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

/2\mathbb{Z}/2 harmonic 1-forms, \mathbb{R}-trees, and
the Morgan-Shalen compactification

Siqi He [email protected] Morningside Center of Mathematics, Chinese Academy of Sciences, Beijing, 100190 China Richard Wentworth [email protected] Department of Mathematics, University of Maryland, College Park, MD 20742, USA  and  Boyu Zhang [email protected] Department of Mathematics, University of Maryland, College Park, MD 20742, USA
(Date: April 10, 2025)
Abstract.

This paper studies the relationship between an analytic compactification of the moduli space of flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connections on a closed, oriented 3-manifold MM defined by Taubes, and the Morgan–Shalen compactification of the SL2()\mathrm{SL}_{2}(\mathbb{C}) character variety of the fundamental group of MM. We exhibit an explicit correspondence between /2\mathbb{Z}/2 harmonic 1-forms, measured foliations, and equivariant harmonic maps to \mathbb{R}-trees, as initially proposed by Taubes. As an application, we prove that /2\mathbb{Z}/2 harmonic 1-forms exist on all Haken manifolds with respect to all Riemannian metrics. We also show that there exist manifolds that support singular /2\mathbb{Z}/2 harmonic 1-forms but have compact SL2()\mathrm{SL}_{2}(\mathbb{C}) character varieties, which resolves a folklore conjecture.

Key words and phrases:
\mathbb{Z}/2 harmonic forms, \mathbb{R}-trees, Morgan-Shalen compactification
2020 Mathematics Subject Classification:
Primary: 58D27; Secondary: 14M35, 57K35

1. Introduction

Let (M,g)(M,g) be a closed, oriented Riemannian 3-manifold with fundamental group Γ=π1(M)\Gamma=\pi_{1}(M). Let 𝒳(Γ)\mathcal{X}(\Gamma) denote the SL2()\mathrm{SL}_{2}(\mathbb{C}) character variety of Γ\Gamma as defined in, for example, [CS83]. Note that as a set, the closed points of 𝒳(Γ)\mathcal{X}(\Gamma) correspond to conjugacy classes of completely reducible representations, i.e. those that are either irreducible or direct sums of a 11–dimensional representation and its dual. The space 𝒳(Γ)\mathcal{X}(\Gamma) has the structure of an affine algebraic variety which, if positive dimensional, admits an Aut(Γ){\rm Aut}(\Gamma)–invariant compactification 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)}, called the Morgan–Shalen compactification [CS83, MS84, MS88a, MS88b]. A boundary point in 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is given by the projective class of a length function for an isometric action of Γ\Gamma on an \mathbb{R}–tree.

In [Tau13b], Taubes introduced a compactification of the space of flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connections on 33–manifolds from an analytic perspective. Consider the set of solutions to the equations

(1) FA=12[ϕ,ϕ],dAϕ=0,dAϕ=0,F_{A}=\frac{1}{2}[\phi,\phi],\quad d_{A}\phi=0,\quad d_{A}^{*}\phi=0,

where AA is a connection on an SU(2)\mathrm{SU}(2) bundle PP over MM, and ϕ\phi is a section of TM𝔤PT^{*}M\otimes\mathfrak{g}_{P}, where 𝔤P=P×ad𝔰𝔲(2)\mathfrak{g}_{P}=P\times_{ad}\mathfrak{su}(2) is the adjoint bundle. Let SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} denote the moduli space of solutions to (1) up to SU(2)\mathrm{SU}(2) gauge transformations. We abuse notation and also use 𝒳(Γ)\mathcal{X}(\Gamma) to denote the topological space of closed points of 𝒳(Γ)\mathcal{X}(\Gamma) with the analytic topology. Then by the results in [Don87] and [Cor88, Thm. 3.3], SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} is canonically homeomorphic to 𝒳(Γ)\mathcal{X}(\Gamma). The compactness results of Taubes [Tau13b] define a compactification of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})}, where the boundary points are described by a class of objects called /2\mathbb{Z}/2 harmonic 11–forms. A /2\mathbb{Z}/2 harmonic 11–form can be regarded as a generalization to 33-manifolds of holomorphic quadratic differentials on a Riemann surface. Its definition and properties will be reviewed in Section 3 below.

This article studies the relationship between the Morgan–Shalen compactification and the Taubes compactification. The Morgan–Shalen compactification is closely related to topological concepts such as singular measured foliations, incompressible surfaces, and \mathbb{R}–trees. Our main results will provide a topological interpretation of the analytic limits defined by Taubes. The relationship between the Morgan–Shalen compactification and the analytical compactification for two-dimensional manifolds was studied in [DDW00]; see also [KNPS15, HMNW23, OSWW20, LTW22, BIPP23] for extensions.

Our results are based on the theory of harmonic maps from manifolds to \mathbb{R}–trees and more generally to metric spaces of nonpositive curvature (NPC spaces) as developed by Wolf and Korevaar–Schoen [Wol95, KS93, KS97]. Let M~\widetilde{M} be the universal cover of MM. By the Corlette–Donaldson theorem, every completely reducible flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connection on MM defines a π1(M)\pi_{1}(M)–equivariant harmonic map from M~\widetilde{M} to the 33–dimensional hyperbolic space form 3\mathbb{H}^{3}. The Gromov–Hausdorff limit of the convex hulls of the images of a divergent sequence of harmonic maps is the corresponding Morgan–Shalen tree (cf. [DDW98]). We establish a direct relationship between the /2\mathbb{Z}/2 harmonic 11–forms appearing in the Taubes limit and the limiting harmonic maps to \mathbb{R}–trees. Such a relationship was proposed by Taubes in [Tau13a, pp. 12-14].

More precisely, let /2\mathcal{M}_{\mathbb{Z}/2} be the moduli space of /2\mathbb{Z}/2 harmonic 1-forms up to rescaling, equipped with the L2L^{2} topology (see (6)), and let ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} denote the Taubes compactification with boundary ¯SL2()/2\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}\subset\mathcal{M}_{\mathbb{Z}/2} (see Definition 3.8). Let 𝒫(Γ)\mathcal{PL}(\Gamma) be the space of projective length functions of minimal Γ\Gamma–actions on complete \mathbb{R}–trees (see Section 2.2). In Section 6.1, we construct a map

(2) :𝒫(Γ)/2.\mathcal{H}:\mathcal{PL}(\Gamma)\to\mathcal{M}_{\mathbb{Z}/2}.

