j},\eta_{j})} for some θj,ηj\theta_{j},\eta_{j}\in\mathbb{R}, and the holonomy of a longitude β\beta is φ(0,2π)\varphi_{(0,2\pi)} (see (22) for notation).

We say that the structure is branched (or that MM is a branched AdS manifold) if θj,ηj2π\theta_{j},\eta_{j}\in 2\pi\mathbb{Z}.

As a consequence of the above definition, each tubular neighborhood TjKjT_{j}\setminus K_{j} is locally isometric to the local model of a spin-cone singular manifold 𝔸d𝕊(θj,ηj)3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(\theta_{j},\eta_{j})}.

Refer to caption
Figure 1. The behavior of the developing map restricted to the slice (0,1)××{y}(0,1)\times\mathbb{R}\times\{y\}.

5. From domination to anti-de Sitter manifolds

The principal result of this section is Theorem 5.3. Our construction builds on Janigro’s thesis [Jan22] but extends it in several essential ways. In her work, Janigro considers a pair of maps defined on the complement of the singular set of a hyperbolic cone surface, with hh an immersion. In contrast, we consider a general pair of dominating maps f,hf,h defined on closed hyperbolic surfaces, with no restriction on hh being an immersion. In particular, we provide a precise definition of the AdS manifold associated with such a pair of dominating maps and introduce its “completed” version that includes the singular locus (see Definition 5.2). This refinement yields new topological results (Proposition 5.5) beyond those established in [Jan22]. These results play an important role in the proof of our main theorem on AdS manifolds (Theorem B).

Throughout this section, we consider an oriented surface Σ\Sigma (not necessarily closed) with fundamental group Γ\Gamma, and two smooth maps f,h:Σ~2f,h:\widetilde{\Sigma}\to\mathbb{H}^{2} such that

  1. (a)

    hh is a local diffeomorphism on an open dense subset. We denote by CΣ~C\subset\widetilde{\Sigma} the subset of points where hh is singular. Note that any harmonic map either has this property or has an image contained in a geodesic [Sam78, Theorem 3].

  2. (b)

    Denote by {Uα}αI\{U_{\alpha}\}_{\alpha\in I} an open cover of Σ~C\widetilde{\Sigma}\setminus C such that h|Uαh|_{U_{\alpha}} is a diffeomorphism onto its image. We require the cover to be Γ\Gamma-invariant, in the sense that the indices II are labeled equivariantly under the deck transformation action

    γα=β if and only if γUα=Uβ.\textit{}\gamma\cdot\alpha=\beta\text{ if and only if }\gamma\cdot U_{\alpha}=U_{\beta}. (31)
  3. (c)

    The map hh dominates ff, i.e. fσ<hσf^{*}\sigma<h^{*}\sigma. In particular, we may shrink the neighborhood UαU_{\alpha} so that for all x,yUαx,y\in U_{\alpha}, we have

    dσ(f(x),f(y))<dσ(h(x),h(y)).d_{\sigma}\bigl(f(x),f(y)\bigr)<d_{\sigma}\bigl(h(x),h(y)\bigr).

5.1. Gluing of the fibration

For each open set UαU_{\alpha}, we define

Mα:={APSL(2,)|!pUα such that A(f(p))=h(p)}.M_{\alpha}:=\Bigl\{\,A\in\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\;\Bigm|\;\exists!\,p\in U_{\alpha}\text{ such that }A(f(p))=h(p)\Bigr\}.

It turns out that MαM_{\alpha} is an open subset of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}, foliated by timelike geodesics, a fact observed in [GK17, Proposition 7.2].

Proposition 5.1 ([Sag24, Proposition 6.1]).

For each αI\alpha\in I, the subset Mα𝔸d𝕊3M_{\alpha}\subset\mathbb{A}\mathrm{d}\mathbb{S}^{3} is open. Moreover, the map

α:MαUα,Ap,\mathcal{F}_{\alpha}:M_{\alpha}\longrightarrow U_{\alpha},\quad A\longmapsto p,

where pp is the unique point in UαU_{\alpha} satisfying A(f(p))=h(p)A(f(p))=h(p), defines a principal 𝕊1\mathbb{S}^{1}–bundle whose fibers are timelike geodesics in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}.

Next, for each branched point xCx\in C, consider the timelike geodesic

x:=f(x),h(x)={APSL(2,)A(h(x))=f(x)}𝕊1.\ell_{x}:=\ell_{f(x),h(x)}=\{\,A\in\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\mid A(h(x))=f(x)\}\cong\mathbb{S}^{1}.

For each αI\alpha\in I and xCx\in C, we denote by α:Mα𝔸d𝕊3\mathcal{I}_{\alpha}:M_{\alpha}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3} and x:x𝔸d𝕊3\mathcal{I}_{x}:\ell_{x}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3} the inclusion maps. We then assemble two disjoint unions:

f,h=αIMα,𝒮f,h=xCx,\mathcal{R}_{f,h}\;=\;\bigsqcup_{\alpha\in I}M_{\alpha},\qquad\mathcal{S}_{f,h}\;=\;\bigsqcup_{x\in C}\ell_{x}, (32)

where the letters \mathcal{R} and 𝒮\mathcal{S} stand for regular and singular, respectively. Define two maps:

  • :f,h𝒮f,h𝔸d𝕊3by{|Mα=α,|x=x,\mathcal{I}\;:\;\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\;\longrightarrow\;\mathbb{A}\mathrm{d}\mathbb{S}^{3}\quad\text{by}\quad\begin{cases}\mathcal{I}\bigl|_{M_{\alpha}}=\mathcal{I}_{\alpha},\\ \mathcal{I}\bigl|_{\ell_{x}}=\mathcal{I}_{x},\end{cases}
  • :f,h𝒮f,hΣ~by{|Mα=α,|x=(constant map equal to x).\mathcal{F}\;:\;\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\;\longrightarrow\;\widetilde{\Sigma}\quad\text{by}\quad\begin{cases}\mathcal{F}\bigl|_{M_{\alpha}}=\mathcal{F}_{\alpha},\\ \mathcal{F}\bigl|_{\ell_{x}}=\text{(constant map equal to }x\text{).}\end{cases}

We equip f,h𝒮f,h\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h} with the initial topology induced by the map

f,h𝒮f,h𝔸d𝕊3×Σ~,A((A),(A)).\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\;\longrightarrow\;\mathbb{A}\mathrm{d}\mathbb{S}^{3}\times\widetilde{\Sigma},\qquad A\;\longmapsto\;\bigl(\mathcal{I}(A),\,\mathcal{F}(A)\bigr).

In particular, a sequence Anf,h𝒮f,hA_{n}\in\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h} converges to Af,h𝒮f,hA\in\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h} if and only if

(An)(A)in 𝔸d𝕊3and(An)(A)in Σ~.\mathcal{I}(A_{n})\;\longrightarrow\;\mathcal{I}(A)\quad\text{in }\mathbb{A}\mathrm{d}\mathbb{S}^{3}\quad\text{and}\quad\mathcal{F}(A_{n})\;\longrightarrow\;\mathcal{F}(A)\quad\text{in }\widetilde{\Sigma}.
Definition 5.2.

Let f,h:X~2f,h:\widetilde{X}\to\mathbb{H}^{2} be as above. We define

𝔸d𝕊f,h:=(f,h𝒮f,h)/,𝔸d𝕊f,h:=f,h/,\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\;:=\;\bigl(\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\bigr)\;\big/\!\sim,\qquad\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\;:=\;\mathcal{R}_{f,h}\;\big/\!\sim,

where A,BA,B in f,h𝒮f,h\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h} (or in f,h\mathcal{R}_{f,h} ) are equivalent if and only if

(A)=(B)and(A)=(B).\mathcal{I}(A)=\mathcal{I}(B)\quad\text{and}\quad\mathcal{F}(A)=\mathcal{F}(B). (33)

Using the equivalence relation above, we may define maps

:f,h𝒮f,h𝔸d𝕊3and:f,h𝒮f,hΣ~,\mathcal{I}\;:\;\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\;\longrightarrow\;\mathbb{A}\mathrm{d}\mathbb{S}^{3}\quad\text{and}\quad\mathcal{F}\;:\;\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\;\longrightarrow\;\widetilde{\Sigma},

which descend to well‐defined maps on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}. We state now the principal result of this section.

Theorem 5.3.

Let f,h:Σ~2f,h:\widetilde{\Sigma}\to\mathbb{H}^{2} be maps satisfying conditions (a)(c). Then,

  • 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} is topologically a solid torus.

  • 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is an anti–de Sitter manifold with an atlas of charts

    {(Pr(Mα),(Pr|Mα)1)}αI.\bigl\{\,(\Pr(M_{\alpha}),\;\mathcal{I}\circ(\Pr|_{M_{\alpha}})^{-1})\bigr\}_{\alpha\in I}.

    Moreover, :𝔸d𝕊f,hΣ~C\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\widetilde{\Sigma}\setminus C is a principal 𝕊1\mathbb{S}^{1}–bundle with timelike geodesic fibers.

Remark 5.4.

Although the constructions of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} and 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} rely on the choice of an open covering 𝒰={Uα}αI\mathcal{U}=\{U_{\alpha}\}_{\alpha\in I} of Σ~C\widetilde{\Sigma}\setminus C, different choices yield a canonical identification. To explain this, we temporarily denote the spaces by 𝔸d𝕊f,h(𝒰)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}(\mathcal{U}) and 𝔸d𝕊f,h(𝒰)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}(\mathcal{U}) (instead of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} and 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}) to keep track of the covering, and let [A]𝒰[A]_{\mathcal{U}} denote the equivalence class of APSL(2,)A\in\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}) in 𝔸d𝕊f,h(𝒰)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}(\mathcal{U}). Then, for another covering 𝒱\mathcal{V} of Σ~C\widetilde{\Sigma}\setminus C, the natural map

𝔸d𝕊f,h(𝒰)𝔸d𝕊f,h(𝒱),[A]𝒰[A]𝒱\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}(\mathcal{U})\longrightarrow\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}(\mathcal{V}),\qquad[A]_{\mathcal{U}}\longmapsto[A]_{\mathcal{V}}

defines a homeomorphism between 𝔸d𝕊f,h(𝒰)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}(\mathcal{U}) and 𝔸d𝕊f,h(𝒱)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}(\mathcal{V}), which restricts to an isometry between the resulting AdS manifolds 𝔸d𝕊f,h(𝒰)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}(\mathcal{U}) and 𝔸d𝕊f,h(𝒱)\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}(\mathcal{V}). We refer the reader to [Jan22, Section 3.3, p. 58] for a discussion of this functorial behavior in her (similar) context.

