Domination between non-Fuchsian representations and anti-de Sitter geometry
Abstract.
Motivated by work of various authors on domination between surface group representations, harmonic maps, and -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 -manifold with singularities, and we construct large families of branched anti-de Sitter -manifolds.
Contents
1. Introduction
Let be an oriented surface with fundamental group and let be the hyperbolic space of constant curvature . As is standard, the group acts on by Möbius transformations. We say that a representation strictly dominates another representation if there exists a map with equivariance , , that contracts hyperbolic distance. Formally, the latter condition means that there exists such that for all
Domination has attracted considerable attention and has been studied for maps between spaces other than . When 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 -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 -dimensional anti-de Sitter space 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 is closed, a strictly dominating pair is equivalent to a closed -manifold locally modeled on (more on this below).
A representation is called Fuchsian if it is discrete and faithful; equivalently, it is the holonomy of a hyperbolic structure on . 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 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 as just -equivariant (or, if is unspecified, equivariant) if for all , , where the first -action is via deck transformations. Toward the definition below, we note that, given an equivariant map , the pullback metric descends from to . For equivariant maps and , we write to mean that for every unit tangent vector , and that the inequality is strict when is positive definite. We also recall that when is closed and equipped with a Riemann surface structure , and is reductive, there exists a -equivariant harmonic map ; the map is not always unique, but the pullback metric is (see Section 2.1).
Definition 1.1.
Let be representations.
-
(1)
We say that dominates in the pullback sense if there exist maps and which are respectively - and -equivariant, such that
-
(2)
When is closed and both and are reductive, we say that dominates in the harmonic maps sense if there exists a Riemann surface structure on such that the -equivariant and -equivariant harmonic maps satisfy
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 is a Fuchsian representation of a closed surface group. The first condition is a little flimsy–one has a lot of freedom with and –while the second one is more rigid.
1.1. Main results
In this paper, the two main questions we address are the following.
-
(1)
Given , can we dominate it in the pullback or harmonic maps sense by a non-Fuchsian representation ?
-
(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 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 and a representation is Fuchsian precisely when , see [Gol80]. Note that when , the representation is reductive and the -equivariant harmonic map is unique. Our first domination result concerns the case where this harmonic map is a branched immersion. Recall that a map between surfaces has a branch point at a point if one can find local coordinates and centered at and respectively on which takes the form . We say that is a branched immersion if it is an immersion apart from at a (discrete) set of branch points.
Theorem A.
Let be a closed Riemann surface of genus with fundamental group and let be a representation with such that the unique equivariant harmonic map is a branched immersion. Then, for every between and , we can choose a representation with and equivariant harmonic map such that .
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 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 (see [Tan94] for explanation). The equivariant harmonic map associated with is also a branched immersion, so Theorem A can be alternatively cast as a domination result between hyperbolic cone structures. The choices of of Euler number are parametrized by effective divisors of degree that dominate the branching divisor of . For 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 where 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 . The three-dimensional anti-de Sitter space is a Lorentzian space form of constant negative sectional curvature, and it can be modeled on with the Killing metric. In this model, the group of space and time-orientation preserving isometries is , acting by right and left multiplication
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 .
For surface group representations , we define by
| (1) |
Kulkarni and Raymond proved that every group acting properly discontinuously on identifies with some , and that (up to switching factors) must be Fuchsian [KR85]. The quotient is a circle bundle over , 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 is a closed surface of genus at least , acts properly and discontinuously on if and only if strictly dominates (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 -manifolds that fiber via timelike geodesics over hyperbolic cone surfaces. Janigro defined the notion of a spin-cone AdS -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 -manifolds. We extend Janigro’s work by showing that domination in the pullback sense gives rise to a singular AdS -manifold (see Theorem 5.3). This extension is obtained by adding a singular set to the AdS -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 , 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 is a branched immersion and analyze the resulting singularities in detail. We arrive at the following theorem.
Theorem B.
Let be a hyperbolic surface with fundamental group and let be two representations. Assume that there exist a –equivariant map and a –equivariant branched immersion such that . Denote by the singular set of and . Then there exists a three–manifold , topologically a solid torus, with the following properties.
-
(1)
The group admits a properly discontinuous action on .
-
(2)
There exists a continuous map that is equivariant with respect to the action of on and the action of on . Moreover, the fibers of are topological circles.
-
(3)
Let and set . Then is an anti-de Sitter manifold, and the restriction of to is an -principal bundle over with timelike geodesic fibers.
-
(4)
The quotient is a branched anti-de Sitter manifold with singular locus . The holonomy representation factors
where the first map is induced by , the second is induced by inclusion, and the third map is .
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 is a global diffeomorphism, both and coincide with the anti–de Sitter space , and we thus recover previously known results about closed AdS -manifolds.
Combining Theorems A and B, we obtain the following result, which provides a large class of examples of branched anti-de Sitter -manifolds.
Corollary B.
Let be a closed Riemann surface of genus with fundamental group and let be a representation with such that the unique equivariant harmonic map is a branched immersion. Then, for every between and , we can choose a representation , , with equivariant map such that the data determines a branched anti-de Sitter -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 -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 is not a branched immersion but is), so Corollary B can also be expanded.
Remark 1.2.
If one really wanted to construct a lot of spin-cone AdS -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 be a closed Riemann surface of genus at least with fundamental group and let be a non-Fuchsian reductive representation with -equivariant harmonic map . Every equivariant harmonic map from determines a holomorphic quadratic differential on called the Hopf differential. By [Wol89], there exists a unique Fuchsian representation together with an equivariant harmonic diffeomorphism with the same Hopf differential as . In [DT16], it is shown that
The map is -equivariant and strictly contracting, and hence shows that strictly dominates . 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 be reductive representations with Fuchsian. Then strictly dominates if and only if there exists a unique Riemann surface structure together with equivariant harmonic maps and for and respectively with the same Hopf differential and such that .
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 (up to translation) is equivalent to a pair consisting of a holomorphic quadratic differential on , the Hopf differential of , and an effective divisor on satisfying certain conditions. When is not zero, is the divisor of the square root of the holomorphic energy of .
See Proposition 2.7 for the precise statement (which includes the case ). When , one recovers Wolf’s parametrization of Teichmüller space [Wol89], and when , 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 and do we have domination? The problems turns out to be quite delicate, but we find a complete solution when 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 and be divisors on . We write to mean that for all and that there exists such that .