Using this, we define a map

(3) Ξ¯:𝒳(Γ)¯¯SL2(),\begin{split}\overline{\Xi}:\overline{\mathcal{X}(\Gamma)}\to\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})},\end{split}

where Ξ¯\overline{\Xi} is given by the Riemann–Hilbert correspondence on 𝒳(Γ)\mathcal{X}(\Gamma) and by \mathcal{H} on 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)}. Then we have the following result.

Theorem 1.1.

The map Ξ¯:𝒳(Γ)¯¯SL2()\overline{\Xi}:\overline{\mathcal{X}(\Gamma)}\to\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} is continuous and surjective.

Let 𝐯¯SL2()\mathbf{v}\in\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} be a /2\mathbb{Z}/2 harmonic form in the boundary of the compactified moduli space. In Section 4, we show that the leaf space of the pull-back of 𝐯\mathbf{v} to M~\widetilde{M} naturally has the structure of an \mathbb{R}–tree, which we denote by 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}}. Let π𝐯~\pi_{\tilde{\mathbf{v}}} denote the projection from M~\widetilde{M} to 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}}. The following result describes the relationship between the leaf space and the Morgan–Shalen tree.

Theorem 1.2.

Let 𝐯¯SL2()\mathbf{v}\in\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} be a /2\mathbb{Z}/2 harmonic form in the boundary of the compactified moduli space, and assume 𝒫(Γ)\ell\in\mathcal{PL}(\Gamma) satisfies ()=𝐯\mathcal{H}(\ell)=\mathbf{v}. Suppose 𝒯\mathcal{T}_{\ell} is the minimal tree with length function \ell and u:M~𝒯u:\widetilde{M}\to\mathcal{T}_{\ell} is the equivariant harmonic map given by the Morgan–Shalen compactification. Assume uu is scaled so that its energy is equal to 11. Then

  1. (1)

    The map uu factors uniquely as the composition of π𝐯~\pi_{\tilde{\mathbf{v}}} and a continuous map f:𝒯M~,𝐯~𝒯f:\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}}\to\mathcal{T}_{\ell}.

  2. (2)

    The map ff is 11–Lipschitz.

  3. (3)

    The map π𝐯~\pi_{\tilde{\mathbf{v}}} is harmonic.

  4. (4)

    |u|=|π𝐯~||\nabla u|=|\nabla\pi_{\tilde{\mathbf{v}}}| (see Section 5.1 for the definitions). Away from the zero set of |u||\nabla u|, we have u=π𝐯~\nabla u=\nabla\pi_{\tilde{\mathbf{v}}} as two-valued differential forms.

The above results establish a bridge between the analytic and algebraic compactifications of the moduli space of flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connections on 33–manifolds. This allows us to prove new results on one side using results from the other side. We discuss several such applications in Section 7.

In Section 7.1, we prove an existence result of /2\mathbb{Z}/2 harmonic forms by considering harmonic maps to \mathbb{R}–trees.

A non-zero /2\mathbb{Z}/2 harmonic 1-form 𝐯\mathbf{v} is called non-trivial if 𝐯\mathbf{v} has non-trivial holonomy; in other words, if it is not given by a single-valued 11–form. We say that 𝐯\mathbf{v} is singular if there exists a point on MM where 𝐯\mathbf{v} cannot be locally lifted to a single-valued form. Clearly, singular implies non-trivial, but the converse is false in general. Using the map \mathcal{H} constructed in Section 6.1, we prove the following theorem.

Theorem 1.3.

Suppose MM is a rational homology sphere, and suppose there exists a closed connected embedded surface SMS\subset M such that SS is two-sided, π1\pi_{1}–injective, and does not bound an embedded ball. Then for every Riemannian metric gg on MM, there exists a non-trivial /2\mathbb{Z}/2 harmonic 1-form on (M,g)(M,g).

We remark that while all previous analytical constructions of non-trivial /2\mathbb{Z}/2 harmonic 1-forms and spinors require the metric to take specific forms ([DW21, TW20, TW24, He22, HP24, CH24]), Theorem 1.3 holds for all metrics on the given manifolds.

A rational homology sphere MM satisfies the conditions in Theorem 1.3 if and only if MM is reducible or Haken. On the other hand, if a closed oriented 3-manifold MM has b1(M)>0b_{1}(M)>0, then there always exist trivial /2\mathbb{Z}/2 harmonic forms that are given by usual harmonic 11–forms. Therefore, Theorem 1.3 has the following consequence.

Corollary 1.4.

Suppose a closed oriented 3-manifold MM is reducible or Haken. Then for every Riemannian metric gg on MM, there exist /2\mathbb{Z}/2 harmonic 1-forms on (M,g)(M,g).

Since /2\mathbb{Z}/2 harmonic 1-forms describe the boundary of compactified moduli spaces of solutions to (1), it is natural to ask whether all of them can be deformed to solutions of the equation. The Kuranishi structure near the boundary of compactified moduli spaces for generalized Seiberg–Witten equations has been studied by Doan–Walpuski [DW20] and Parker [Par24b]. Equation (1) is a special case of the generalized Seiberg–Witten equations; in general, the compactified moduli spaces are described by /2\mathbb{Z}/2 harmonic spinors, which is a generalization of the concept of /2\mathbb{Z}/2 harmonic forms.

For the Seiberg–Witten equations with two spinors, Parker [Par24b] proved that all /2\mathbb{Z}/2 harmonic spinors, under some nondegenerate conditions, can be realized as limits of 1-parameter families of solutions to the equation. Motivated by Parker’s result, there has been a folklore conjecture that /2\mathbb{Z}/2 harmonic 1-forms could not exists on any 3-manifold with compact SL2()\mathrm{SL}_{2}(\mathbb{C}) character variety. A precise formulation of this conjecture was recently stated in [HP24, Conjecture 1.14].

Theorem 1.3 implies that there are counterexamples to this folklore conjecture. In fact, we have the following result.

Corollary 1.5.

There exist infinitely many closed 3-manifolds MM such that the SL2()\mathrm{SL}_{2}(\mathbb{C}) character variety of π1(M)\pi_{1}(M) is compact, and MM supports a singular /2\mathbb{Z}/2 harmonic 1-form with respect to every Riemannian metric.

Proof.