5.2. Topology of the gluing

We start with the following proposition, which describes the topology of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}.

Proposition 5.5.

There is a homeomorphism 𝔸d𝕊f,hh(Σ~)×𝕊1\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to h(\widetilde{\Sigma})\times\mathbb{S}^{1} that restricts to a homeomorphism between 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} and h(Σ~C)×𝕊1h(\widetilde{\Sigma}\setminus C)\times\mathbb{S}^{1}.

To prepare the argument, consider the disjoint union

𝒳=αIUαC,\mathcal{X}\;=\;\bigsqcup_{\alpha\in I}U_{\alpha}\;\;\bigsqcup\;C, (34)

where each UαΣ~CU_{\alpha}\subset\widetilde{\Sigma}\setminus C is an open set as before. We now define a map H:𝒳2H\;:\;\mathcal{X}\;\longrightarrow\;\mathbb{H}^{2} by

H(y):=h(y).H(y):=h(y). (35)

for yy in UαU_{\alpha} or in CC. We endow 𝒳\mathcal{X} with the initial topology making HH continuous; that is, a subset V𝒳V\subset\mathcal{X} is open if and only if

V=H1(W)for some open Wh(Σ~).V\;=\;H^{-1}(W)\quad\text{for some open }W\subset h(\widetilde{\Sigma}).

We then define on 𝒳\mathcal{X} the equivalence relation

xyH(x)=H(y).x\sim y\quad\Longleftrightarrow\quad H(x)\;=\;H(y).
Lemma 5.6.

The quotient 𝒳/\mathcal{X}/\!\sim is homeomorphic to the image h(Σ~)2h(\widetilde{\Sigma})\subset\mathbb{H}^{2}.

Proof.

Let q:𝒳𝒳/q:\mathcal{X}\to\;\mathcal{X}/\!\sim be the quotient map, and define

H¯:𝒳/h(Σ~),H¯(q(x))=H(x).\overline{H}\;:\;\mathcal{X}/\!\sim\;\longrightarrow\;h(\widetilde{\Sigma}),\quad\overline{H}\bigl(q(x)\bigr)\;=\;H(x).

Since HH satisfies

H(x)=H(y)xy,H(x)=H(y)\quad\Longleftrightarrow\quad x\sim y,

the induced map H¯\overline{H} is well-defined. Moreover, by the definition of the initial topology, this map is continuous. The inverse of H¯\overline{H} is given by

H¯1:h(Σ~)𝒳/,H¯1(z)=q(p),\overline{H}^{-1}\;:\;h(\widetilde{\Sigma})\;\longrightarrow\;\mathcal{X}/\!\sim,\quad\overline{H}^{-1}(z)\;=\;q(p),

where p𝒳p\in\mathcal{X} is any point satisfying H(p)=zH(p)=z. We claim that H¯1\overline{H}^{-1} is continuous. Let U𝒳/U^{\prime}\subset\mathcal{X}/\!\sim be any open set. We must show that

(H¯1)1(U)h(Σ~)\left(\overline{H}^{-1}\right)^{-1}(U^{\prime})\;\subset\;h(\widetilde{\Sigma})

is open. By the definition of the quotient topology and the initial topology on 𝒳\mathcal{X}, we have that q1(U)=H1(W)q^{-1}(U^{\prime})=H^{-1}(W) for some open set Wh(Σ~)W\subset h(\widetilde{\Sigma}). This implies that

(H¯1)1(U)=W,\left(\overline{H}^{-1}\right)^{-1}(U^{\prime})=W,

which is open in h(Σ~)h(\widetilde{\Sigma}). Hence H¯1\overline{H}^{-1} is continuous. Because H¯\overline{H} is a continuous bijection with a continuous inverse, it is a homeomorphism. This completes the proof. ∎

Next, we define a map F:𝒳2F:\mathcal{X}\longrightarrow\mathbb{H}^{2} by setting F(y):=f(y)F(y):=f(y) for yy in UαU_{\alpha} or in CC. The following lemma shows that, since hh dominates ff, the map FF descends to the quotient 𝒳/\mathcal{X}/\sim.

Lemma 5.7.

The map F:𝒳2F:\mathcal{X}\to\mathbb{H}^{2} is continuous and induces a continuous map F¯:𝒳/2\overline{F}:\mathcal{X}/\sim\to\mathbb{H}^{2}.

Proof.

Since hh dominates ff, the inequality

dσ(F(x),F(y))dσ(H(x),H(y))d_{\sigma}\bigl(F(x),\,F(y)\bigr)\;\leq\;d_{\sigma}\bigl(H(x),\,H(y)\bigr)

holds for all x,y𝒳x,y\in\mathcal{X}. In particular, if a sequence xnxx_{n}\to x in 𝒳\mathcal{X}, then H(xn)H(x)H(x_{n})\to H(x), and hence

dσ(F(xn),F(x))dσ(H(xn),H(x)) 0.d_{\sigma}\bigl(F(x_{n}),\,F(x)\bigr)\;\leq\;d_{\sigma}\bigl(H(x_{n}),\,H(x)\bigr)\;\longrightarrow\;0.

This shows that F:𝒳2F:\mathcal{X}\to\mathbb{H}^{2} is continuous with respect to the topology of 𝒳\mathcal{X}. Moreover, if xyx\sim y, then H(x)=H(y)H(x)=H(y), so

dσ(F(x),F(y))dσ(H(x),H(y))= 0,d_{\sigma}\bigl(F(x),\,F(y)\bigr)\;\leq\;d_{\sigma}\bigl(H(x),\,H(y)\bigr)\;=\;0,

and therefore F(x)=F(y)F(x)=F(y). Hence, we can define F¯:𝒳/2\overline{F}:\mathcal{X}/\sim\to\mathbb{H}^{2} by F¯(q(x))=F(x)\overline{F}(q(x))=F(x). ∎

We are now in position to prove Proposition 5.5.

Proof of Proposition 5.5.

For each p2p\in\mathbb{H}^{2}, we denote by BpB_{p} the unique hyperbolic isometry that sends ii to pp and whose axis is the oriented geodesic joining ii to pp.

To construct the homeomorphism between 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} and h(Σ~)×𝕊1h(\widetilde{\Sigma})\times\mathbb{S}^{1}, we consider the map

Φ:𝒳×𝕊1f,h𝒮f,h(y,θ)BH(y)RθBF(y)1.\begin{array}[]{ccccc}\Phi&:&\mathcal{X}\times\mathbb{S}^{1}&\longrightarrow&\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\\ &&(y,\theta)&\longmapsto&B_{H(y)}R^{\theta}B_{F(y)}^{-1}.\end{array}

Observe that Φ\Phi is continuous–this follows from the continuity of HH and FF. Next, note that if pqp\sim q in 𝒳\mathcal{X}, then by definition H(p)=H(q)H(p)=H(q) but also F(p)=F(q)F(p)=F(q) by the proof of Lemma 5.7. This implies that Φ(p,θ)=Φ(q,θ)\Phi(p,\theta)=\Phi(q,\theta) for any θ𝕊1\theta\in\mathbb{S}^{1}. Therefore, we may define a continuous map between 𝒳/×𝕊1\mathcal{X}/\!\sim\times\mathbb{S}^{1} and 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} as follows:

Φ¯:𝒳/×𝕊1𝔸d𝕊f,h([y],θ)[BH(y)RθBF(y)1],\begin{array}[]{ccccc}\overline{\Phi}&:&\mathcal{X}/\!\sim\times\mathbb{S}^{1}&\longrightarrow&\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\\ &&([y],\theta)&\longmapsto&[B_{H(y)}R^{\theta}B_{F(y)}^{-1}],\end{array}

where [][\cdot] denotes the equivalence classes in both 𝒳/\mathcal{X}/\!\sim and 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}. We aim to show Φ¯\overline{\Phi} is a homeomorphism.

First, we define an inverse for Φ¯\overline{\Phi}. To this end, we identify the circle 𝕊1\mathbb{S}^{1} with the timelike geodesic i,i\ell_{i,i} via the map eiθRθe^{i\theta}\mapsto R^{\theta}. Then, the inverse of Φ\Phi is given by

Φ1:f,h𝒮f,h𝒳×𝕊1A((A),BH((A))1(A)BF((A))),\begin{array}[]{ccccc}\Phi^{-1}&:&\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}&\longrightarrow&\mathcal{X}\times\mathbb{S}^{1}\\ &&A&\longmapsto&\left(\mathcal{F}(A),\;B_{H(\mathcal{F}(A))}^{-1}\,\mathcal{I}(A)\,B_{F(\mathcal{F}(A))}\right),\end{array}

where :f,h𝒮f,h𝔸d𝕊3\mathcal{I}:\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3} and :f,h𝒮f,hΣ~\mathcal{F}:\mathcal{R}_{f,h}\sqcup\mathcal{S}_{f,h}\to\widetilde{\Sigma} are the previously defined maps. The map Φ1\Phi^{-1} is continuous. Again, observe that if ABA\sim B, then Φ1(A)=Φ1(B)\Phi^{-1}(A)=\Phi^{-1}(B). This allows us to define a continuous inverse

Φ¯1:𝔸d𝕊f,h𝒳/×𝕊1[A]([(A)],BF((A))1(A)BH((A))).\begin{array}[]{ccccc}\overline{\Phi}^{-1}&:&\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}&\longrightarrow&\mathcal{X}/\!\sim\times\mathbb{S}^{1}\\ &&[A]&\longmapsto&\left([\mathcal{F}(A)],\;B_{F(\mathcal{F}(A))}^{-1}\,\mathcal{I}(A)\,B_{H(\mathcal{F}(A))}\right).\end{array}

As a consequence, 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} is homeomorphic to 𝒳/\mathcal{X}/\sim, which in turn is homeomorphic to h(Σ~)×𝕊1h(\widetilde{\Sigma})\times\mathbb{S}^{1} by Lemma 5.6. It is straightforward to see that the above construction also gives rise to a homeomorphism between 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} and h(Σ~C)h(\widetilde{\Sigma}\setminus C). This completes the proof. ∎

5.3. Anti-de Sitter structure

We turn our attention to the regular part 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. We will show that it admits an anti-de Sitter structure.