Proposition 1.6.
An equivariant harmonic map associated with a pair , , is a branched immersion if and only if or .
When , the harmonic map is weakly conformal and automatically a branched immersion. The conditions and 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 be equivariant harmonic maps with holomorphic data and respectively, with . Assuming is a branched immersion, then
Corollary C.
Let be equivariant harmonic maps with holomorphic data and respectively, with . Assuming is a branched immersion, then
The assumption on the degrees is not necessary, but just keeps the statement cleaner. If , then we can flip back to the case by using an outer automorphism of the fundamental group. A harmonic map with data becomes a map with data , and the pullback metric is unaffected, and thus the condition becomes . Similar if It is worth noting that if is not a branched immersion, the condition 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 with the same Hopf differential, the map determines a maximal surface (zero mean curvature at immersed points). The domination condition says that is a spacelike immersion off the singular set of .
Remark 1.8.
In the works [DT16], [Tho17], [Sag23], [Sag24], the main domain results concern pairs where is an action by isometries on a Hadamard manifold and is a Fuchsian representation to 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 , 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 . That being said, after Theorem C, it is natural to inquire about the equality case 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 , there exists a Riemann surface with fundamental group such that the following holds.
-
(1)
There exist representations with equivariant harmonic maps respectively, such that and .
-
(2)
There exists a representation with equivariant harmonic map with the property that there exist no pairs consisting of a non-Fuchsian representation and an equivariant harmonic map with the same Hopf differential such that .
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 such that, for all , where 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 , with the space replaced with a Riemannian symmetric space of . For generalizations of [DT16], with Fuchsian representations replaced with Hitchin representations to 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 -Higgs bundles, which are equivalent to solutions to affine Toda equations (see [ST25]). There is even a Toledo number for cyclic -Higgs bundles, which generalizes the Euler number (see [GPG24]).
1.3.3. More on singular AdS -manifolds
Our constructions in Section 5 (in particular, Theorem 5.3) allow the construction of general singular AdS -manifolds from dominating pairs, rather than only branched AdS -manifolds arising from branched immersions. By developing more exotic examples of singular AdS -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 -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 . 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 -manifolds, and in Section 5, we construct singular AdS -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 -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 be a closed oriented surface of genus at least and let be a Riemann surface structure on with universal cover . Let be a Riemannian manifold and let be a map. The derivative defines a section of the endomorphism bundle . Complexifying , let be the decomposition into and parts. We denote by the connection on induced by the Levi-Civita connection of , and we also use for its extension to -valued forms. The definition below depends on the Riemann surface structure and the metric
Definition 2.1.
The map is harmonic if
In this paper, we are concerned with equivariant harmonic maps from to the space , where is the -dimensional upper half-plane and is a hyperbolic metric of constant curvature . The starting point is the existence theorem of Donaldson [Don87] and Corlette [Cor88]. A representation is irreducible if it is not contained in a parabolic subgroup, and is reductive if the Zariski closure of is a reductive group. Geometrically, and perhaps more intuitively, is irreducible if the induced action of on the Gromov boundary has no global fixed point, and is reductive if preserves a geodesic in
Theorem 2.2 (Donaldson, Corlette).
Let be a Riemann surface structure on and let be a representation, acting by isometries on Then there exists a -equivariant harmonic map if and only if is reductive.
By Sampson’s argument in [Sam78, Theorem 3], if is irreducible, then is unique. When is just reductive and not irreducible, again by [Sam78, Theorem 3], there is a -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 be a Riemann surface as above and fix a metric on that’s compatible with the Riemann surface structure. We use as well to denote the lift to a metric on the universal cover .
Keeping things general for now, let be an action by isometries and let be a -equivariant map. The (possibly degenerate) pullback metric extends bilinearly to the complexified tangent bundle of and then decomposes into types as
| (2) |
We write where is called the energy density function. As well, the quadratic differential is called the Hopf differential of (note also that ). Since is -equivariant, , and are invariant under the action of and hence descend to . The formula (2) is rewritten as
| (3) |
The significance of stems from the fact that the equation can be seen as the Euler-Lagrange equation for the Dirichlet energy
where is the area form of (note does not depend on the choice of ). That is, harmonic maps are equivalently critical points of . As for the Hopf differential , it is holomorphic when is harmonic, and when the target has dimension at most so for example when the converse holds as well.
We now set . In this special case, there are more analytic quantities to probe the harmonic map. Working in local coordinates, where , , we set
The function is called the holomorphic energy and is called the anti-holomorphic energy. They satisfy
and the vanishing of the Jacobian of is equivalent to the vanishing of the function
When is harmonic, the functions and are either identically zero or have isolated zeros and satisfy the Bochner formulae: away from the zeros of
| (4) |
where and are the sectional curvatures of and , respectively (so ), and away from the zeros of
(see [SY97, Section 1.7]). Note that the Bochner formula for can be seen as a consequence of the Bochner formula for and the equation . Note that for a non-irreducible representation, again using that harmonic maps are related by isometries, is independent of the choice of harmonic map.
We end this subsection with a discussion on Euler numbers. The Euler number of a representation denoted can be defined in many ways; for instance, using characteristic classes, or as an obstruction to lifting to . Here, we give the most naive definition:
| (5) |
where is any -equivariant map, and is the area form of (which might be degenerate). By an application of Stoke’s theorem, the integral above indeed depends only on and not on . It is also useful to recall the equality of -forms
| (6) |
The following lemma allows us to make Definition 2.4 below.
Lemma 2.3.
Let be a Riemann surface structure on and let be a reductive representation with equivariant harmonic map . If , then cannot be identically zero.
Proof.
Definition 2.4.
Let be a Riemann surface structure on and let be a representation with and equivariant harmonic map . The divisor of , denoted is the divisor of the square root of the holomorphic energy of .
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 be reductive with . Then
If we reverse the orientation of or precompose with an outer automorphism of , then the Euler numbers flip sign. For this reason, we often restrict to the case . For , one has results analogous to above, with replaced with
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 of a holomorphic line bundle on , we denote the divisor by Let be the canonical bundle of so that is the space of holomorphic quadratic differentials on .