Boyer–Zhang [BZ98, Theorem 1.8] and Motegi [Mot88] constructed infinitely many closed oriented Haken 3-manifolds whose SL2()\mathrm{SL}_{2}(\mathbb{C}) character varieties are zero-dimensional. Since character varieties are affine varieties, zero-dimensional character varieties are compact. There are infinitely many examples MM in [Mot88] such that H1(M)H_{1}(M) are cyclic groups with odd orders. If MM is a rational homology sphere such that H1(M)H_{1}(M) has no 2-torsion, every /2\mathbb{Z}/2 harmonic form on MM is singular. Hence the result is proved by Theorem 1.3. ∎

Corollary 1.5 gives the first example of /2\mathbb{Z}/2 harmonic 1-forms (and more generally, /2\mathbb{Z}/2 harmonic spinors) that cannot be deformed into solutions of the corresponding gauge-theoretic equations. It also suggests that some properties of solutions to the Seiberg–Witten equations with two spinors, as studied in [DW20, DW21, Par24b], may not generalize to flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connections.

In Sections 7.2 to 7.5, we give several more applications of Theorems 1.1 and 1.2 and discuss some related results. Using the regularity results of Parker [Par23a], we prove a stronger regularity result for the Morgan–Shalen convergence (Corollary 7.8). Using Theorem 1.2, we show that the /2\mathbb{Z}/2 harmonic forms that arise as limits in Taubes’ construction must satisfy certain additional properties (Corollary 7.9). The Hubbard–Masur map [HM79] takes the space of measured foliations on a Riemann surface, modulo certain equivalence relations, to the space of quadratic differentials. Wolf [Wol95] showed that this map is a homeomorphism. Our construction of \mathcal{H} can be interpreted as a generalization of the Hubbard–Masur map in dimension three, and it associates a canonical measured foliation with every boundary point in the Morgan–Shalen limit (see Section 7.5).

Finally, we address a folklore non-existence conjecture for /2\mathbb{Z}/2 harmonic 11–forms on manifolds that are diffeomorphic to S3S^{3}, which is motivated by the relation of /2\mathbb{Z}/2 harmonic 11–forms and the SL2()\mathrm{SL}_{2}(\mathbb{C}) representation variety. We refer to [HP24] for more discussion about this conjecture.

Conjecture 1.6.

Let gg be a Riemannian metric on S3S^{3}. There exists no /2\mathbb{Z}/2 harmonic 11–form on (S3,g)(S^{3},g).

Note that if a manifold has positive Ricci curvature, then the non-existence of /2\mathbb{Z}/2 harmonic 11–forms follows directly from the Weitzenböck formula (see Eqns. (7), (8)).

We approach this conjecture using the relationship between /2\mathbb{Z}/2 harmonic 11–forms and measured foliations, as developed in Sections 4 and 7 below. Given a /2\mathbb{Z}/2 harmonic 1-form 𝐯\mathbf{v} on S3S^{3}, the leaf space of the measured foliation defined by 𝐯\mathbf{v} is an \mathbb{R}–tree (see Theorem 4.4). Under certain assumptions, we show that the projection map from S3S^{3} to the leaf space satisfies a maximum principal, which leads to a contradiction. As a consequence, we prove the following partial resolution of Conjecture 1.6.

Theorem 1.7.

There is no /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v} on (S3,g)(S^{3},g) satisfying both of the following conditions:

  1. (1)

    Every zero point of 𝐯\mathbf{v} is cylindrical (see Definition 7.3).

  2. (2)

    For every arc γ\gamma in MZM\setminus Z transverse to ker𝐯\ker\mathbf{v}, where ZZ denotes the zero locus of 𝐯\mathbf{v}, we have μ𝐯(γ)=dM,𝐯(x,y).\mu_{\mathbf{v}}(\gamma)=d_{M,\mathbf{v}}(x,y).

Condition (1) is a local regularity requirement for the zero locus of 𝐯\mathbf{v}; it is weaker than the usual regularity assumption in the study of /2\mathbb{Z}/2 harmonic 11–forms as stated [Don21, He23, Par23b]. The notations μ𝐯\mu_{\mathbf{v}} and dM,𝐯d_{M,\mathbf{v}} in Condition (2) refer to the transverse invariant measure and the distance function on the leaf space defined by 𝐯\mathbf{v} (see (12) and (13)). We will show that Condition (2) always holds for /2\mathbb{Z}/2 harmonic 11–forms that appear on the boundary of Taubes’ compactification (Corollary 7.9, part (2)). We note that a completely different approach to Conjecture 1.6 is given in a forthcoming work of Parker [Par24a], whch uses the gluing constructions in gauge theory.

Acknowledgements. The authors wish to express their gratitude to many people for their interest and helpful comments. Among them are Nathan Dunfield, Xinghua Gao, Andriy Haydys, Zhenkun Li, Yi Liu, Ciprian Manolescu, Rafe Mazzeo, Tomasz Mrowka, Yi Ni, Jean-Pierre Otal, Greg Parker, Ao Sun, Clifford Taubes, Thomas Walpuski and Mike Wolf.

S.H. is partially supported by NSFC grant No.12288201 and No.2023YFA1010500. R.W.’s research is supported by NSF grant DMS-2204346. B.Z. is partially supported by NSF grant DMS-2405271 and a travel grant from the Simons Foundation.

2. The Morgan-Shalen compactification, length functions, and \mathbb{R}-trees

In this section, we briefly review relevant results about the Morgan–Shalen compactification and \mathbb{R}–trees from [CS83, MS84, MS88a, MS88b]. For a more comprehensive introduction, we refer the reader to the survey by Otal [Ota15].

2.1. \mathbb{R}–trees and length functions

An \mathbb{R}–tree is a metric space (𝒯,d𝒯)(\mathcal{T},d_{\mathcal{T}}) such that every pair of points is connected by a unique arc, and every arc is isometric to a closed interval in \mathbb{R} as a subspace of 𝒯\mathcal{T}.

Let Γ\Gamma be a finitely generated group. A Γ\Gamma–tree is an \mathbb{R}–tree 𝒯\mathcal{T} with an isometric action ρ:ΓIsom(𝒯)\rho:\Gamma\to\text{Isom}(\mathcal{T}). A Γ\Gamma–tree is called minimal if there is no proper Γ\Gamma–invariant subtree. The length function is defined by

(4) ρ(γ):=infx𝒯d𝒯(x,ρ(γ)x).\ell_{\rho}(\gamma):=\inf_{x\in\mathcal{T}}d_{\mathcal{T}}(x,\rho(\gamma)x).