An anti-de Sitter structure on a three manifold MM is a (Isom0(𝔸d𝕊3),𝔸d𝕊3)(\mathrm{Isom}_{0}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}),\mathbb{A}\mathrm{d}\mathbb{S}^{3}) structure. By definition, this is a maximal atlas of coordinate charts on MM with values in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} such that the transition maps are given by elements of Isom0(𝔸d𝕊3)\mathrm{Isom}_{0}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}).

Consider the projection map

Pr:αIMα𝔸d𝕊f,h=(αIMα)/A[A].\begin{array}[]{ccccc}\mathrm{Pr}&:&\bigsqcup_{\alpha\in I}M_{\alpha}&\to&\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}=\left(\bigsqcup_{\alpha\in I}M_{\alpha}\right)\Big/\sim\\ &&A&\mapsto&[A].\\ \end{array}

Using the topology induced on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}, we have the following.

  • The restriction of Pr\mathrm{Pr} to MαM_{\alpha} is a homeomorphism onto its image.

  • Pr(Mα)\mathrm{Pr}(M_{\alpha}) is an open set of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}.

Using this, we can show the following.

Proposition 5.8.

𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is an anti-de Sitter manifold with an atlas of charts

{(Pr(Mα),(Pr|Mα)1)}αI.\bigl\{\,(\Pr(M_{\alpha}),\;\mathcal{I}\circ(\Pr|_{M_{\alpha}})^{-1})\bigr\}_{\alpha\in I}.

Moreover, :𝔸d𝕊f,hΣ~C\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\widetilde{\Sigma}\setminus C is a principal 𝕊1\mathbb{S}^{1}–bundle with timelike-geodesic fibers.

Proof.

The atlas of charts clearly defines an anti-de Sitter structure on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. Since

=α(Pr|Mα)1,\mathcal{F}=\mathcal{F}_{\alpha}\circ\mathcal{I}\circ\left(\Pr|_{M_{\alpha}}\right)^{-1},

and the fibers of α\mathcal{F}_{\alpha} are timelike geodesics, it follows that the fibers of \mathcal{F} are timelike geodesics with respect to the anti-de Sitter structure induced by the atlas of charts

{(Pr(Mα),(Pr|Mα)1)}αI.\left\{\,\left(\Pr(M_{\alpha}),\;\mathcal{I}\circ\left(\Pr|_{M_{\alpha}}\right)^{-1}\right)\right\}_{\alpha\in I}.

The above atlas of charts was introduced in [Jan22, Section 3.2.3]. We now show that \mathcal{F} is a fibration by circles, i.e., it is locally trivial. For each αI\alpha\in I, the map α:MαUα\mathcal{F}_{\alpha}:M_{\alpha}\to U_{\alpha} is a fibration over the contractible open set UαU_{\alpha}, and thus it is trivial. Therefore, there exists a diffeomorphism

ϕα:MαUα×𝕊1\phi_{\alpha}:M_{\alpha}\to U_{\alpha}\times\mathbb{S}^{1}

such that Pr1ϕα=α\Pr_{1}\circ\phi_{\alpha}=\mathcal{F}_{\alpha}, where Pr1:Uα×𝕊1Uα\Pr_{1}:U_{\alpha}\times\mathbb{S}^{1}\to U_{\alpha} is the projection onto the first factor. For each index αI\alpha\in I, we consider the map

Tα:1(Uα)Mα,[A]A,T_{\alpha}:\mathcal{F}^{-1}(U_{\alpha})\to M_{\alpha},\quad[A]\mapsto A,

where [][\cdot] denotes the equivalence class in 1(Uα)𝔸d𝕊f,h\mathcal{F}^{-1}(U_{\alpha})\subset\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. We claim that TαT_{\alpha} is a homeomorphism.

First, for each [A]1(Uα)[A]\in\mathcal{F}^{-1}(U_{\alpha}), we necessarily have AMαA\in M_{\alpha}. Indeed, by definition, ([A])=α(A)Uα\mathcal{F}([A])=\mathcal{F}_{\alpha}(A)\in U_{\alpha}, hence A(f(p))=h(p)A(f(p))=h(p), that is, AMαA\in M_{\alpha}. Second, the map is well-defined. Indeed, if [A]=[B][A]=[B] with [A],[B]1(Uα)[A],[B]\in\mathcal{F}^{-1}(U_{\alpha}), then we have seen that A,BMαA,B\in M_{\alpha}, and because AA and BB are equivalent, we have (A)=(B)\mathcal{I}(A)=\mathcal{I}(B). Since, |Mα\mathcal{I}|_{M_{\alpha}} is the inclusion map, we deduce A=BA=B.

Next, the inverse of TαT_{\alpha} is clearly given by

Tα1:Mα1(Uα),A[A].T_{\alpha}^{-1}:M_{\alpha}\to\mathcal{F}^{-1}(U_{\alpha}),\quad A\mapsto[A].

The continuity of both TαT_{\alpha} and Tα1T_{\alpha}^{-1} follows from the definition of the topology on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. Now, since Tαϕα:1(Uα)Uα×𝕊1T_{\alpha}\circ\phi_{\alpha}:\mathcal{F}^{-1}(U_{\alpha})\to U_{\alpha}\times\mathbb{S}^{1} is a local trivialization of \mathcal{F}, this completes the proof. ∎

The proof of Theorem 5.3 follows by combining Propositions 5.5 and 5.8.

We end this section by studying the case where the maps f,h:Σ~2f,h:\widetilde{\Sigma}\to\mathbb{H}^{2} are equivariant with respect to the representations ρ\rho and j:ΓPSL(2,)j:\Gamma\to\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R}), respectively. In what follows, we denote

Γj,ρ={(j(γ),ρ(γ)):γΓ}.\Gamma_{j,\rho}=\{(j(\gamma),\rho(\gamma)):\gamma\in\Gamma\}.

We will show that Γj,ρ\Gamma_{j,\rho} acts properly discontinuously on both 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} and 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. In particular, it will follow from Proposition 5.8 that the quotient of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is an anti–de Sitter manifold.

First, we describe this action. For each γΓ\gamma\in\Gamma, let H(γ)=(j(γ),ρ(γ))H(\gamma)=(j(\gamma),\rho(\gamma)). Then H(γ)H(\gamma) induces a map

H¯(γ):𝔸d𝕊f,h𝔸d𝕊f,h\overline{H}(\gamma):\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} (36)

defined by

H¯(γ)[A]:=[j(γ)(A)ρ(γ)1].\overline{H}(\gamma)\cdot[A]:=[\,j(\gamma)\,\mathcal{I}(A)\,\rho(\gamma)^{-1}\,]. (37)

It is straightforward to verify that the map :𝔸d𝕊f,hΣ~\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\widetilde{\Sigma} satisfies the following equivariance property: for all γΓ\gamma\in\Gamma and [A]𝔸d𝕊f,h[A]\in\mathbb{A}\mathrm{d}\mathbb{S}^{f,h},

(H¯(γ)[A])=γ([A]).\mathcal{F}\!\left(\overline{H}(\gamma)\cdot[A]\right)=\gamma\cdot\mathcal{F}([A]).

Since \mathcal{F} is continuous and equivariant, and Γ\Gamma acts properly discontinuously on its universal cover Σ~\widetilde{\Sigma}, it follows that the action of Γj,ρ\Gamma_{j,\rho} on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} is properly discontinuous.

On the other hand, 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} is a Hausdorff topological space by Proposition 5.5, and thus the quotient 𝔸d𝕊f,h/Γj,ρ\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}/\Gamma_{j,\rho} is a well–defined Hausdorff topological space. Moreover, the quotient 𝔸d𝕊f,h/Γj,ρ\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}/\Gamma_{j,\rho} inherits an anti–de Sitter structure. We summarize this discussion in the following lemma.

Lemma 5.9.

Let f,h:Σ~2f,h:\widetilde{\Sigma}\to\mathbb{H}^{2} be maps equivariant with respect to the representations j,ρ:ΓPSL(2,)j,\rho:\Gamma\to\mathrm{PSL}(2,\mathbb{R}), respectively, and satisfying conditions (a)(c). Then the map :𝔸d𝕊f,hΣ~\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\widetilde{\Sigma} is equivariant with respect to the action of Γj,ρ\Gamma_{j,\rho} on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} via H¯\overline{H} (see (36)) and the action of Γ\Gamma on Σ~\widetilde{\Sigma}. In particular, Γj,ρ\Gamma_{j,\rho} acts properly discontinuously on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}, and the quotient

f,h:=𝔸d𝕊f,h/Γj,ρ\mathcal{M}^{f,h}:=\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\big/\Gamma_{j,\rho}

is a well-defined Hausdorff topological space. Moreover, the map \mathcal{F} induces a principal 𝕊1\mathbb{S}^{1}-bundle over the anti-de Sitter manifold

f,h:=𝔸d𝕊f,h/Γj,ρ,\mathcal{M}_{*}^{f,h}:=\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\big/\Gamma_{j,\rho},

with base Σ\Sigma_{*}, where Σ\Sigma_{*} is the quotient of Σ~C\widetilde{\Sigma}\setminus C by the action of Γ\Gamma. We continue to denote this map by

:f,hΣ.\mathcal{F}:\mathcal{M}_{*}^{f,h}\to\Sigma_{*}.

6. Branched AdS manifolds from branched immersions

In this section, we study the anti-de Sitter manifold obtained from a pair f,hf,h as above and under the additional assumption that hh is a branched immersion. Recall that a map hh between surfaces has a branch point at pp if there exist local complex coordinates zz at pp and ww at h(p)h(p) such that hh takes the form zw=znz\mapsto w=z^{n}. We say that hh is a branched immersion if it is an immersion outside a discrete set of branch points. The main result of this section is the proof of Theorem B, which is completed at the end.