Definition 2.6.
Let be the set of pairs where is an effective divisor on with and , such that, if , then (as functions). We add the further condition that if then .
Proposition 2.7.
The set is in bijection with the space of conjugacy classes of reductive representations from to with non-negative Euler class. If is associated with , then and is the Hopf differential of any -equivariant harmonic map.
To parametrize representations with non-positive Euler numbers, one takes pairs such that and ; simply use an outer automorphism. About the case , the equation (7) shows that if , i.e, , then no reductive representation can carry a harmonic map with , for then we would have .
Proposition 2.7 should be known to experts. The parametrization of the character variety by pairs is proved in Hitchin’s original paper [Hit87, Theorem 10.8] (outside of Euler number , 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 comes from the holomorphic energy.
2.3.1. -Higgs bundles
Since we’re considering the adjoint group , we work with -Higgs bundles. Representations with even Euler class lift to and so for those representations one could use linear Higgs bundles. In general, one could transfer to the linear setting using the isomorphism , but the proofs are a bit faster and more natural in the principal bundle setting.
There are many excellent sources on -Higgs bundles, and we don’t need to recall everything here. We mostly draw on [ST25], which we refer to for more details. Let be a non-compact simple complex Lie group with maximal compact subgroup . The space equipped with the metric induced by the Killing form on the Lie algebra of is a Riemannian symmetric space of non-compact type. For , this symmetric space is . Let and be the Lie algebras of and respectively, and let be the Killing orthogonal complement of in . We write etc., for complexifications.
Definition 2.8.
A -Higgs bundle over a Riemann surface is a pair where is a principal -bundle and is a holomorphic -form valued in the associated bundle called the Higgs field.
See [ST25, Section 2.2] for more details on the discussion below. As is well known, an equivariant harmonic map to gives rise to a -Higgs bundle. Very briefly, is a principal -bundle that carries a principal connection induced from the Maurer-Cartan form on . The map pulls back a -bundle over with a principal connection. Using the Maurer-Cartan isomorphism, the derivative of identifies as a -form valued in , say . Upon complexifying to a principal -bundle , we can split into types as , and the harmonicity of is equivalent to the assertion that, with respect to the Koszul-Malgrange holomorphic structure on associated with the principal connection on , is a Higgs field. Given a -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 -Higgs bundle is equivalent to a -Higgs bundle equipped with a holomorphic gauge transformation of satisfying certain properties and such that . In the language of [ST25], is Coxeter cyclic. By [ST25, Proposition 1.2], the Higgs bundle is equivalent to the data of a line bundle and sections and of and respectively. An argument of Hitchin from [Hit87, Section 10] shows that we can always specify things so that, when the -Higgs bundle arises from an equivariant harmonic map, then is the Euler number of the underlying representation. In this case, the Hopf differential, up to dividing by a positive scalar , is the product . From the proof of [ST25, Proposition 1.2], under this constraint, two -Higgs bundles giving triples and as above are isomorphic if and only if there exists an isomorphism from that intertwines with and with .
From [ST25, Theorems A and 4.3], once we’ve fixed a conformal metric on , a solution to Hitchin’s self-duality equations is equivalent to a Hermitian metric on (with dual metric on ) solving the equation, for functions and on , away from their zeros,
| (8) |
It is also shown in [ST25] that , where is the Hopf differential of the harmonic map, and that 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 -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 , solving (8) requires solving only for , 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 , if , then (8) has a unique solution for with a prescribed vanishing divisor. By virtue of and having non-vanishing sections, it’s implicit in this case that . If either of or is equal to zero, [McI25, Theorem 1.3] shows that one has a solution if and only if and ( can occur only if the common degree is ) and moreover the solution is unique. We prove the following.
Proposition 2.9.
Assume that and that (8) admits a solution with associated representation and harmonic map . Then and .
Proof.
Let and be the energy and Hopf differential of respectively. Consider the function, on ,
Over each point of , and are the zeros of . 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 , Since so that , it follows that , for their divisors agree with that of and respectively. By Proposition 2.5, captures the vanishing divisor of , and hence and ∎
Note that the proof used only in the last line. If , then, as unordered sets of functions, .
Remark 2.10.
We did not strictly need to introduce and , nor their Bochner formulae. Indeed, we could have just defined using and . We prefer to use and and to show the equality with and respectively because these are classical functions and carry geometric meaning.
Lemma 2.11.
Let be a reductive representation with and carrying equivariant harmonic map , giving rise to data . Then the conjugacy class of is determined by .
Recall that the above is the constant chosen so that . In making the statement we used that for , does not vanish identically. In the proof below, when unspecified, a product of sections is the tensor product.
Proof.
We say that and are isomorphic if the corresponding -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 is determined by .
Setting , with and , the section defines a -family of isomorphisms from to the trivial bundle ; each isomorphism takes to a constant in , and specifying that constant determines the isomorphism. We choose the isomorphism so that is sent to . There is an induced isomorphism from to . Under the dual isomorphism from to , since becomes ∎
With preliminaries established, we can now prove Proposition 2.7.
Proof of Proposition 2.7.
Let be the set of isomorphism classes of triples such that we can solve (8) (uniquely) and let be the subset of such that has degree . For , we define by where and is the divisor of the square root of the holomorphic energy of the harmonic map. For , this is just , but for we’re not quite able to distinguish. Each indeed lands in : for , Proposition 2.9 gives , and hence . Still in the case , the condition is obvious. For , we can’t pick out whether or , but the existence and uniqueness theory in this case shows that neither function vanishes identically, and as well using that we get that We prove the proposition by showing that, for every , is a bijection onto the set
By Lemma 2.11, each is injective. There is no issue for : none of , , or are zero sections, so we can express or For surjectivity, fix a pair with . Looking for a -preimage we take . Note that, via the inclusion , the constant section of determines a canonical section of We take to be this section and define by Then as desired. ∎
3. Harmonic maps and domination
In this section, we prove Theorems A, C, and D, and Corollary C. Throughout, let be a closed Riemann surface of genus at least and let be a conformal metric on .
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 be a holomorphic quadratic differential on .
Definition 3.1.
We say that a function on the complement of a discrete subset of is a Bochner solution for if or for some equivariant harmonic map with Hopf differential .