By [CM87, Sec. (1.3)], every γΓ\gamma\in\Gamma acts semisimply, i.e., the infimum is realized at some point in 𝒯\mathcal{T}. By [MS84, Prop. II.2.15], ρ\ell_{\rho} is identically zero if and only if Γ\Gamma has a fixed point.

In general, ρ(γ)\ell_{\rho}(\gamma) depends only on the conjugacy class of ρ\rho and the conjugacy class of γ\gamma. So let CC be the set of conjugacy classes of Γ\Gamma, and let (C):=(C)\mathbb{P}(C):=\mathbb{P}(\mathbb{R}^{C}) be the real projective space of nonzero functions on CC. If ρ0\ell_{\rho}\not\equiv 0, then the class of ρ\ell_{\rho} in (C)\mathbb{P}(C) is called the projective length function.

A length function is called abelian if it is given by |μ(γ)||\mu(\gamma)| for some homomorphism μ:Γ\mu:\Gamma\to\mathbb{R}. A ray RR in an \mathbb{R}–tree is the image of an isometric embedding of [0,+)[0,+\infty). A Γ\Gamma–tree is said to have a fixed end if there exists a ray RR such that for every γΓ\gamma\in\Gamma, γ(R)R\gamma(R)\cap R is also a ray. We will need the following result.

Theorem 2.1 ([CM87, Thm. 3.7]).

Assume 𝒯\mathcal{T} is a minimal Γ\Gamma–tree with a non-trivial length function \ell. Then \ell is non-abelian if and only if Γ\Gamma acts without fixed ends. Moreover, if \ell is non-abelian and 𝒯\mathcal{T}^{\prime} is another minimal Γ\Gamma–tree with the same length function, then there exists a unique Γ\Gamma-equivariant isometry between 𝒯\mathcal{T} and 𝒯\mathcal{T}^{\prime}.

When \ell is abelian, the Γ\Gamma–tree is not always uniquely determined by the length function. A counterexample is given in [CM87, Ex. 3.9]. To simplify terminology, we will usually refer to a “Γ\Gamma–tree” as a “tree” when the Γ\Gamma–action is clear from the context.

2.2. The Morgan–Shalen compactification

The character χρ:Γ\chi_{\rho}:\Gamma\to\mathbb{C} of a representation ρ:ΓSL2()\rho:\Gamma\to\mathrm{SL}_{2}(\mathbb{C}) is defined by χρ(γ):=Tr(ρ(γ))\chi_{\rho}(\gamma):=\mathrm{Tr}(\rho(\gamma)). Clearly, χρ\chi_{\rho} depends only on ρ\rho up to overall conjugation. By definition, 𝒳(Γ)\mathcal{X}(\Gamma) is the set of all possible characters. By invariant theory, 𝒳(Γ)\mathcal{X}(\Gamma) has the structure of an affine complex algebraic variety. Locally, points are determined by finitely many characters from a sufficiently large generating set for Γ\Gamma. For more details, we refer to [CS83, Cor. 1.4.5].

The Morgan–Shalen compactification 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} can be described as follows [MS84]: Let C\mathbb{R}^{C} be the vector space of real-valued functions on CC with the weak topology, and (C):=(C)\mathbb{P}(C):=\mathbb{P}(\mathbb{R}^{C}) the projective space of nonzero real-valued functions on CC with the quotient topology. Define a map ϑ:𝒳(Γ)C\vartheta:\mathcal{X}(\Gamma)\to\mathbb{R}^{C} by

ϑ(ρ):C,[γ]log(|χρ(γ)|+2).\vartheta(\rho):C\to\mathbb{R}\ ,\ [\gamma]\mapsto\log(|\chi_{\rho}(\gamma)|+2).

Since ϑ(ρ)(1)0\vartheta(\rho)(1)\neq 0, the element ϑ(ρ)\vartheta(\rho) has an image in (C)\mathbb{P}(C), which we denote by [ϑ(ρ)][\vartheta(\rho)]. Let 𝒳(Γ)+\mathcal{X}(\Gamma)^{+} denote the one-point compactification of 𝒳(Γ)\mathcal{X}(\Gamma) with the inclusion map ι:𝒳(Γ)𝒳(Γ)+\iota:\mathcal{X}(\Gamma)\to\mathcal{X}(\Gamma)^{+}. The Morgan–Shalen compactification 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is then the closure of the embedded image of 𝒳(Γ)\mathcal{X}(\Gamma) in 𝒳(Γ)+×(C)\mathcal{X}(\Gamma)^{+}\times\mathbb{P}(C) by the map ι×[ϑ]\iota\times[\vartheta].

Note that the group PSL2():=SL2()/2\mathrm{PSL}_{2}(\mathbb{C}):=\mathrm{SL}_{2}(\mathbb{C})/\mathbb{Z}_{2} is equal to the identity component of the isometry group of the 3–dimensional hyperbolic space form 3\mathbb{H}^{3}. For each representation ρ\rho, we define

ρ:C,[γ]infx3d3(x,ρ(γ)x).\ell_{\rho}:C\to\mathbb{R}\ ,\ [\gamma]\mapsto\inf_{x\in\mathbb{H}^{3}}d_{\mathbb{H}^{3}}(x,\rho(\gamma)x).

It is straightforward to verify that |ρ(γ)2log|χρ(γ)||2|\ell_{\rho}(\gamma)-2\log|\chi_{\rho}(\gamma)||\leq 2 for all γ\gamma (cf. [CS83]). Hence, if a sequence ρn𝒳(Γ)\rho_{n}\in\mathcal{X}(\Gamma) converges to a point ρ\rho_{\infty} in 𝒳(Γ)¯𝒳(Γ)\overline{\mathcal{X}(\Gamma)}\setminus\mathcal{X}(\Gamma), then the limit of [ϑ(ρn)][\vartheta(\rho_{n})] is the same as the limit of [ρn][\ell_{\rho_{n}}].

We record the following result for later reference.

Lemma 2.2 (cf. [CS83, Ota15]).

𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is Hausdorff and compact.

Proof.