Our approach follows, in spirit, the arguments developed by Janigro in her thesis [Jan22], within her framework for analyzing local singularities of anti–de Sitter 33-manifolds. To extend her framework to our more general setting (outlined in Section 5), we introduce several technical refinements and additional results, which are established in the course of the proof of Proposition 6.1.

6.1. The fundamental example

Before proving Theorem B, we first treat the special case where Σ=2\Sigma=\mathbb{H}^{2} and the map h:Σ~2h:\widetilde{\Sigma}\to\mathbb{H}^{2} has a single branch point. Understanding this example is an essential step toward the proof of the general theorem. We consider smooth maps f,h:22f,h:\mathbb{H}^{2}\to\mathbb{H}^{2} such that hh dominates ff (i.e. fσ<hσf^{*}\sigma<h^{*}\sigma). Without loss of generality, by composing with isometries, we may assume that h(i)=f(i)=ih(i)=f(i)=i. We have seen in Section 5 how to construct an anti-de Sitter manifold from the maps ff and hh. Recall that

𝔸d𝕊f,h:=(αIMαi,i)/,\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}:=\left(\bigsqcup_{\alpha\in I}M_{\alpha}\sqcup\ell_{i,i}\right)\Big/\sim,

where Mα:={APSL(2,)|!xUα such that A(f(x))=h(x)},M_{\alpha}:=\left\{A\in\mathrm{P}\mathrm{S}\mathrm{L}(2,\mathbb{R})\;\middle|\;\exists!\,x\in U_{\alpha}\text{ such that }A(f(x))=h(x)\right\}, and {Uα}αI\{U_{\alpha}\}_{\alpha\in I} is an open cover of 2:=2{i}\mathbb{H}^{2}_{*}:=\mathbb{H}^{2}\setminus\{i\} on which hh is a local diffeomorphism. Note that in this case, MαM_{\alpha} does not contain the timelike geodesic i,i\ell_{i,i}, because ff and hh take values in 2\mathbb{H}^{2}_{*}. From Proposition 5.8, we can define the local isometry

𝒟:𝔸d𝕊f,h𝔸d𝕊3[A](A).\begin{array}[]{rcl}\mathcal{D}&:&\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}\\ &&[A]\mapsto\mathcal{I}(A).\end{array} (38)

Note that 𝒟\mathcal{D} takes values in 𝔸d𝕊3:=𝔸d𝕊3i,i\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}:=\mathbb{A}\mathrm{d}\mathbb{S}^{3}\setminus\ell_{i,i}. Moreover, 𝒟\mathcal{D} can also be defined on all of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} by setting 𝒟(A)=(A)\mathcal{D}(A)=\mathcal{I}(A) for any [A]i,i[A]\in\ell_{i,i}. We record here some notations and remarks that will be used later.

  • We denote by Dev~:𝔸d𝕊f,h~𝔸d𝕊3~\widetilde{\mathrm{Dev}}:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}} the developing map of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. This can be thought of as the lift to the universal cover of the local isometry 𝒟:𝔸d𝕊f,h𝔸d𝕊3\mathcal{D}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}.

  • We denote by Dev:𝔸d𝕊f,h~𝔸d𝕊3\mathrm{Dev}:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} the developing map onto 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}. This is given by

    Dev=TDev~,\mathrm{Dev}=T\circ\widetilde{\mathrm{Dev}}, (39)

    where T:𝔸d𝕊3~𝔸d𝕊3T:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} is the universal covering map defined in (21).

  • Finally, we denote by Hol~:π1(𝔸d𝕊f,h)Isom0(𝔸d𝕊3~)\widetilde{\mathrm{Hol}}:\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*})\to\mathrm{Isom}_{0}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}) the holonomy representation of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}.

According to Proposition 5.8, there exists an 𝕊1\mathbb{S}^{1}–principal bundle :𝔸d𝕊f,h2.\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\longrightarrow\mathbb{H}^{2}_{*}. This bundle is trivial. Indeed, one can obtain a global trivialization directly from the proof of Proposition 5.5. Alternatively, since any map from a punctured disc to the classifying space \mathbb{CP}^{\infty} is nullhomotopic, any 𝕊1\mathbb{S}^{1}–bundle over a punctured disc must be trivial.

In what follows, we denote by α\alpha a nontrivial loop in π1(𝔸d𝕊f,h)\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}) that is not homotopic to the fiber of \mathcal{F} and whose projection under \mathcal{F} is a generator γ\gamma of π1(2)\pi_{1}(\mathbb{H}^{2}_{*})\cong\mathbb{Z}. (Here, γ\gamma may be viewed as a loop around the branched point of 2\mathbb{H}^{2}_{*}.) We also denote by β\beta the generator of π1(𝔸d𝕊f,h)\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}) that is homotopic to the future-directed timelike geodesic fiber of the fibration :𝔸d𝕊f,h2\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{H}^{2}_{*}. Since the bundle \mathcal{F} is trivial, the fundamental group of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} decomposes as

π1(𝔸d𝕊f,h)=αβ.\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*})\;=\;\mathbb{Z}\alpha\;\oplus\;\mathbb{Z}\beta. (40)

The spacetime 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is time-oriented, so that 11\in\mathbb{Z} corresponds to the future-directed fiber, while 1-1 corresponds to the same fiber with the opposite orientation.

We can now state the main result of this section, which is inspired by [Jan22, pp. 61–69].

Proposition 6.1.

We consider two smooth maps f,h:22f,h:\mathbb{H}^{2}\to\mathbb{H}^{2} such that hh dominates ff. Assume that h(i)=f(i)=ih(i)=f(i)=i and that h:22h:\mathbb{H}^{2}_{*}\to\mathbb{H}^{2}_{*} is a degree-nn covering. Then 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is a branched anti-de Sitter manifold with singular locus [i,i][\ell_{i,i}]. Moreover,

Hol~(α)=φ(2nπ,2kπ)andHol~(β)=φ(0,2π),\widetilde{\mathrm{Hol}}(\alpha)=\varphi_{(2n\pi,2k\pi)}\ \text{and}\ \ \ \widetilde{\mathrm{Hol}}(\beta)=\varphi_{(0,2\pi)},

for some k.k\in\mathbb{Z}. Therefore, the developing map Dev~:𝔸d𝕊f,h~𝔸d𝕊3~\widetilde{\mathrm{Dev}}:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}} induces a local isometry between 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} and 𝔸d𝕊(2nπ,0)3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{(2n\pi,0)}.

Remark 6.2.

It is worth noting that if we change the meridian α\alpha by α=αkβ\alpha^{\prime}=\alpha-k\beta, then the holonomy of α\alpha^{\prime} is φ(2nπ,0)\varphi_{(2n\pi,0)}. Thus, what really matters in the holonomy of the meridian is the degree of hh, and this explains why we do not require any further assumption on the map ff.

Remark 6.3.

The statement of Proposition 6.1 remains valid if the maps ff and hh are defined on a small neighborhood of ii in 2\mathbb{H}^{2} instead of the whole plane. The proof carries over verbatim in this local setting.

We record the following lemma for later use.

Lemma 6.4.

Let f,h:22f,h:\mathbb{H}^{2}\to\mathbb{H}^{2} be as in Proposition 6.1. Then, for all x2x\in\mathbb{H}^{2}, we have the strict inequality

dσ(i,f(x))<dσ(i,h(x)).d_{\sigma}(i,f(x))<d_{\sigma}(i,h(x)).

In particular, the timelike geodesics i,i\ell_{i,i} and h(x),f(x)\ell_{h(x),f(x)} do not intersect.

Proof.

The domination condition fσhσf^{*}\sigma\leq h^{*}\sigma implies that

(fγ)(hγ)\ell(f\circ\gamma)\leq\ell(h\circ\gamma)

for every curve γ\gamma in 2\mathbb{H}^{2}, where ()\ell(\cdot) denotes the length of the curve with respect to the hyperbolic metric σ\sigma. We claim that dσ(i,f(x))dσ(i,h(x))d_{\sigma}(i,f(x))\leq d_{\sigma}(i,h(x)) for all x𝔻2x\in\mathbb{D}^{2}_{*}. For simplicity, we identify 2\mathbb{H}^{2} with the Poincaré disc 𝔻2\mathbb{D}^{2} via an identification sending the point ii to the origin 0 (see (41)). Assume that h(x)=r0eiθ0h(x)=r_{0}e^{i\theta_{0}} for some r0,θ0r_{0},\theta_{0}\in\mathbb{R}. Consider the geodesic ray η:(0,1)𝔻2\eta:(0,1)\to\mathbb{D}^{2}_{*} defined by η(r)=reiθ0\eta(r)=re^{i\theta_{0}}. By the path lifting property, there exists a curve γ:(0,1)𝔻2\gamma:(0,1)\to\mathbb{D}^{2}_{*} such that hγ=ηh\circ\gamma=\eta and γ(r0)=x\gamma(r_{0})=x. This implies that for each 0<t<r00<t<r_{0}, we have

dσ(f(γ(t)),f(x))\displaystyle d_{\sigma}(f(\gamma(t)),f(x)) (fγ|[t,r0])\displaystyle\leq\ell(f\circ\gamma|_{[t,r_{0}]})
(hγ|[t,r0])\displaystyle\leq\ell(h\circ\gamma|_{[t,r_{0}]})
=(η|[t,r0])=dσ(h(γ(t)),h(x)).\displaystyle=\ell(\eta|_{[t,r_{0}]})=d_{\sigma}(h(\gamma(t)),h(x)).