Equivalently, is a Bochner solution if there is a divisor satisfying certain conditions (for example, dominated by if ) such that on the complement of the support of , is defined, , and solves
| (9) |
where . Moreover, at a point in the support of , is asymptotic to .
Proposition 3.2.
Let and be Bochner solutions for with divisors and respectively. If , then on the complement of the support of .
Proof.
We first show , and then we promote the result to outside of the support of . Set . By our assumptions, is bounded above and tends to on a non-empty discrete subset of . Assume for the sake of contradiction that at a point. Then the open subset is non-empty. Since tends to somewhere, is a proper open subset of . Taking the Laplacian of , (9) yields
| (10) |
Since is just the zero set of , is continuous on . By (10), is subharmonic on . By the weak maximum principle, is maximized on . But this contradicts . We deduce that .
We now prove that the inequality is strict. The strict inequality clearly holds near the support of . On the complement of the support of , which we will call , we have
Since and ,
where Hence, by a consequence of the strong maximum principle [Min87], either or on all of . Since tends to as we approach the support of , the former cannot occur. We conclude that on the set in question. ∎
Remark 3.3.
Proposition 3.2 extends easily to the case of a closed surface with compact boundary . If and extend continuously to and agree on , then the open subset from the proof above does not intersect , 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 be an equivariant harmonic map with holomorphic data , .
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 , so if , we have near , and if , we have near . Thus, for a branched immersion, does not flip sign.
For the main “if” direction, if , we apply Proposition 3.2 with and . Then, and , and hence our assumption shows that on the complement of the support of . By the argument from [BBDH21, pp. 12] (or using the Hartman-Wintner formula as in [Woo77]), the isolated singular points of are branch points. If , we apply Proposition 3.2 with and 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 and . Define
and
Note that is the singular set of and is the singular set of .
Proposition 3.4.
Assume that and have the same Hopf differential. For to hold everywhere, it is necessary and sufficient that the following four conditions are satisfied.
-
(1)
On .
-
(2)
On .
-
(3)
On .
-
(4)
On , .
For in a neighbourhood of a point , it is necessary and sufficient that (1) if , (2) if , and similar for (3) and (4).
Here, (as opposed to ) means that for every unit tangent vector .
Proof.
From the formula (3), if and only if . Rewriting as
| (11) |
and recalling the formula the condition is equivalent to demanding that
| (12) |
To make the notation easier on the eyes, note that (12) is equivalent to
With this in mind, we check that is necessary and sufficient. We explicitly write out the proof only for and and leave the rest to the reader, since the arguments for and are totally analogous and don’t add anything new.
On assuming (1), we have . From we get Hence , and taking absolute values yields the result. If (1) fails then we reverse the inequalities to see that
On assuming (2), Similar to above, implies that and hence If (2) does not hold, we get
As we said above, we omit the arguments for and . 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 , , and from Section 3.3. The main lemma is an application of Proposition 3.2.
Lemma 3.5.
Let be equivariant harmonic maps with holomorphic data and respectively. If , then , strictly away from the support of . If , then , strictly away from the support of .
Proof.
For the first statement, set and , which are Bochner solutions with divisors and 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 , which, as a Bochner solution, has divisor ∎
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, . Assume that . This rules out (for then we would have everywhere). The equality is ruled out too, since , and being a branched immersion implies that . If at a point , then
| (13) |
As well, , which implies that . Then, Proposition 3.4, specifically situation (1), contradicts . Using similar reasoning, we will show that . Suppose that at a point . Then,
which implies . Analogous to (13),
which via Proposition 3.4 implies that near , and thus gives a contradiction. We conclude that and
Proof of Corollary C.
Proof of Theorem A.
Let , , and be as in the statement of the theorem and let be the holomorphic data of . If , we pick any divisor of degree such that , and we apply Proposition 2.7 and Corollary C part (1) to produce the desired representation and equivariant harmonic map . If , we pick of degree and dominated by both 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 “ is a branched harmonic immersion” is necessary. This result shows that the domination problem can become delicate.
Proposition 3.6.
For , consider equivariant harmonic maps and with holomorphic data and respectively. Assume that . If there exist distinct points and at which and respectively, then neither nor hold everywhere.
Proof.
By Lemma 3.5, everywhere, strictly away from the support of . Since , we obtain
everywhere, with the same strictness condition. We deduce that
| (14) |
By assumption, the open subset is non-empty (it contains the point ). Thus, by (14), , and by Proposition 3.4 case (1), fails around . On the other hand, the presence of the point shows that is non-empty, and (14) shows that . By Proposition 3.4 case (4), does not hold around . ∎
3.5. Theorem D
Both types of examples from Theorem D come from variations on the same construction. Let be a closed Riemann surface with one boundary component and form the Riemann surface double , which comes with an antiholomorphic involution whose set of fixed points is . Necessarily, 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 so that we can define functions such as and . Let be a non-zero holomorphic quadratic differential with no zeros on and that is symmetric with respect to , i.e, such that . We view and as subsets of . We choose a first divisor on dominated by . If , we define . We point out that for such pairs, , and hence the representation corresponding to has Euler number (recall Proposition 2.5). Let be the lift of to the universal cover .
For this and as above, let be the corresponding representation and let be a -equivariant harmonic map.
Lemma 3.7.
and on
Proof.
In a local coordinate , since is anti-holomorphic, It follows using the definitions stemming from (3) that and The second equality implies that the divisor associated with is . Thus, and have the same holomorphic data. By Proposition 2.7, if the representation associated with is irreducible, then , and if the representation is reductive but not irreducible, then and are related by translation along a geodesic. Even in the latter case, since fixes , we can conclude that the two maps agree. The fact that on follows from the equality ∎
Let be the preimage of under the universal covering map. We record an analog of Proposition 1.6 for maps restricted to .
Lemma 3.8.
is a branched immersion if and only if or .
Proof.
Similar to above, we provide an analog for Corollary C. Let and be constructed as above with equivariant harmonic maps and respectively. For , we let and be the divisors corresponding to and respectively.
Lemma 3.9.
Assume that . Then if and only if or .
Proof.
By using an outer automorphism of the fundamental group, we can assume that , and under this assumption the case is removed. By Lemma 3.7, and , so on if and only if everywhere. By Lemma 3.8, is a branched immersion.