Since both 𝒳(Γ)+\mathcal{X}(\Gamma)^{+} and (C)\mathbb{P}(C) are Hausdorff, the space 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is Hausdorff. By [Ota15, Prop. 8], the projection of the closure of the image of ι×[ϑ]\iota\times[\vartheta] to (C)\mathbb{P}(C) is compact. Since 𝒳(Γ)+\mathcal{X}(\Gamma)^{+} is compact, this implies that 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is compact.∎

The following result relates the limit of length functions with \mathbb{R}–trees.

Theorem 2.3 ([MS84, MS88a, MS88b]).

Given a sequence of representations ρn:ΓSL2()\rho_{n}:\Gamma\to\mathrm{SL}_{2}(\mathbb{C}), the following holds:

  1. (i)

    Suppose χρn(γ)\chi_{\rho_{n}}(\gamma) is bounded for each γΓ\gamma\in\Gamma. Then, after passing to a subsequence if necessary, ρn\rho_{n} converges to an element ρ\rho_{\infty} in 𝒳(Γ)\mathcal{X}(\Gamma).

  2. (ii)

    Suppose limn|χρn(γ)|=\lim_{n\to\infty}|\chi_{\rho_{n}}(\gamma)|=\infty for some γΓ\gamma\in\Gamma. Then, after passing to a subsequence, there exists a minimal Γ\Gamma–tree (𝒯,d𝒯)(\mathcal{T},d_{\mathcal{T}}), where the action is denoted by ρ:ΓIsom(𝒯)\rho_{\infty}:\Gamma\to\mathrm{Isom}(\mathcal{T}), such that its length function ρ\ell_{\rho_{\infty}} satisfies ρ0\ell_{\rho_{\infty}}\not\equiv 0, and [ρn][ρ][\ell_{\rho_{n}}]\to[\ell_{\rho_{\infty}}] in (C)\mathbb{P}(C).

Not every element [](C)[\ell]\in\mathbb{P}(C) can be realized as a length function of a tree (see [Chi76] and [CM87, p. 586]). If an element of (C)\mathbb{P}(C) is realized by the length function of a Γ\Gamma–tree, then it can be realized by a minimal Γ\Gamma–tree [CM87, Prop. 3.1]. Define 𝒫(Γ)\mathcal{PL}(\Gamma) to be the subspace of (C)\mathbb{P}(C) consisting of all elements that can be realized by the length functions of minimal Γ\Gamma–trees that are complete as metric spaces. For each []𝒫(Γ)[\ell]\in\mathcal{PL}(\Gamma), since \ell is not identically zero, the Γ\Gamma–action on the corresponding tree has no fixed point.

3. /2\mathbb{Z}/2 Harmonic 1-forms and Taubes’ compactness results

In this section, we review the compactness results on stable SL2()\mathrm{SL}_{2}(\mathbb{C}) flat connections due to Taubes [Tau13b, Tau13a]. We will use these results to define a compactification of the moduli space of flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connections using /2\mathbb{Z}/2 harmonic 1-forms.

3.1. /2\mathbb{Z}/2 harmonic 1-forms.

Let (M,g)(M,g) be a Riemannian manifold with metric gg. Here, MM is allowed to be non-compact or non-complete but we assume that MM has no boundary. The concept of a /2\mathbb{Z}/2 harmonic 11–form was first introduced by Taubes [Tau13b, Tau13a] in order to describe a compactification of the moduli spaces of solutions to a class of gauge-theoretic equations.

Definition 3.1.

A /2\mathbb{Z}/2 harmonic 11–form on MM is given by a closed subset ZMZ\subsetneq M and a two-valued section 𝐯\mathbf{v} of TMT^{*}M on the complement of ZZ, such that the following conditions hold:

  • (i)

    For each xZx\notin Z, there exists an open neighborhood UMZU\subset M\setminus Z of xx such that the values of 𝐯\mathbf{v} on UU have the form ±v\pm v, where vv is a non-vanishing section of TM|UT^{*}M|_{U} such that dv=0dv=0, dv=0d^{*}v=0.

  • (ii)

    For every open subset UU of MM such that U¯\overline{U} is compact, we have UZ|𝐯|2<+\int_{U\setminus Z}|\mathbf{v}|^{2}<+\infty, and UZ|𝐯|2<+\int_{U\setminus Z}|\nabla\mathbf{v}|^{2}<+\infty.

  • (iii)

    Let nn be the dimension of MM. There exist constants CC and ϵ>0\epsilon>0 such that for every xZx\in Z and every r<1r<1 less than the injectivity radius of MM at xx, we have Br(x)|𝐯|2<Crn+ϵ\int_{B_{r}(x)}|\mathbf{v}|^{2}<C\cdot r^{n+\epsilon}.

Here, |𝐯||\nabla\mathbf{v}| and |𝐯||\mathbf{v}| are defined pointwise as the values of |v||\nabla v| and |v||v| where vv satisfies Condition (i). Condition (iii) above is needed for the proofs of some key properties of /2\mathbb{Z}/2 harmonic 11–forms. The set ZZ is called the zero locus of 𝐯\mathbf{v}.

Example 3.2 ([Tau13b, p. 9]).

When MM is a Riemann surface, a /2\mathbb{Z}/2 harmonic 1-form is given by the real part of the square root of a holomorphic quadratic differential. More explicitly, given a quadratic differential qH0(KM2)q\in H^{0}(K_{M}^{2}), where KMK_{M} is the canonical bundle, 𝐯:=Re(q)\mathbf{v}:=\mathrm{Re}(\sqrt{q}) defines a /2\mathbb{Z}/2 harmonic 1-form. Conversely, given a /2\mathbb{Z}/2 harmonic 1-form, let 𝐯(1,0)\mathbf{v}^{(1,0)} be the (1,0)(1,0) component of 𝐯\mathbf{v}, then 𝐯(1,0)𝐯(1,0)\mathbf{v}^{(1,0)}\otimes\mathbf{v}^{(1,0)} also defines a quadratic differential. This correspondence is a bijection. In particular, there exists no /2\mathbb{Z}/2 harmonic 1-form on S2S^{2}.

We have the following analytic property for the zero loci of /2\mathbb{Z}/2 harmonic 1-forms.

Theorem 3.3 ([Tau14, Zha22]).

Assume MM is connected and the dimension of MM is n4n\leq 4. Then the zero locus of every /2\mathbb{Z}/2 harmonic 11–form on MM is (n2)(n-2)–rectifiable with locally finite (n2)(n-2)–Hausdorff measure.