To establish the claim, we need to show that γ(t)0\gamma(t)\to 0 as t0t\to 0. Indeed, for t<r0t<r_{0}, we have γ(t)h1({z𝔻2|z|r0})\gamma(t)\in h^{-1}(\{z\in\mathbb{D}^{2}\mid|z|\leq r_{0}\}), which is a compact subset of 𝔻2\mathbb{D}^{2} by the properness of hh. To see properness, recall that hh is equal to znz^{n} in an appropriate choice of local coordinates around 0. Therefore, γ(t)\gamma(t) must converge to some point x𝔻2x\in\mathbb{D}^{2}. But since h(γ(t))=η(t)h(\gamma(t))=\eta(t) and η(t)0\eta(t)\to 0 as t0t\to 0, it follows that γ(t)0\gamma(t)\to 0. This completes the proof of the claim. To prove the strict inequality, consider yy a point on the geodesic segment between ii and xx. Recall that domination is strict outside the singularities. Hence, for yy sufficiently close to xx, we have dσ(f(x),f(y))<dσ(h(x),h(y))d_{\sigma}\bigl(f(x),f(y)\bigr)<d_{\sigma}\bigl(h(x),h(y)\bigr). It follows that

dσ(f(x),i)dσ(f(x),f(y))+dσ(f(y),i)<dσ(h(x),h(y))+dσ(h(y),i)=dσ(h(x),i).d_{\sigma}\bigl(f(x),i\bigr)\leq d_{\sigma}\bigl(f(x),f(y)\bigr)+d_{\sigma}\bigl(f(y),i\bigr)<d_{\sigma}\bigl(h(x),h(y)\bigr)+d_{\sigma}\bigl(h(y),i\bigr)=d_{\sigma}\bigl(h(x),i\bigr).

This concludes the proof.∎

Remark 6.5.

We denote by [i,i][\ell_{i,i}] the set of classes in 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} that are equivalent to some element lying on i,i\ell_{i,i}. That is,

[i,i]={[A]𝔸d𝕊f,h|(A)=iand(A)i,i}.[\ell_{i,i}]=\left\{[A]\in\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\;\middle|\;\mathcal{F}(A)=i\ \text{and}\ \mathcal{I}(A)\in\ell_{i,i}\right\}.

Observe that if [A][i,i][A]\in[\ell_{i,i}], then AA cannot belong to any MαM_{\alpha}, since this would contradict Lemma 6.4.

The next lemma shows that the developing map of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} satisfies Definition 4.5.

Lemma 6.6.

The developing map Dev:𝔸d𝕊f,h~𝔸d𝕊3\mathrm{Dev}:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} extends continuously to the universal branched cover 𝔸d𝕊f,h~[i,i]~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\sqcup\widetilde{[\ell_{i,i}]}. Moreover, the restriction

Dev|[i,i]~:[i,i]~i,i\mathrm{Dev}\big|_{\widetilde{[\ell_{i,i}]}}:\widetilde{[\ell_{i,i}]}\to\ell_{i,i}

is a universal covering map of the timelike geodesic i,i𝕊1\ell_{i,i}\cong\mathbb{S}^{1}.

Proof.

Let Π:𝔸d𝕊f,h~𝔸d𝕊f,h\Pi:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} be the universal covering map. By definition, the developing map is given by Dev=𝒟Π\mathrm{Dev}=\mathcal{D}\circ\Pi, where 𝒟:𝔸d𝕊f,h𝔸d𝕊3\mathcal{D}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3} is defined by 𝒟([A])=(A)\mathcal{D}([A])=\mathcal{I}(A), as in equation (38). By Proposition 5.5, 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} is a solid torus. Therefore, the universal covering map Π\Pi can be described using the cylindrical coordinates (28), and hence Π\Pi extends continuously to 𝔸d𝕊f,h[i,i]~\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\sqcup\widetilde{[\ell_{i,i}]}. Since 𝒟\mathcal{D} is already defined on all of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h} (and not only on 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}), it follows that Dev\mathrm{Dev} also extends to 𝔸d𝕊f,h~[i,i]~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\sqcup\widetilde{[\ell_{i,i}]}.

For the second claim, note that 𝒟|[i,i]:[i,i]i,i\mathcal{D}\big|_{[\ell_{i,i}]}:[\ell_{i,i}]\to\ell_{i,i} is a homeomorphism, and Dev=𝒟Π\mathrm{Dev}=\mathcal{D}\circ\Pi. Since Π\Pi restricts to the universal cover of [i,i][\ell_{i,i}], the composition Dev|[i,i]~\mathrm{Dev}\big|_{\widetilde{[\ell_{i,i}]}} is the universal covering map of i,i\ell_{i,i}. ∎

6.2. Holonomy around the puncture in the fundamental example

The aim of this section is to compute the holonomy of a nontrivial loop in π1(𝔸d𝕊f,h)\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}) that is not homotopic to a fiber of the fibration :𝔸d𝕊f,h2\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{H}^{2}_{*} and that generates the π1\pi_{1} of the base. As in Section 6.1, we denote such a loop by α\alpha, and we denote by γ\gamma its projection to 2\mathbb{H}^{2}_{*}, which generates π1(2)\pi_{1}(\mathbb{H}^{2}_{*}). We now prove the following.

Proposition 6.7.

Let ff and hh be as in Proposition 6.1. Then we have

Hol~(α)=φ(2nπ,2kπ),\widetilde{\mathrm{Hol}}(\alpha)=\varphi_{(2n\pi,2k\pi)},

for some kk\in\mathbb{Z}.

The proof of Proposition 6.7 needs some preparation.

Definition 6.8.

Let π:(0,+)×2\pi:(0,+\infty)\times\mathbb{R}\to\mathbb{H}^{2}_{*} be the universal covering map of 2\mathbb{H}^{2}_{*} defined in (16). Then we define the angular form ω\omega as the 11-form on 2\mathbb{H}^{2}_{*}

πω=dθ.\pi^{*}\omega=\mathrm{d}\theta.
Remark 6.9.

We record the following observations.

  1. (1)

    If we consider the biholomorphism 𝒞:𝔻22\mathcal{C}:\mathbb{D}^{2}\to\mathbb{H}^{2} defined by

    𝒞:ziziz+i,which mapsito0,\mathcal{C}:z\mapsto i\frac{z-i}{z+i},\quad\text{which maps}\quad i\quad\text{to}\quad 0, (41)

    then the angular 11-form ω\omega has the following expression in 𝔻2{0}\mathbb{D}^{2}\setminus\{0\}

    𝒞ω=yx2+y2dx+xx2+y2dy.\mathcal{C}^{*}\omega=\frac{-y}{x^{2}+y^{2}}\,dx+\frac{x}{x^{2}+y^{2}}\,dy.
  2. (2)

    The angular form ω\omega is preserved by any rotation RθR^{\theta}. That is,

    (Rθ)ω=ω.(R^{\theta})^{*}\omega=\omega.

The following elementary and well-known lemma expresses the degree of a branched covering in terms of the angular form ω\omega.

Lemma 6.10.

Let γ:[0,1]2\gamma:[0,1]\to\mathbb{H}^{2}_{*} be a curve around the puncture generating π1(2)\pi_{1}(\mathbb{H}^{2}_{*}), and let h:22h:\mathbb{H}^{2}_{*}\to\mathbb{H}^{2}_{*} be a degree-nn cover branched on ii. Then,

hγω=2nπ.\int_{h\circ\gamma}\omega=2n\pi.
Proof.

Since hh is a degree-nn cover of 2\mathbb{H}^{2}_{*}, γhω=nγω\int_{\gamma}h^{*}\omega=n\int_{\gamma}\omega. Hence,

hγω=γhω=nγω.\int_{h\circ\gamma}\omega=\int_{\gamma}h^{*}\omega=n\int_{\gamma}\omega.

Now it is enough to take a parametrization of a representative of a loop around the puncture and compute γω\int_{\gamma}\omega. This is possible as the integral depends only on the homotopy class of γ\gamma. So we take γ(t)=R2πtc(r0)2\gamma(t)=R^{2\pi t}c(r_{0})\in\mathbb{H}^{2}_{*} (see (16)). We then obtain γω=2π\int_{\gamma}\omega=2\pi, which yields the desired conclusion. ∎

The next lemma gives an expression for a lift of a curve in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} to 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}

Lemma 6.11.

Let T:(0,+)××𝔸d𝕊3T:(0,+\infty)\times\mathbb{R}\times\mathbb{R}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} be the universal covering map of 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} defined in (21). Let c:𝔸d𝕊3c:\mathbb{R}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} be a curve in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} and let c~:𝔸d𝕊3~\widetilde{c}:\mathbb{R}\to\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}} be a lift of the curve cc such that c~(0)=(r0,x0,y0)\widetilde{c}(0)=(r_{0},x_{0},y_{0}). Then,

c~(t)=(dσ(i,c(t)i),x0+c([0,t])Val(ω),y0+c([0,t])Val¯(ω)),\widetilde{c}(t)=\left(d_{\sigma}(i,c(t)\cdot i),x_{0}+\int_{c([0,t])}\mathrm{Val}^{*}(\omega),y_{0}+\int_{c([0,t])}\mathrm{\overline{Val}}^{*}(\omega)\right),

where Val,Val¯:𝔸d𝕊32\mathrm{Val},\overline{\mathrm{Val}}:\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}\to\mathbb{H}^{2}_{*} are the valuations map defined by Val(A)=A(i)\mathrm{Val}(A)=A(i) and Val¯(A)=A1(i)\overline{\mathrm{Val}}(A)=A^{-1}(i).

Proof.

Assume that c~(t)=(r(t),θ(t),η(t))\widetilde{c}(t)=(r(t),\theta(t),\eta(t)), so that

c(t)=Rθ(t)A(r(t))Rη(t).c(t)=R^{\theta(t)}A(r(t))R^{-\eta(t)}.