The exact same local analysis from the proof of Theorem C shows that is necessary for the domination. Assume that . By Lemma 3.7, both and vanish on , and from we obtain that Setting and , we apply Proposition 3.2 and Remark 3.3 to obtain on (analogous to Lemma 3.5), and the inequality is strict when . By the argument from Theorem C, on . By symmetry, the same is true on . ∎
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 and giving rise to data and respectively with and , exactly as in Lemma 3.9.
For (2), we consider as above and, for as above, we specify that . Let be the corresponding representation and equivariant harmonic map. By Lemmas 3.7 and 3.8, and are branched immersions, and is singular on the shared frontier of and . In fact, since (and , and are immersions.
Let be any other pair with holomorphic data and assume that . Since is non-degenerate on and , the singular set of is contained in , and and are immersions. By applying an outer automorphism to (which will not alter ), we can assume that . Then is contained in the singular set of , and the latter being empty forces . Now, we have that either or . If , the non-degeneracy of on shows that , which forces to be Fuchsian. Similarly, if , then the divisor of the square root of the anti-holomorphic energy on , which is by definition , is contained in the singular set of , and thus . We therefore deduce by Proposition 2.7 that and are conjugate. But then , and we have a contradiction. ∎
4. Anti-de Sitter -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 be the Lie algebra of . On , we consider the bilinear form , which is invariant under the adjoint representation. It can be shown that coincides with times the Killing form of . This form has signature and induces a bi-invariant Lorentzian metric on , which we denote by . We define the -dimensional anti-de Sitter space to be:
The group of orientation- and time-orientation–preserving isometries is , acting on by left and right multiplication. That is,
A tangent vector is timelike if , lightlike if , and spacelike if . We call a geodesic 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
| (15) |
for . These are topological circles and have Lorentzian length . Note that under this identification, it can be checked that for any isometry of , we have
Hence, the 1-to-1 correspondence is equivariant with respect to the action of on and on the set of timelike geodesics.
We end this subsection by briefly recalling the notion of a geometric structure on a -manifold . Let be a manifold and a Lie group acting transitively on by analytic diffeomorphisms. Then a -structure on is a maximal atlas of coordinate charts on with values in such that the transition maps are given by elements of . An important result from the theory is that is equipped with a holonomy representation and a -equivariant local diffeomorphism , called the developing map. The pair is defined up to the action of , where acts by conjugation on the holonomy representation and post-composition on the developing map. In this paper, we focus on the case where is the three-dimensional anti-de Sitter space and . The corresponding -structures on 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 -dimensional hyperbolic geometry.
4.2.1. Hyperbolic cone singularities
Let , and let be the geodesic parametrized by hyperbolic arc length given by
so that and in the boundary at infinity of . The universal cover of is given by
| (16) |
The group of deck transformations of , which is isomorphic to the fundamental group of , is given by
We endow with the Riemannian metric obtained by pulling back the hyperbolic metric via . In this model, the isometry group satisfies . More precisely,
For , we define the local model of a hyperbolic cone structure by
| (17) |
We call the point the branched point of . To justify this terminology, we define the universal branched cover of as the quotient:
| (18) |
where for all , identifying all points at to a single point, denoted . The universal covering map extends to this space by setting for all . This is well defined since and . Thus, , making the branch point of the covering .
As a particular case, it is worth observing that when , the surface is a degree- cover of . Indeed, the universal covering map induces the degree- cover
where denotes the class of in . We also note that can be identified with via the map
| (19) |
Therefore, it will be useful to view as the punctured hyperbolic plane equipped with the degree- covering map
In the disc model, this map corresponds to .
4.2.2. Spin-cone structure
In what follows, for each , we consider the following isometries of the hyperbolic Poincaré half plane:
The matrix acts as a translation of length along the geodesic in with endpoints and , whereas acts as a rotation of angle fixing the point (alternatively, one may view it as a rotation fixing the origin in the Poincaré disc model). Let denote the set of elliptic isometries fixing , see (15). Observe that the curve , given by , is the arc-length parametrization of . Hence, the Lorentzian length of is equal to , that is,
| (20) |
Now, consider the space . In [Jan22, Proposition 3.5.1], Janigro defines the universal cover of by the map
| (21) |
The group of deck transformations of , which is isomorphic to the fundamental group of , is given by
We endow with the Lorentzian metric obtained by pulling back the anti-de Sitter metric via . In this model, . More precisely, we have
| (22) |
Note that the isometry group of identifies with isometries of that fix the timelike geodesic , that is,
The map induces a homomorphism . If , then satisfies:
| (23) |
The kernel of is the group of deck transformations of the covering , given by
| (24) |
Remark 4.1.
If we take to be another timelike geodesic in , then there exists an isometry of sending to . Hence, there is an isometry between and , and we still denote the isometries of by as in (22).
For , we define as the lattice in generated by and . That is,
| (25) |
Observe that for any . Hence, may be regarded as an element of . Without loss of generality, we may assume . Following [Jan22], we give the definition below.
Definition 4.2.
Let and . We define the local model of a spin-cone singularity as
The next lemma can be viewed as a generalization, in the singular setting, of the fact that the anti-de Sitter space is a circle bundle over , with fibers being timelike geodesics of length .
Lemma 4.3.
Let be different from zero, and . Consider the projection map , which induces a map . Then, is a fibration with the property that each fiber is a timelike geodesic of length .
Proof.
Consider the projection map , which induces a map . The fiber above is given by
We claim that is a timelike geodesic with arc-length parametrization
Let be such that . Then there exist such that
Since , it follows that and hence . Since , this implies and , so is injective.
Next, the curve
parametrizes a timelike geodesic of by arc length. Hence, the fiber is a timelike geodesic of length . This completes the proof. ∎
We now consider the special case for some and , which will be our main focus. First of all, it can be shown without difficulties that the universal covering map induces a diffeomorphism , which we continue to denote by (see [Jan22, Proposition 3.5.1]). For each integer , we denote by the map from that sends to . According to the previous result, it is straightforward to check that induces a diffeomorphism , for which we keep the notation . We denote by the equivalence class of under the action of . We observe that
| (26) |
is a degree- covering map, which is also the case for the map defined by
| (27) |
We summarize this discussion in the following lemma.
Lemma 4.4.