In particular, the zero locus ZZ of every /2\mathbb{Z}/2 harmonic 11–form has Lebesgue measure zero. Therefore, we will sometimes write UZ\int_{U\setminus Z} as U\int_{U} when the integrand is defined on the complement of ZZ and extends to an L1L^{1} function over UMU\subset M.

It is convenient to regard a /2\mathbb{Z}/2 harmonic 1-form as a section of the bundle TM/{±1}T^{*}M/\{\pm 1\} over MM, where the fiber of TM/{±1}T^{*}M/\{\pm 1\} at each point xMx\in M is the quotient space (TM|x)/{±1}(T^{*}M|_{x})/\{\pm 1\}. A distance function on the quotient space (TM|x)/{±1}(T^{*}M|_{x})/\{\pm 1\} is defined by

(5) d(±v1,±v2)=min{|v1v2|,|v1+v2|}.d(\pm v_{1},\pm v_{2})=\min\{|v_{1}-v_{2}|,|v_{1}+v_{2}|\}.

We define the moduli space of /2\mathbb{Z}/2 harmonic 1-forms as

(6) /2={𝐯𝐯as in Definition 3.1, and 𝐯L2(M)=1},\mathcal{M}_{\mathbb{Z}/2}=\{\mathbf{v}\mid\mathbf{v}\;\text{as in Definition \ref{def_ZTharmonicform}, and }\|\mathbf{v}\|_{L^{2}(M)}=1\},

and endow /2\mathcal{M}_{\mathbb{Z}/2} with the L2L^{2} topology induced from the distance function (5).

Proposition 3.4.

Assume MM is closed and has dimension no greater than 4. Then the space /2\mathcal{M}_{\mathbb{Z}/2} is compact.

Proof.

We first show that there exists a constant CC depending only on MM and the Riemannian metric, such that for each 𝐯/2\mathbf{v}\in\mathcal{M}_{\mathbb{Z}/2} with zero locus ZZ, we have MZ|𝐯|2C.\int_{M\setminus Z}|\nabla\mathbf{v}|^{2}\leq C.

By the Weitzenböck formula, near every xMZx\in M\setminus Z, if we write 𝐯\mathbf{v} as {±v}\{\pm v\} such that dv=0,dv=0dv=0,d^{*}v=0, then

(7) 12dd|v|2+|v|2+Ric(v,v)=0,\frac{1}{2}d^{*}d|v|^{2}+|\nabla v|^{2}+\mathrm{Ric}(v,v)=0,

where Ric\mathrm{Ric} is the Ricci curvature of MM.

We show that

(8) MZdd|𝐯|2=0.\int_{M\setminus Z}d^{*}d|\mathbf{v}|^{2}=0.

For each positive integer ii, let ρi\rho_{i} be a smooth function on \mathbb{R} such that ρi(x)=0\rho_{i}(x)=0 when xix\leq-i, ρi(x)=1\rho_{i}(x)=1 when xi+2x\geq-i+2, and ρi(x)[0,1]\rho_{i}^{\prime}(x)\in[0,1] for all xx. Define ηi\eta_{i} by

ηi={ρi(ln|𝐯|) if |𝐯|00 if |𝐯|=0,\eta_{i}=\begin{cases}\rho_{i}(\ln|\mathbf{v}|)&\text{ if }|\mathbf{v}|\neq 0\\ 0&\text{ if }|\mathbf{v}|=0,\end{cases}

then ηi\eta_{i} is a smooth function on MM.

On MZM\setminus Z, we have

|ηi|=|ρi(ln|𝐯|)||𝐯||𝐯||𝐯||𝐯||𝐯||𝐯|.|\nabla\eta_{i}|=|\rho_{i}^{\prime}(\ln|\mathbf{v}|)|\cdot\frac{\nabla|\mathbf{v}|}{|\mathbf{v}|}\leq\frac{\nabla|\mathbf{v}|}{|\mathbf{v}|}\leq\frac{|\nabla\mathbf{v}|}{|\mathbf{v}|}.

Hence the following inequality holds:

|ηi||𝐯||𝐯|.|\nabla\eta_{i}|\cdot|\mathbf{v}|\leq|\nabla\mathbf{v}|.

Also note that the support of ηi\nabla\eta_{i} is a subset of {xM:|𝐯(x)|[ei,ei+2]}\{x\in M:|\mathbf{v}(x)|\in[e^{-i},e^{-i+2}]\}. Since |𝐯|L2(MZ)|\nabla\mathbf{v}|\in L^{2}(M\setminus Z), we conclude that

limisupp ηi|𝐯|2=0,\lim_{i\to\infty}\int_{\textrm{supp }\eta_{i}}|\nabla\mathbf{v}|^{2}=0,

and hence |ηi||𝐯||\nabla\eta_{i}|\cdot|\mathbf{v}| converges to zero in L2(MZ)L^{2}(M\setminus Z) as ii goes to infinity.

By integration by parts, we have

(9) MZ(dd|𝐯|2)ηi=MZ(d|𝐯|2)(dηi).\int_{M\setminus Z}(d^{*}d|\mathbf{v}|^{2})\cdot\eta_{i}=\int_{M\setminus Z}(d|\mathbf{v}|^{2})(d\eta_{i}).

By (7), the function dd|𝐯|2d^{*}d|\mathbf{v}|^{2} is integrable on MZM\setminus Z, and hence the limit of the left-hand side of (9) equals MZdd|𝐯|2\int_{M\setminus Z}d^{*}d|\mathbf{v}|^{2}. Since 𝐯L2(MZ)<+\|\nabla\mathbf{v}\|_{L^{2}(M\setminus Z)}<+\infty and |ηi||𝐯|0|\nabla\eta_{i}|\cdot|\mathbf{v}|\to 0 in L2L^{2} as ii goes to infinity, the right-hand side of (9) converges to zero as ii goes to infinity. Hence (8) is proved. By (7), we conclude that MZ|𝐯|2C\int_{M\setminus Z}|\nabla\mathbf{v}|^{2}\leq C for some constant CC depending only on MM.