The equality r(t)=dσ(i,c(t)i)r(t)=d_{\sigma}(i,c(t)\cdot i) follows easily. We now focus on deriving the formula for θ(t)\theta(t) using the angular form ω\omega. Observe that

Val(c(t))=π(r(t),θ(t)),\mathrm{Val}(c(t))=\pi(r(t),\theta(t)), (42)

where π:(0,+)×2\pi:(0,+\infty)\times\mathbb{R}\to\mathbb{H}^{2}_{*} is the map from (16). We compute,

c([0,t])Val(ω)\displaystyle\int_{c([0,t])}\mathrm{Val}^{*}(\omega) =0tωVal(c(s))(dc(s)Val(c(s)))ds\displaystyle=\int_{0}^{t}\omega_{\mathrm{Val}(c(s))}\left(\mathrm{d}_{c(s)}\mathrm{Val}(c^{\prime}(s))\right)\mathrm{d}s
=0tωπ(r(s),θ(s))(d(r(s),θ(s))π(r(s),θ(s)))ds\displaystyle=\int_{0}^{t}\omega_{\pi(r(s),\theta(s))}\left(\mathrm{d}_{(r(s),\theta(s))}\pi(r^{\prime}(s),\theta^{\prime}(s))\right)\mathrm{d}s
=0tdθ(r(s),θ(s))ds\displaystyle=\int_{0}^{t}\mathrm{d}\theta(r^{\prime}(s),\theta^{\prime}(s))\mathrm{d}s
=0tθ(s)ds\displaystyle=\int_{0}^{t}\theta^{\prime}(s)\mathrm{d}s
=θ(t)x0,\displaystyle=\theta(t)-x_{0},

where in the second equality, we used Equation (42), and in the third equality, we used Lemma 6.8. The proof for the third coordinate of c~\widetilde{c} is essentially the same. This completes the proof. ∎

The final lemma toward Proposition 6.7 is below. The proof follows line by line the argument of [Jan22, Lemma 3.5.3].

Lemma 6.12.

[Jan22, Lemma 3.5.3] Let γ1:2\gamma_{1}:\mathbb{R}\to\mathbb{H}^{2}_{*} and γ2:2\gamma_{2}:\mathbb{R}\to\mathbb{H}^{2}_{*} be two paths in 2\mathbb{H}^{2}_{*} satisfying γj(t+t0)=Rθ0γj(t)\gamma_{j}(t+t_{0})=R^{\theta_{0}}\gamma_{j}(t), for some t0,θ0t_{0},\theta_{0}\in\mathbb{R}. Assume that for all tt\in\mathbb{R}, the geodesic segment [γ1(t),γ2(t)][\gamma_{1}(t),\gamma_{2}(t)] does not contain ii. Then,

γ1([0,t0])ω=γ2([0,t0])ω.\int_{\gamma_{1}([0,t_{0}])}\omega=\int_{\gamma_{2}([0,t_{0}])}\omega.

We now turn to the proof of Proposition 6.7. Recall that α\alpha is a nontrivial loop in π1(𝔸d𝕊f,h)\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}) that is not homotopic to the fiber of \mathcal{F}, and that α\alpha projects to a loop γ\gamma around the puncture of 2\mathbb{H}^{2}_{*} that generates the π1\pi_{1}. Let α~\widetilde{\alpha} be a lift of α\alpha in 𝔸d𝕊f,h~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}. For every tt\in\mathbb{R}, we denote

A~t:=Dev~(α~(t))\widetilde{A}_{t}:=\widetilde{\text{Dev}}(\widetilde{\alpha}(t))

and

At:=Dev(α~(t))=T(Dev~(α~(t)))A_{t}:=\mathrm{Dev}(\widetilde{\alpha}(t))=T(\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(t)))

(see (39)). Consequently, we have the following.

  • By the construction of the manifold 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}, the path AtA_{t} in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*} has the property

    At(f(γ(t)))=h(γ(t))A_{t}(f(\gamma(t)))=h(\gamma(t))

    for every tt.

  • By definition of the action of the fundamental group, we have

    Dev~(α~(1))=Dev~(αα~(0))=H~(α)Dev~(α~(0)),\widetilde{\text{Dev}}(\widetilde{\alpha}(1))=\widetilde{\text{Dev}}(\alpha\cdot\widetilde{\alpha}(0))=\widetilde{H}(\alpha)\widetilde{\text{Dev}}(\widetilde{\alpha}(0)), (43)

    where H~(α)=φ(θ0,η0)\widetilde{H}(\alpha)=\varphi_{(\theta_{0},\eta_{0})} is the holonomy of α\alpha, see (22).

Before proving Proposition 6.7, we need a basic lemma on hyperbolic geometry. For each p2p\in\mathbb{H}^{2}_{*}, we denote by BpB_{p} the unique hyperbolic isometry that sends ii to pp and whose axis is the oriented geodesic joining ii to pp.

Lemma 6.13.

Let p,q2p,q\in\mathbb{H}^{2}_{*} be such that dσ(i,p)dσ(i,q)d_{\sigma}(i,p)\neq d_{\sigma}(i,q). Then, BpBq1B_{p}B_{q}^{-1} does not fix ii.

Proof.

By contradiction, assume that R:=BpBq1R:=B_{p}B_{q}^{-1} fixes ii. By definition, Rq=pRq=p, and hence dσ(i,q)=dσ(Ri,Rp)=dσ(i,p)d_{\sigma}(i,q)=d_{\sigma}(Ri,Rp)=d_{\sigma}(i,p), which is a contradiction. ∎

Proof of Proposition 6.7.

In this proof, we take explicit γ\gamma and α\alpha in order to make computations. We set γ(t)=re2iπt\gamma(t)=re^{2i\pi t} and

At=Bh(γ(t))Bf(γ(t))1,A_{t}=B_{h(\gamma(t))}B^{-1}_{f(\gamma(t))},

where BpB_{p} the unique hyperbolic isometry that sends ii to pp and whose axis is the oriented geodesic joining ii to pp. Since γ(t+1)=γ(t)\gamma(t+1)=\gamma(t), it follows that At+1=AtA_{t+1}=A_{t}, and therefore α(t)=[At]\alpha(t)=[A_{t}] is a loop in 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} that is not homotopic to a fiber of the fibration :𝔸d𝕊f,h2\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{H}^{2}_{*}.

The goal is now to compute the holonomy of the loop α\alpha. Let α~\widetilde{\alpha} be the lift of α\alpha to 𝔸d𝕊f,h~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}. By definition, we have

Dev(α~(t))=𝒟(α(t))=At,\mathrm{Dev}(\widetilde{\alpha}(t))=\mathcal{D}(\alpha(t))=A_{t},

and so Dev~(α~(t))\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(t)) is the lift to 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}} of Dev(α~(t))\mathrm{Dev}(\widetilde{\alpha}(t)). Consider θ0,η0\theta_{0},\eta_{0}\in\mathbb{R} such that Hol~(α)=φ(θ0,η0)\widetilde{\mathrm{Hol}}(\alpha)=\varphi_{(\theta_{0},\eta_{0})} (see notation in (22)). Thus, by Proposition 6.11, we have

Dev~(α~(t))=(d2(i,Dev(α~(t))i),x0+Devα~([0,t])Val(ω),y0+Devα~([0,t])Val¯(ω)),\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(t))=\left(\mathrm{d}_{\mathbb{H}^{2}}\!\left(i,\mathrm{Dev}(\widetilde{\alpha}(t))\cdot i\right),\ x_{0}+\int_{\mathrm{Dev\circ\widetilde{\alpha}}([0,t])}\mathrm{Val}^{*}(\omega),\ y_{0}+\int_{\mathrm{Dev}\circ\widetilde{\alpha}([0,t])}\overline{\mathrm{Val}}^{*}(\omega)\right), (44)

for some x0,y0x_{0},y_{0}\in\mathbb{R}. This leads to

Dev~(α~(1))\displaystyle\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(1)) =(dσ(i,Dev(α~(1))i),x0+Devα~([0,1])Val(ω),y0+Devα~([0,1])Val¯(ω)).\displaystyle=\left(\mathrm{d}_{\sigma}\!\left(i,\mathrm{Dev}(\widetilde{\alpha}(1))\cdot i\right),\ x_{0}+\int_{\mathrm{Dev\circ\widetilde{\alpha}}([0,1])}\mathrm{Val}^{*}(\omega),\ y_{0}+\int_{\mathrm{Dev}\circ\widetilde{\alpha}([0,1])}\overline{\mathrm{Val}}^{*}(\omega)\right). (45)

We claim that

Devα~([0,1])Val(ω)=2nπ.\int_{\mathrm{Dev}\circ\widetilde{\alpha}([0,1])}\mathrm{Val}^{*}(\omega)=2n\pi. (46)

To prove the claim, let us consider the paths

γ1(t):=At(i),γ2(t):=h(γ(t))=At(f(γ(t))).\gamma_{1}(t):=A_{t}(i),\qquad\gamma_{2}(t):=h(\gamma(t))=A_{t}(f(\gamma(t))).

Clearly γ2\gamma_{2} is a loop in 2\mathbb{H}^{2}_{*}, and the same holds for γ1\gamma_{1} by Lemma 6.13, which ensures that γ1(t)2\gamma_{1}(t)\in\mathbb{H}^{2}_{*} for all tt\in\mathbb{R}. Next, we want to apply Lemma 6.12. The curves γj\gamma_{j} satisfy γj(t+1)=γj(t)\gamma_{j}(t+1)=\gamma_{j}(t) for j=1,2j=1,2 (because γ(t+1)=γ(t)\gamma(t+1)=\gamma(t) and At+1=AtA_{t+1}=A_{t}). By Lemma 6.4, we have

dσ(γ1(t),γ2(t))=dσ(i,f(γ(t)))<dσ(i,h(γ(t)))=dσ(i,γ2(t)).d_{\sigma}(\gamma_{1}(t),\gamma_{2}(t))=d_{\sigma}(i,f(\gamma(t)))<d_{\sigma}(i,h(\gamma(t)))=d_{\sigma}(i,\gamma_{2}(t)).

This implies in particular that ii does not belong to the geodesic segment [γ1(t),γ2(t)][\gamma_{1}(t),\gamma_{2}(t)]. Therefore,

Devα~([0,1])Val(ω)=γ2ω=hγω=2nπ,\int_{\mathrm{Dev}\circ\widetilde{\alpha}([0,1])}\mathrm{Val}^{*}(\omega)=\int_{\gamma_{2}}\omega=\int_{h\circ\gamma}\omega=2n\pi, (47)

where the last equality follows from the fact that hh is a branched covering of degree nn (see Lemma 6.10). Next, since

Dev~(α~(1))=φ(θ0,η0)(Dev~(α~(0))),\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(1))=\varphi_{(\theta_{0},\eta_{0})}\left(\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(0))\right),

it follows from (44) that

x0+Devα~([0,1])Val(ω)=x0+θ0.x_{0}+\int_{\mathrm{Dev}\circ\widetilde{\alpha}([0,1])}\mathrm{Val}^{*}(\omega)=x_{0}+\theta_{0}.