The space is a degree- covering of .
We now move on to define anti-de Sitter manifolds with spin-cone singularities. Let be an oriented three-manifold, and let be a link in , i.e., a finite disjoint union of embedded circles . For each , we consider , a tubular neighborhood of . Each such neighborhood is homeomorphic to the solid torus , and the complement is homeomorphic to , where denotes the punctured disc.
On the universal cover of , we introduce cylindrical coordinates such that the universal covering map is given by
| (28) |
We define the universal branched cover of branched over as the quotient:
| (29) |
where for any and . The real line attached to can thus be identified with , collapsing the -plane along the -axis. Furthermore, observe that the covering map defined in (28) extends to by taking . Since , we may similarly define the universal cover of branched over and denote it by .
Definition 4.5.
Let be a link in as defined above. We say that an anti-de Sitter structure on has spin-cone singularities along if the following conditions hold:
-
•
The restriction of the developing map
extends continuously to . Namely, in cylindrical coordinates , the limit
(30) exists and is independent of . See Figure 1.
-
•
The map sends onto a complete timelike geodesic .
-
•
The lifted holonomy around a meridian encircling is given by for some , and the holonomy of a longitude is (see (22) for notation).
We say that the structure is branched (or that is a branched AdS manifold) if .
As a consequence of the above definition, each tubular neighborhood is locally isometric to the local model of a spin-cone singular manifold .
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 an immersion. In contrast, we consider a general pair of dominating maps defined on closed hyperbolic surfaces, with no restriction on 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 (not necessarily closed) with fundamental group , and two smooth maps such that
-
(a)
is a local diffeomorphism on an open dense subset. We denote by the subset of points where is singular. Note that any harmonic map either has this property or has an image contained in a geodesic [Sam78, Theorem 3].
-
(b)
Denote by an open cover of such that is a diffeomorphism onto its image. We require the cover to be -invariant, in the sense that the indices are labeled equivariantly under the deck transformation action
(31) -
(c)
The map dominates , i.e. . In particular, we may shrink the neighborhood so that for all , we have
5.1. Gluing of the fibration
For each open set , we define
It turns out that is an open subset of , foliated by timelike geodesics, a fact observed in [GK17, Proposition 7.2].
Proposition 5.1 ([Sag24, Proposition 6.1]).
For each , the subset is open. Moreover, the map
where is the unique point in satisfying , defines a principal –bundle whose fibers are timelike geodesics in .
Next, for each branched point , consider the timelike geodesic
For each and , we denote by and the inclusion maps. We then assemble two disjoint unions:
| (32) |
where the letters and stand for regular and singular, respectively. Define two maps:
-
•
-
•
We equip with the initial topology induced by the map
In particular, a sequence converges to if and only if
Definition 5.2.
Let be as above. We define
where in (or in ) are equivalent if and only if
| (33) |
Using the equivalence relation above, we may define maps
which descend to well‐defined maps on . We state now the principal result of this section.
Theorem 5.3.
Remark 5.4.
Although the constructions of and rely on the choice of an open covering of , different choices yield a canonical identification. To explain this, we temporarily denote the spaces by and (instead of and ) to keep track of the covering, and let denote the equivalence class of in . Then, for another covering of , the natural map
defines a homeomorphism between and , which restricts to an isometry between the resulting AdS manifolds and . 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 .
Proposition 5.5.
There is a homeomorphism that restricts to a homeomorphism between and .
To prepare the argument, consider the disjoint union
| (34) |
where each is an open set as before. We now define a map by
| (35) |
for in or in . We endow with the initial topology making continuous; that is, a subset is open if and only if
We then define on the equivalence relation
Lemma 5.6.
The quotient is homeomorphic to the image .
Proof.
Let be the quotient map, and define
Since satisfies
the induced map is well-defined. Moreover, by the definition of the initial topology, this map is continuous. The inverse of is given by
where is any point satisfying . We claim that is continuous. Let be any open set. We must show that
is open. By the definition of the quotient topology and the initial topology on , we have that for some open set . This implies that
which is open in . Hence is continuous. Because is a continuous bijection with a continuous inverse, it is a homeomorphism. This completes the proof. ∎
Next, we define a map by setting for in or in . The following lemma shows that, since dominates , the map descends to the quotient .
Lemma 5.7.
The map is continuous and induces a continuous map .
Proof.
Since dominates , the inequality
holds for all . In particular, if a sequence in , then , and hence
This shows that is continuous with respect to the topology of . Moreover, if , then , so
and therefore . Hence, we can define by . ∎
We are now in position to prove Proposition 5.5.
Proof of Proposition 5.5.
For each , we denote by the unique hyperbolic isometry that sends to and whose axis is the oriented geodesic joining to .
To construct the homeomorphism between and , we consider the map
Observe that is continuous–this follows from the continuity of and . Next, note that if in , then by definition but also by the proof of Lemma 5.7. This implies that for any . Therefore, we may define a continuous map between and as follows:
where denotes the equivalence classes in both and . We aim to show is a homeomorphism.
First, we define an inverse for . To this end, we identify the circle with the timelike geodesic via the map . Then, the inverse of is given by
where and are the previously defined maps. The map is continuous. Again, observe that if , then . This allows us to define a continuous inverse
As a consequence, is homeomorphic to , which in turn is homeomorphic to by Lemma 5.6. It is straightforward to see that the above construction also gives rise to a homeomorphism between and . This completes the proof. ∎
5.3. Anti-de Sitter structure
We turn our attention to the regular part . We will show that it admits an anti-de Sitter structure.
An anti-de Sitter structure on a three manifold is a structure. By definition, this is a maximal atlas of coordinate charts on with values in such that the transition maps are given by elements of .
Consider the projection map
Using the topology induced on , we have the following.
-
•
The restriction of to is a homeomorphism onto its image.
-
•
is an open set of .
Using this, we can show the following.
Proposition 5.8.
is an anti-de Sitter manifold with an atlas of charts
Moreover, is a principal –bundle with timelike-geodesic fibers.
Proof.