The Sobolev space W1,pW^{1,p} for multi-valued sections of a vector bundle was studied in [DLS11, Ch. 4]. By [Zha22, Lem. 2.1], the /2\mathbb{Z}/2 harmonic 1-form 𝐯\mathbf{v} can be regarded as a 22–valued section with W1,2W^{1,2} regularity over MM, and its W1,2W^{1,2}–norm is equal to (up to constant multiplicative factors) MZ|𝐯|2+|𝐯|2\int_{M\setminus Z}|\mathbf{v}|^{2}+|\nabla\mathbf{v}|^{2}. The desired proposition then follows from the Rellich compactness theorem for multi-valued sections [DLS11, Prop. 4.6(i)]. ∎

By [Tau14, Lem. 4.6], every 𝐯/2\mathbf{v}\in\mathcal{M}_{\mathbb{Z}/2} is Hölder continuous.

Proposition 3.5.

Assume MM is closed. Then the L2L^{2} topology and the 𝒞0\mathcal{C}^{0} topology coincide on /2\mathcal{M}_{\mathbb{Z}/2}.

Proof.

Since both topologies are metric spaces, we only need to show that they define the same convergence condition for sequences. It is obvious that 𝒞0\mathcal{C}^{0} convergence implies L2L^{2} convergence; we show that the converse also holds. Namely, assuming {𝐯i}\{\mathbf{v}_{i}\} is a sequence in /2\mathcal{M}_{\mathbb{Z}/2} that converges to 𝐯\mathbf{v} in L2L^{2}, we show that {𝐯i}\{\mathbf{v}_{i}\} converges to 𝐯\mathbf{v} in 𝒞0\mathcal{C}^{0}.

Let inj(M)\mathrm{inj}(M) denote the injectivity radius of MM. By [Tau14, Lem. 2.3] and a rescaling argument, there exists a constant C1C\geq 1 depending only on MM such that

(10) supBr/2(x)|𝐮|2CBr(x)|𝐮|2\sup_{B_{r/2}(x)}|\mathbf{u}|^{2}\leq C\fint_{B_{r}(x)}|\mathbf{u}|^{2}

for all 𝐮/2\mathbf{u}\in\mathcal{M}_{\mathbb{Z}/2}, xMx\in M, and r<inj(M)r<\mathrm{inj}(M).

Let {𝐯i}\{\mathbf{v}_{i}\} be a sequence in /2\mathcal{M}_{\mathbb{Z}/2} that converges to 𝐯\mathbf{v} in the L2L^{2} topology. Let D>1D>1 be a constant such that vol(Bd(x))<Dvol(Bd/2(x))\mathrm{vol}\big{(}B_{d}(x)\big{)}<D\cdot\mathrm{vol}\big{(}B_{d/2}(x)\big{)} for every xMx\in M and d<inj(M)d<\mathrm{inj}(M). For each ϵ>0\epsilon>0, let BϵB_{\epsilon} be the set consisting of all xMx\in M such that |𝐯(x)|2<ϵ2/(8CD)|\mathbf{v}(x)|^{2}<\epsilon^{2}/(8CD), where CC is the constant in (10). Let dϵd_{\epsilon} be the distance between ZZ and MBϵM\setminus B_{\epsilon}. Take ϵ\epsilon sufficiently small so that dϵ<inj(M)d_{\epsilon}<\mathrm{inj}(M).

Since {𝐯i}\{\mathbf{v}_{i}\} converges to 𝐯\mathbf{v} in L2L^{2}, for ii sufficiently large, we have

Bdϵ(x)|𝐯i|2<ϵ24CD\fint_{B_{d_{\epsilon}}(x)}|\mathbf{v}_{i}|^{2}<\frac{\epsilon^{2}}{4CD}

for all xZx\in Z. Let Vϵ=xZBdϵ/2(x)V_{\epsilon}=\cup_{x\in Z}B_{d_{\epsilon}/2}(x). By (10), this implies

supVϵ|𝐯i|2<ϵ24.\sup_{V_{\epsilon}}|\mathbf{v}_{i}|^{2}<\frac{\epsilon^{2}}{4}.

So the 𝒞0\mathcal{C}^{0}–distance between 𝐯i\mathbf{v}_{i} and 𝐯\mathbf{v} on VϵV_{\epsilon} is no greater than

supVϵ(|𝐯i|+|𝐯|)<ϵ(1/2+1/(8CD))<ϵ.\sup_{V_{\epsilon}}(|\mathbf{v}_{i}|+|\mathbf{v}|)<\epsilon(1/2+\sqrt{1/(8CD)})<\epsilon.

On the other hand, 𝐯i\mathbf{v}_{i} converges to 𝐯\mathbf{v} uniformly on MVϵM\setminus V_{\epsilon} by standard elliptic estimates. Hence the desired result is proved. ∎

3.2. The moduli space SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} and its compactification

Let MM be a closed oriented 3-manifold. Let PP be a principal SU(2)\mathrm{SU}(2) bundle over MM, and let 𝔤P\mathfrak{g}_{P} be the associated 𝔰𝔲(2)\mathfrak{su}(2) bundle given by the adjoint action. Consider the system of equations

(11) FA=12[ϕ,ϕ],dAϕ=0,dAϕ=0,\begin{split}F_{A}=\frac{1}{2}[\phi,\phi],\quad d_{A}\phi=0,\quad d_{A}^{*}\phi=0,\end{split}

where AA is a connection on PP and ϕ\phi is a section of TM𝔤PT^{*}M\otimes\mathfrak{g}_{P}. We define SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} to be the set of solutions to (11), modulo SU(2)\mathrm{SU}(2) gauge transformations. The topology on SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} is given by the Wk,pW^{k,p}–Sobolev norm for k,pk,p sufficiently large. By the standard elliptic bootstrapping argument, the topology on SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} does not depend on the choice of (k,p)(k,p) when p>1p>1 and kk is sufficiently large.

Note that if 𝐯\mathbf{v} is a /2\mathbb{Z}/2 harmonic 11–form, then 𝐯𝐯\mathbf{v}\otimes\mathbf{v} is a single-valued section of TMTMT^{*}M\otimes T^{*}M. The following result is a consequence of Taubes’ compactness theorems.

Theorem 3.6 ([Tau13b, WZ21, Par23a]).

Let (Ai,ϕi)(A_{i},\phi_{i}) be a sequence of solutions to (11), and let ri:=ϕiL2(M)r_{i}:=\|\phi_{i}\|_{L^{2}(M)}.