Combined with (47), we obtain θ0=2nπ,\theta_{0}=2n\pi, which finishes the proof of the claim.

The remaining part is to compute

Devα~([0,1])Val¯(ω),\int_{\mathrm{Dev}\circ\widetilde{\alpha}([0,1])}\overline{\mathrm{Val}}^{*}(\omega),

which is equal to γ3ω\int_{\gamma_{3}}\omega, where γ3(t):=At1i\gamma_{3}(t):=A_{t}^{-1}\cdot i. Since γ3\gamma_{3} is a closed loop, it follows from classical degree theory and the calculation at the end of Lemma 6.10 that this last integral equals 2kπ2k\pi for some kk\in\mathbb{Z}. This concludes the proof, as

Dev~(α~(1))=H~(α)Dev~(α~(0))\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(1))=\widetilde{H}(\alpha)\,\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(0))

and

Dev~(α~(1))\displaystyle\widetilde{\mathrm{Dev}}(\widetilde{\alpha}(1)) =(dσ(i,Dev(α~(0))i),x0+2nπ,y0+2kπ).\displaystyle=\left(\mathrm{d}_{\sigma}\!\left(i,\mathrm{Dev}(\widetilde{\alpha}(0))\cdot i\right),\ x_{0}+2n\pi,\ y_{0}+2k\pi\right). (48)

6.3. Holonomy of the fiber in the fundamental example

The main goal of this subsection is to prove the following Proposition.

Proposition 6.14.

Let β\beta be a generator of π1(𝔸d𝕊f,h)\pi_{1}(\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}) that is homotopic to a future-directed timelike geodesic fiber of :𝔸d𝕊f,h2\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{H}^{2}_{*} (see (40)). Then,

Hol~(β)=φ(0,2π).\widetilde{\mathrm{Hol}}(\beta)=\varphi_{(0,2\pi)}.

The proof of this proposition proceeds through the study of the stabilizers of timelike geodesics in 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}. For x,y2x,y\in\mathbb{H}^{2}, let x,y\ell_{x,y} be the timelike geodesic in 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} defined in (15). If x,y\ell_{x,y} belongs to 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}, we denote by ~x,y\widetilde{\ell}_{x,y} its lift to 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}. We denote by Stab(x,y)\mathrm{Stab}(\ell_{x,y}) the stabilizer of x,y\ell_{x,y} in Isom(𝔸d𝕊3)\mathrm{Isom}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}), and by Stab(~x,y)\mathrm{Stab}(\widetilde{\ell}_{x,y}) the stabilizer of ~x,y\widetilde{\ell}_{x,y} in Isom(𝔸d𝕊3~)\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}). That is,

Stab(x,y):={φIsom(𝔸d𝕊3)φ(x,y)=x,y}\mathrm{Stab}(\ell_{x,y}):=\left\{\varphi\in\mathrm{Isom}(\mathbb{A}\mathrm{d}\mathbb{S}^{3})\ \mid\ \varphi(\ell_{x,y})=\ell_{x,y}\right\}

and

Stab(~x,y):={φIsom(𝔸d𝕊3~)φ(~x,y)=~x,y}.\mathrm{Stab}(\widetilde{\ell}_{x,y}):=\left\{\varphi\in\mathrm{Isom}(\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}})\ \mid\ \varphi(\widetilde{\ell}_{x,y})=\widetilde{\ell}_{x,y}\right\}.

We have the following lemma.

Lemma 6.15.

[Jan22, Propositions 2.2.5 and 2.2.6] Let x,y2x,y\in\mathbb{H}^{2} be such that dσ(i,x)<dσ(i,y)d_{\sigma}(i,x)<d_{\sigma}(i,y). Consider the timelike geodesic x,y\ell_{x,y}, which is disjoint from i,i\ell_{i,i}, and let x,y~\widetilde{\ell_{x,y}} be its lift in 𝔸d𝕊3~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}. Then,

Stab(x,y~)={0}×2π.\mathrm{Stab}(\widetilde{\ell_{x,y}})=\{0\}\times 2\pi\mathbb{Z}.
Proof of Proposition 6.14.

Let x2x\in\mathbb{H}^{2}_{*}. By Lemma 6.4,

dσ(i,f(x))<dσ(i,h(x))d_{\sigma}(i,f(x))<d_{\sigma}(i,h(x))

and the timelike geodesics h(x),f(x)\ell_{h(x),f(x)} and i,i\ell_{i,i} are distinct. In particular, h(x),f(x)𝔸d𝕊3\ell_{h(x),f(x)}\subset\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}. Let β\beta be a generator of an oriented fiber of the fibration :𝔸d𝕊f,h2\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{H}^{2}_{*}. We may choose a representative of β\beta so that 𝒟(β)\mathcal{D}(\beta) is the oriented timelike geodesic h(x),f(x)\ell_{h(x),f(x)} (see (38)). Let β~\widetilde{\beta} be the lift of β\beta to 𝔸d𝕊f,h~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}} that is preserved by the deck transformation induced by β\beta. By the equivariance of the developing map Dev~:𝔸d𝕊f,h~𝔸d𝕊3~\widetilde{\mathrm{Dev}}:\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{3}_{*}}, the holonomy Hol~(β)\widetilde{\mathrm{Hol}}(\beta) preserves the timelike geodesic Dev~(β~)=~h(x),f(x).\widetilde{\mathrm{Dev}}(\widetilde{\beta})=\widetilde{\ell}_{h(x),f(x)}. Hence, by Lemma 6.15, there exists mm\in\mathbb{Z} such that

Hol~(β)=(0,2mπ).\widetilde{\mathrm{Hol}}(\beta)=(0,2m\pi). (49)

On the other hand, T(Hol~(β))T_{*}(\widetilde{\mathrm{Hol}}(\beta)) induces a map from

𝕊1h(x),f(x)𝕊1h(x),f(x),\mathbb{S}^{1}\cong\ell_{h(x),f(x)}\to\mathbb{S}^{1}\cong\ell_{h(x),f(x)},

which induces an isomorphism \mathbb{Z}\to\mathbb{Z} on the level of the fundamental group. This is because T(Hol~(β))T_{*}(\widetilde{\mathrm{Hol}}(\beta)) is an isometry and, in particular, a homeomorphism. The isomorphism \mathbb{Z}\to\mathbb{Z} is given by pmpp\mapsto mp, where mm is the integer in (49). Thus, m=±1m=\pm 1. However, m=1m=-1 is excluded since T(Hol~(β))T_{*}(\widetilde{\mathrm{Hol}}(\beta)) is a time-orientation-preserving isometry. This concludes the proof. ∎

Returning to the original goal of this subsection 6.1, by combining Lemma 6.6 with Propositions 6.7 and 6.14, we obtain the proof of Proposition 6.1.

6.4. Proof of Theorem B

In this final subsection, we prove Theorem B.

Proof of Theorem B.

Let Σ\Sigma be a hyperbolic surface with fundamental group Γ\Gamma and let f,h:Σ~2f,h:\widetilde{\Sigma}\to\mathbb{H}^{2} be smooth maps that are equivariant with respect to representations ρ,j:ΓPSL(2,)\rho,j:\Gamma\to\mathrm{PSL}(2,\mathbb{R}), as in Theorem B. As in Section 5, let CΣ~C\subset\widetilde{\Sigma} be the singular set of the branched immersion hh, and set C0=C/ΓC_{0}=C/\Gamma. Around each point of CC, the branched immersion hh is a degree-nn branched covering.

Due to Theorem 5.3 and Lemma 5.9, the only remaining point to verify in Theorem B is that 𝔸d𝕊f,h/Γj,ρ\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}/\Gamma_{j,\rho} is a branched AdS manifold.

Fix a point x0Cx_{0}\in C, and let UΣ~U\subset\widetilde{\Sigma} be a small disc centered at x0x_{0} on which hh is a branched covering of degree nn. We may identify UU with a small disc in 2\mathbb{H}^{2} centered at ii; by applying isometries, we can further assume that x0=ix_{0}=i and f(i)=h(i)=if(i)=h(i)=i. Denote by f¯,h¯:U2\overline{f},\overline{h}:U\to\mathbb{H}^{2}_{*} the restrictions of ff and hh respectively. Applying Proposition 6.1 together with Remark 6.3, we obtain that 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} is a branched anti-de Sitter manifold that fibers over UU.

Since the construction of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} is local, the behavior of 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} around the fiber [i,i][\ell_{i,i}] is the same as that of 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}}. In fact, we will show that 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} can be identified with an open subset of the quotient f,h=𝔸d𝕊f,h/Γj,ρ\mathcal{M}^{f,h}_{*}=\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}/\Gamma_{j,\rho}. Indeed, one can see without difficulties that 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} is an open set inside 𝔸d𝕊f,h\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}. Considering the natural projection p:𝔸d𝕊f,hf,h,p:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathcal{M}^{f,h}_{*}, we claim that, upon shrinking the neighborhood UU, we may assume that p:𝔸d𝕊f¯,h¯p(𝔸d𝕊f¯,h¯)f,hp:\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}}\to p(\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}})\subset\mathcal{M}^{f,h}_{*} is injective.

To prove this, we use the fact that Γ\Gamma acts properly discontinuously on Σ~\widetilde{\Sigma} to shrink the open set UU so that γUU=\gamma U\cap U=\emptyset for all γe\gamma\neq e. Now, by contradiction, if the restriction of pp to 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} were not injective, then there would exist A𝔸d𝕊f¯,h¯A\in\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} and γΓ\gamma\in\Gamma such that γA𝔸d𝕊f¯,h¯\gamma\cdot A\in\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}}. Using the equivariant fibration :𝔸d𝕊f,hΣ~\mathcal{F}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}\to\widetilde{\Sigma}, we deduce that (A)U\mathcal{F}(A)\in U and γ(A)U\gamma\mathcal{F}(A)\in U, and hence γ=e\gamma=e. This shows that p:𝔸d𝕊f¯,h¯p(𝔸d𝕊f¯,h¯)p:\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}}\to p(\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}}) is injective.