The atlas of charts clearly defines an anti-de Sitter structure on . Since
and the fibers of are timelike geodesics, it follows that the fibers of are timelike geodesics with respect to the anti-de Sitter structure induced by the atlas of charts
The above atlas of charts was introduced in [Jan22, Section 3.2.3]. We now show that is a fibration by circles, i.e., it is locally trivial. For each , the map is a fibration over the contractible open set , and thus it is trivial. Therefore, there exists a diffeomorphism
such that , where is the projection onto the first factor. For each index , we consider the map
where denotes the equivalence class in . We claim that is a homeomorphism.
First, for each , we necessarily have . Indeed, by definition, , hence , that is, . Second, the map is well-defined. Indeed, if with , then we have seen that , and because and are equivalent, we have . Since, is the inclusion map, we deduce .
Next, the inverse of is clearly given by
The continuity of both and follows from the definition of the topology on . Now, since is a local trivialization of , this completes the proof. ∎
We end this section by studying the case where the maps are equivariant with respect to the representations and , respectively. In what follows, we denote
We will show that acts properly discontinuously on both and . In particular, it will follow from Proposition 5.8 that the quotient of is an anti–de Sitter manifold.
First, we describe this action. For each , let . Then induces a map
| (36) |
defined by
| (37) |
It is straightforward to verify that the map satisfies the following equivariance property: for all and ,
Since is continuous and equivariant, and acts properly discontinuously on its universal cover , it follows that the action of on is properly discontinuous.
On the other hand, is a Hausdorff topological space by Proposition 5.5, and thus the quotient is a well–defined Hausdorff topological space. Moreover, the quotient inherits an anti–de Sitter structure. We summarize this discussion in the following lemma.
Lemma 5.9.
Let be maps equivariant with respect to the representations , respectively, and satisfying conditions (a)–(c). Then the map is equivariant with respect to the action of on via (see (36)) and the action of on . In particular, acts properly discontinuously on , and the quotient
is a well-defined Hausdorff topological space. Moreover, the map induces a principal -bundle over the anti-de Sitter manifold
with base , where is the quotient of by the action of . We continue to denote this map by
6. Branched AdS manifolds from branched immersions
In this section, we study the anti-de Sitter manifold obtained from a pair as above and under the additional assumption that is a branched immersion. Recall that a map between surfaces has a branch point at if there exist local complex coordinates at and at such that takes the form . We say that 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 -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 and the map has a single branch point. Understanding this example is an essential step toward the proof of the general theorem. We consider smooth maps such that dominates (i.e. ). Without loss of generality, by composing with isometries, we may assume that . We have seen in Section 5 how to construct an anti-de Sitter manifold from the maps and . Recall that
where and is an open cover of on which is a local diffeomorphism. Note that in this case, does not contain the timelike geodesic , because and take values in . From Proposition 5.8, we can define the local isometry
| (38) |
Note that takes values in . Moreover, can also be defined on all of by setting for any . We record here some notations and remarks that will be used later.
-
•
We denote by the developing map of . This can be thought of as the lift to the universal cover of the local isometry .
-
•
We denote by the developing map onto . This is given by
(39) where is the universal covering map defined in (21).
-
•
Finally, we denote by the holonomy representation of .
According to Proposition 5.8, there exists an –principal bundle 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 is nullhomotopic, any –bundle over a punctured disc must be trivial.
In what follows, we denote by a nontrivial loop in that is not homotopic to the fiber of and whose projection under is a generator of . (Here, may be viewed as a loop around the branched point of .) We also denote by the generator of that is homotopic to the future-directed timelike geodesic fiber of the fibration . Since the bundle is trivial, the fundamental group of decomposes as
| (40) |
The spacetime is time-oriented, so that corresponds to the future-directed fiber, while 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 such that dominates . Assume that and that is a degree- covering. Then is a branched anti-de Sitter manifold with singular locus . Moreover,
for some Therefore, the developing map induces a local isometry between and .
Remark 6.2.
It is worth noting that if we change the meridian by , then the holonomy of is . Thus, what really matters in the holonomy of the meridian is the degree of , and this explains why we do not require any further assumption on the map .
Remark 6.3.
The statement of Proposition 6.1 remains valid if the maps and are defined on a small neighborhood of in 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 be as in Proposition 6.1. Then, for all , we have the strict inequality
In particular, the timelike geodesics and do not intersect.
Proof.
The domination condition implies that
for every curve in , where denotes the length of the curve with respect to the hyperbolic metric . We claim that for all . For simplicity, we identify with the Poincaré disc via an identification sending the point to the origin (see (41)). Assume that for some . Consider the geodesic ray defined by . By the path lifting property, there exists a curve such that and . This implies that for each , we have
To establish the claim, we need to show that as . Indeed, for , we have , which is a compact subset of by the properness of . To see properness, recall that is equal to in an appropriate choice of local coordinates around . Therefore, must converge to some point . But since and as , it follows that . This completes the proof of the claim. To prove the strict inequality, consider a point on the geodesic segment between and . Recall that domination is strict outside the singularities. Hence, for sufficiently close to , we have . It follows that
This concludes the proof.∎
Remark 6.5.
We denote by the set of classes in that are equivalent to some element lying on . That is,
Observe that if , then cannot belong to any , since this would contradict Lemma 6.4.
The next lemma shows that the developing map of satisfies Definition 4.5.
Lemma 6.6.
The developing map extends continuously to the universal branched cover . Moreover, the restriction
is a universal covering map of the timelike geodesic .
Proof.
Let be the universal covering map. By definition, the developing map is given by , where is defined by , as in equation (38). By Proposition 5.5, is a solid torus. Therefore, the universal covering map can be described using the cylindrical coordinates (28), and hence extends continuously to . Since is already defined on all of (and not only on ), it follows that also extends to .
For the second claim, note that is a homeomorphism, and . Since restricts to the universal cover of , the composition is the universal covering map of . ∎
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 that is not homotopic to a fiber of the fibration and that generates the of the base. As in Section 6.1, we denote such a loop by , and we denote by its projection to , which generates . We now prove the following.
Proposition 6.7.
The proof of Proposition 6.7 needs some preparation.
Definition 6.8.
Let be the universal covering map of defined in (16). Then we define the angular form as the -form on
Remark 6.9.
We record the following observations.
-
(1)
If we consider the biholomorphism defined by
(41) then the angular -form has the following expression in
-
(2)
The angular form is preserved by any rotation . That is,
The following elementary and well-known lemma expresses the degree of a branched covering in terms of the angular form .
Lemma 6.10.
Let be a curve around the puncture generating , and let be a degree- cover branched on . Then,
Proof.
Since is a degree- cover of , . Hence,
Now it is enough to take a parametrization of a representative of a loop around the puncture and compute . This is possible as the integral depends only on the homotopy class of . So we take (see (16)). We then obtain , which yields the desired conclusion. ∎
The next lemma gives an expression for a lift of a curve in to
Lemma 6.11.
Let be the universal covering map of defined in (21). Let be a curve in and let be a lift of the curve such that . Then,
where are the valuations map defined by and .
Proof.
Assume that , so that
The equality follows easily. We now focus on deriving the formula for using the angular form . Observe that
| (42) |
where is the map from (16). We compute,
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 and be two paths in satisfying , for some . Assume that for all , the geodesic segment does not contain . Then,
We now turn to the proof of Proposition 6.7. Recall that is a nontrivial loop in that is not homotopic to the fiber of , and that projects to a loop around the puncture of that generates the . Let be a lift of in . For every , we denote
and
(see (39)). Consequently, we have the following.
-
•
By the construction of the manifold , the path in has the property
for every .
-
•
By definition of the action of the fundamental group, we have
(43) where is the holonomy of , see (22).
Before proving Proposition 6.7, we need a basic lemma on hyperbolic geometry. For each , we denote by the unique hyperbolic isometry that sends to and whose axis is the oriented geodesic joining to .
Lemma 6.13.
Let be such that . Then, does not fix .
Proof.
By contradiction, assume that fixes . By definition, , and hence , which is a contradiction. ∎
Proof of Proposition 6.7.
In this proof, we take explicit and in order to make computations. We set and
where the unique hyperbolic isometry that sends to and whose axis is the oriented geodesic joining to . Since , it follows that , and therefore is a loop in that is not homotopic to a fiber of the fibration .
The goal is now to compute the holonomy of the loop . Let be the lift of to . By definition, we have
and so is the lift to of . Consider such that (see notation in (22)). Thus, by Proposition 6.11, we have
| (44) |
for some . This leads to
| (45) |
We claim that
| (46) |
To prove the claim, let us consider the paths
Clearly is a loop in , and the same holds for by Lemma 6.13, which ensures that for all . Next, we want to apply Lemma 6.12. The curves satisfy for (because and ). By Lemma 6.4, we have
This implies in particular that does not belong to the geodesic segment . Therefore,
| (47) |
where the last equality follows from the fact that is a branched covering of degree (see Lemma 6.10). Next, since
it follows from (44) that
Combined with (47), we obtain which finishes the proof of the claim.
The remaining part is to compute
which is equal to , where . Since 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 for some . This concludes the proof, as
and
| (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 be a generator of that is homotopic to a future-directed timelike geodesic fiber of (see (40)). Then,
The proof of this proposition proceeds through the study of the stabilizers of timelike geodesics in . For , let be the timelike geodesic in defined in (15). If belongs to , we denote by its lift to . We denote by the stabilizer of in , and by the stabilizer of in . That is,
and
We have the following lemma.
Lemma 6.15.
[Jan22, Propositions 2.2.5 and 2.2.6] Let be such that . Consider the timelike geodesic , which is disjoint from , and let be its lift in . Then,
Proof of Proposition 6.14.
Let . By Lemma 6.4,
and the timelike geodesics and are distinct. In particular, . Let be a generator of an oriented fiber of the fibration . We may choose a representative of so that is the oriented timelike geodesic (see (38)). Let be the lift of to that is preserved by the deck transformation induced by . By the equivariance of the developing map , the holonomy preserves the timelike geodesic Hence, by Lemma 6.15, there exists such that
| (49) |
On the other hand, induces a map from
which induces an isomorphism on the level of the fundamental group. This is because is an isometry and, in particular, a homeomorphism. The isomorphism is given by , where is the integer in (49). Thus, . However, is excluded since is a time-orientation-preserving isometry. This concludes the proof. ∎
6.4. Proof of Theorem B
In this final subsection, we prove Theorem B.
Proof of Theorem B.
Let be a hyperbolic surface with fundamental group and let be smooth maps that are equivariant with respect to representations , as in Theorem B. As in Section 5, let be the singular set of the branched immersion , and set . Around each point of , the branched immersion is a degree- branched covering.
Due to Theorem 5.3 and Lemma 5.9, the only remaining point to verify in Theorem B is that is a branched AdS manifold.
Fix a point , and let be a small disc centered at on which is a branched covering of degree . We may identify with a small disc in centered at ; by applying isometries, we can further assume that and . Denote by the restrictions of and respectively. Applying Proposition 6.1 together with Remark 6.3, we obtain that is a branched anti-de Sitter manifold that fibers over .
Since the construction of is local, the behavior of around the fiber is the same as that of . In fact, we will show that can be identified with an open subset of the quotient . Indeed, one can see without difficulties that is an open set inside . Considering the natural projection we claim that, upon shrinking the neighborhood , we may assume that is injective.
To prove this, we use the fact that acts properly discontinuously on to shrink the open set so that for all . Now, by contradiction, if the restriction of to were not injective, then there would exist and such that . Using the equivariant fibration , we deduce that and , and hence . This shows that is injective.
Therefore, is identified via with an open subset of . By repeating this argument around each point of the singular locus of , we conclude that is a branched anti-de Sitter manifold, since is one.
Finally, we discuss the holonomy representation of . This is computed by studying the developing map from the universal cover to and how it transforms under the action of . This developing map is realized explicitly as the projection post-composed with the map defined in Section 5.1 (see also equations (38) and (39)). Since is a circle bundle, the fundamental group fits into the (splitting) exact sequence
By the formula for from Proposition 6.1, acts trivially on and hence the holonomy factors through the map . By the formula for from Proposition 6.1, applied around each of the singular points, we see that the holonomy further factors through the map induced by inclusion. From the definition of in Section 5.1 and the definition of the action of , i.e., equation (37), we see that the induced homomorphism is 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 -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 . 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 ‐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 . 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 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. -dimensional Lorentz space-forms and Seifert fiber spaces. J. Differential Geom., 21(2):231–268, 1985.
- [MB23] Florestan Martin-Baillon. Dominating CAT 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 -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 -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 . 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.