Therefore, 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} is identified via pp with an open subset of f,h=𝔸d𝕊f,h/Γj,ρ\mathcal{M}^{f,h}_{*}=\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}/\Gamma_{j,\rho}. By repeating this argument around each point of the singular locus of hh, we conclude that f,h\mathcal{M}^{f,h}_{*} is a branched anti-de Sitter manifold, since 𝔸d𝕊f¯,h¯\mathbb{A}\mathrm{d}\mathbb{S}_{*}^{\overline{f},\overline{h}} is one.

Finally, we discuss the holonomy representation of f,h\mathcal{M}^{f,h}_{*}. This is computed by studying the developing map from the universal cover 𝔸d𝕊f,h~\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}} to 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} and how it transforms under the action of π1(f,h)\pi_{1}(\mathcal{M}^{f,h}_{*}). This developing map is realized explicitly as the projection 𝔸d𝕊f,h~𝔸d𝕊f,h\widetilde{\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}}\to\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*} post-composed with the map :𝔸d𝕊f,h𝔸d𝕊3\mathcal{I}:\mathbb{A}\mathrm{d}\mathbb{S}^{f,h}_{*}\to\mathbb{A}\mathrm{d}\mathbb{S}^{3} defined in Section 5.1 (see also equations (38) and (39)). Since :f,hΣC0\mathcal{F}:\mathcal{M}^{f,h}*\to\Sigma\setminus C_{0} is a circle bundle, the fundamental group π1(f,h)\pi_{1}(\mathcal{M}^{f,h}_{*}) fits into the (splitting) exact sequence

0π1(𝕊1)π1(f,h)π1(Σ\C0)0.0\to\pi_{1}(\mathbb{S}^{1})\to\pi_{1}(\mathcal{M}^{f,h}_{*})\to\pi_{1}(\Sigma\backslash C_{0})\to 0.

By the formula for Hol~(β)\widetilde{\textrm{Hol}}(\beta) from Proposition 6.1, Hol~(β)\widetilde{\textrm{Hol}}(\beta) acts trivially on 𝔸d𝕊3\mathbb{A}\mathrm{d}\mathbb{S}^{3} and hence the holonomy factors through the map π1(f,h)π1(Σ\C0)\pi_{1}(\mathcal{M}^{f,h}_{*})\to\pi_{1}(\Sigma\backslash C_{0}). By the formula for Hol~(α)\widetilde{\textrm{Hol}}(\alpha) from Proposition 6.1, applied around each of the singular points, we see that the holonomy further factors through the map π1(Σ\C0)π1(Σ)\pi_{1}(\Sigma\backslash C_{0})\to\pi_{1}(\Sigma) induced by inclusion. From the definition of \mathcal{I} in Section 5.1 and the definition of the action of Γj,ρ\Gamma_{j,\rho}, i.e., equation (37), we see that the induced homomorphism π1(Σ)Isom0(𝔸d𝕊3)\pi_{1}(\Sigma)\to\textrm{Isom}_{0}(\mathbb{A}\mathrm{d}\mathbb{S}^{3}) is (j,ρ).(j,\rho). This completes the proof. ∎

References

  • [AY19] Sébastien Alvarez and Jiagang Yang. Physical measures for the geodesic flow tangent to a transversally conformal foliation. Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 36(1):27–51, 2019.
  • [BBDH21] Indranil Biswas, Steven Bradlow, Sorin Dumitrescu, and Sebastian Heller. Uniformization of branched surfaces and higgs bundles. International Journal of Mathematics, 32(13):2150096, 2021.
  • [BG25] Pabitra Barman and Subhojoy Gupta. Dominating surface-group representations via fock–goncharov coordinates. Geometriae Dedicata, 219, 01 2025.
  • [BM12] Thierry Barbot and Catherine Meusburger. Particles with spin in stationary flat spacetimes. Geom. Dedicata, 161:23–50, 2012.
  • [BS20] Francesco Bonsante and Andrea Seppi. Anti-de Sitter geometry and Teichmüller theory. In In the tradition of Thurston. Geometry and topology, pages 545–643. Cham: Springer, 2020.
  • [Cor88] Kevin Corlette. Flat GG-bundles with canonical metrics. J. Differential Geom., 28(3):361–382, 1988.
  • [CTT19] Brian Collier, Nicolas Tholozan, and Jérémy Toulisse. The geometry of maximal representations of surface groups into SO0(2,n)\mathrm{SO}_{0}(2,n). Duke Math. J., 168(15):2873–2949, 2019.
  • [DL20] Song Dai and Qiongling Li. On cyclic Higgs bundles. Mathematische Annalen, 376(3–4):1225–1260, November 2020.
  • [DL22] Song Dai and Qiongling Li. Domination results in nn‐fuchsian fibers in the moduli space of higgs bundles. Proceedings of the London Mathematical Society, 124:427–477, 03 2022.
  • [Don87] S. K. Donaldson. Twisted harmonic maps and the self-duality equations. Proc. London Math. Soc. (3), 55(1):127–131, 1987.
  • [DT16] Bertrand Deroin and Nicolas Tholozan. Dominating surface group representations by Fuchsian ones. Int. Math. Res. Not., 2016(13):4145–4166, 2016.
  • [Far21] Gianluca Faraco. Geometrisation of purely hyperbolic representations in PSL2\mathrm{PSL}_{2}\mathbb{R}. Adv. Geom., 21(1):99–108, 2021.
  • [GK17] François Guéritaud and Fanny Kassel. Maximally stretched laminations on geometrically finite hyperbolic manifolds. Geom. Topol., 21(2):693–840, 2017.
  • [GKW15] François Guéritaud, Fanny Kassel, and Maxime Wolff. Compact anti-de Sitter 3-manifolds and folded hyperbolic structures on surfaces. Pac. J. Math., 275(2):325–359, 2015.
  • [Gol80] William M. Goldman. Discontinuous Groups and the Euler Class. PhD thesis, University of California, Berkeley, Ann Arbor, MI, 1980. Thesis (Ph.D.)–University of California, Berkeley.
  • [GPG24] Oscar García-Prada and Miguel González. Cyclic higgs bundles and the toledo invariant, 2024.
  • [GS22] Subhojoy Gupta and Weixu Su. Dominating surface-group representations into PSL(2,)\mathrm{PSL}(2,\mathbb{C}) in the relative representation variety. manuscripta mathematica, 172, 12 2022.
  • [Hit87] Nigel J. Hitchin. The Selfduality equations on a Riemann surface. Proc. Lond. Math. Soc., 55:59–131, 1987.
  • [Jan22] Agnese Janigro. Compact 3-dimensional Anti-de Sitter manifolds with spin-cone singularities. PhD thesis, Università degli Studi di Milano-Bicocca, 2022.
  • [Kas10] Fanny Kassel. Quotients compacts des groupes ultramétriques de rang un. Ann. Inst. Fourier (Grenoble), 60(5):1741–1786, 2010.
  • [KR85] Ravi S. Kulkarni and Frank Raymond. 33-dimensional Lorentz space-forms and Seifert fiber spaces. J. Differential Geom., 21(2):231–268, 1985.
  • [MB23] Florestan Martin-Baillon. Dominating CAT(1)(-1) surface group representations by Fuchsian ones. Geom. Dedicata, 217(3):19, 2023. Id/No 45.
  • [McI25] Ian McIntosh. The geometric toda equations for noncompact symmetric spaces. Differential Geometry and its Applications, 99:102249, 2025.
  • [Min87] C. Minda. The strong form of ahlfors’ lemma. Rocky Mountain Journal of Mathematics, 17, 09 1987.
  • [Sag23] Nathaniel Sagman. Infinite energy equivariant harmonic maps, domination, and anti-deăSitter 33-manifolds. Journal of Differential Geometry, 124(3):553 – 598, 2023.
  • [Sag24] Nathaniel Sagman. Almost strict domination and anti-de Sitter 3-manifolds. J. Topol., 17(1):51, 2024. Id/No e12323.
  • [Sal00] François Salein. Variétés anti-de Sitter de dimension 3 exotiques. Ann. Inst. Fourier (Grenoble), 50(1):257–284, 2000.
  • [Sam78] J. H. Sampson. Some properties and applications of harmonic mappings. Ann. Sci. École Norm. Sup. (4), 11(2):211–228, 1978.
  • [SG24] Pedro M. Silva and Peter B. Gothen. The conformal limit and projective structures. Int. Math. Res. Not., 2024(16):11812–11831, 2024.
  • [ST25] Nathaniel Sagman and Ognjen Tošić. On Hitchin’s equations for cyclic G-higgs bundles. Advances in Mathematics, 482:110599, 2025.
  • [SY97] R. Schoen and S. T. Yau. Lectures on harmonic maps. Conference Proceedings and Lecture Notes in Geometry and Topology, II. International Press, Cambridge, MA, 1997.
  • [Tan94] Ser Peow Tan. Branched P1\mathbb{C}{P}^{1}-structures on surfaces with prescribed real holonomy. Math. Ann., 300(4):649–667, 1994.
  • [Tho15] Nicolas Tholozan. Entropy of hilbert metrics and length spectrum of Hitchin representations in PSL(3,)\mathrm{PSL}(3,\mathbb{R}). Duke Mathematical Journal, 166, 06 2015.
  • [Tho17] Nicolas Tholozan. Dominating surface group representations and deforming closed anti-de Sitter 3-manifolds. Geom. Topol., 21(1):193–214, 2017.
  • [Tro91] Marc Troyanov. Prescribing curvature on compact surfaces with conical singularities. Trans. Amer. Math. Soc., 324(2):793–821, 1991.
  • [Wol89] Michael Wolf. The Teichmüller theory of harmonic maps. J. Differential Geom., 29(2):449–479, 1989.
  • [Woo77] John C. Wood. Singularities of harmonic maps and applications of the Gauss-Bonnet formula. Amer. J. Math., 99(6):1329–1344, 1977.