  • (i)

    If {ri}\{r_{i}\} is bounded, then there exists a subsequence of (Ai,ϕi)(A_{i},\phi_{i}) that converges in 𝒞\mathcal{C}^{\infty} after gauge transformations.

  • (ii)

    If limiri=+\lim_{i\to\infty}r_{i}=+\infty, then there exists a subsequence (which we still denote by (Ai,ϕi,ri)(A_{i},\phi_{i},r_{i})) and 𝐯/2\mathbf{v}\in\mathcal{M}_{\mathbb{Z}/2} such that ri2Tr(ϕiϕi)r_{i}^{-2}\mathrm{Tr}(\phi_{i}\otimes\phi_{i}) converges to 𝐯𝐯\mathbf{v}\otimes\mathbf{v} in 𝒞0\mathcal{C}^{0}.

Theorem 3.6 is implicitly contained in [Tau13b, WZ21, Par23a]. In the following, we deduce the statement of Theorem 3.6 from the above references.

Proof of Theorem 3.6.

Case (i) follows from standard elliptic bootstrapping. For case (ii), results in [Tau13b, WZ21, Par23a] imply that there exists a subsequence of (Ai,ϕi,ri)(A_{i},\phi_{i},r_{i}), which we denote by the same notation, and a 𝐯/2\mathbf{v}\in\mathcal{M}_{\mathbb{Z}/2} with zero locus ZZ, such that

  1. (1)

    |ϕi|/ri|\phi_{i}|/r_{i} converges to |𝐯||\mathbf{v}| in 𝒞0\mathcal{C}^{0}.

  2. (2)

    For every compact set KK contained in an open ball in MZM\setminus Z, there exists ϕ\phi on KK such that, after a sequence of gauge transformations, ϕi/ri\phi_{i}/r_{i} converges to ϕ\phi in the weak W2,2W^{2,2} topology on KK.

  3. (3)

    The spinor ϕ\phi in (ii) satisfies Tr(ϕϕ)=𝐯𝐯\mathrm{Tr}(\phi\otimes\phi)=\mathbf{v}\otimes\mathbf{v}.

Statements (1) and (2) above are directly given by [WZ21, Thm. 1.28]. A stronger result was proved in [Par23a, Thm. 1.3], where ϕi/ri\phi_{i}/r_{i} was shown to converge to ϕ\phi in the 𝒞\mathcal{C}^{\infty} topology on KK. A similar compactness result was given in [Tau13b, Thm. 1.1a] under weaker assumptions, but with weaker Sobolev regularity in the convergence statement. Statement (3) above follows from part (3) of the second bullet point of [Tau13b, Thm. 1.1a].

Now we prove Case (ii) of Theorem 3.6. For each ϵ>0\epsilon>0, let BϵB_{\epsilon} be the set of xMx\in M such that |𝐯𝐯|ϵ/4|\mathbf{v}\otimes\mathbf{v}|\leq\epsilon/4. Then BϵB_{\epsilon} is a closed subset of MM that contains an open neighborhood of ZZ. By Statement (1) above, for nn sufficiently large, we have

supBϵri2|Tr(ϕiϕi)|<ϵ/2.\sup_{B_{\epsilon}}r_{i}^{-2}|\mathrm{Tr}(\phi_{i}\otimes\phi_{i})|<\epsilon/2.

Hence the 𝒞0\mathcal{C}^{0} distance between ri2|Tr(ϕiϕi)|r_{i}^{-2}|\mathrm{Tr}(\phi_{i}\otimes\phi_{i})| and 𝐯𝐯\mathbf{v}\otimes\mathbf{v} is less than ϵ\epsilon on BϵB_{\epsilon} when ii is sufficiently large. By Statement (2) above and the Sobolev embedding theorems, after passing to a subsequence if necessary, the sequence ϕi/ri\phi_{i}/r_{i} converges to ϕ\phi on KK in the 𝒞0\mathcal{C}^{0} topology. Since Tr(ϕiϕi)\mathrm{Tr}(\phi_{i}\otimes\phi_{i}) is gauge invariant, the result is proved. ∎

Next, we define a compactification of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} using Theorem 3.6. Let 𝕄\mathbb{M} be the disjoint union of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} and /2\mathcal{M}_{\mathbb{Z}/2}. We define a topology on 𝕄\mathbb{M} as follows. The open sets on 𝕄\mathbb{M} are generated by the following two collections of subsets:

  1. (i)

    Open subsets of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})}.

  2. (ii)

    For 𝐯/2\mathbf{v}\in\mathcal{M}_{\mathbb{Z}/2}, ϵ>0\epsilon>0, N>0N>0, the subset of 𝕄\mathbb{M} that contains all 𝐯\mathbf{v}^{\prime} such that the 𝒞0\mathcal{C}^{0} distance between 𝐯𝐯\mathbf{v}\otimes\mathbf{v} and 𝐯𝐯\mathbf{v}^{\prime}\otimes\mathbf{v}^{\prime} is less than ϵ\epsilon, and all (A,ϕ)SL2()(A,\phi)\in\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} such that

    1. (a)

      ϕL2>N\|\phi\|_{L^{2}}>N,

    2. (b)

      the 𝒞0\mathcal{C}^{0} distance between ϕL22Tr(ϕϕ)\|\phi\|_{L^{2}}^{-2}\mathrm{Tr}(\phi\otimes\phi) and 𝐯𝐯\mathbf{v}\otimes\mathbf{v} is less than ϵ\epsilon.

Note that for 𝐯,𝐯/2\mathbf{v},\mathbf{v}^{\prime}\in\mathcal{M}_{\mathbb{Z}/2}, we have 𝐯=𝐯\mathbf{v}=\mathbf{v}^{\prime} if and only if 𝐯𝐯=𝐯𝐯\mathbf{v}\otimes\mathbf{v}=\mathbf{v}^{\prime}\otimes\mathbf{v}^{\prime}, and that the 𝒞0\mathcal{C}^{0} topology on /2\mathcal{M}_{\mathbb{Z}/2} is the same as the pull-back topology from 𝒞0(TMTM)\mathcal{C}^{0}(T^{*}M\otimes T^{*}M) via the map 𝐯𝐯𝐯\mathbf{v}\mapsto\mathbf{v}\otimes\mathbf{v}. Therefore, SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} and /2\mathcal{M}_{\mathbb{Z}/2} are homeomorphic to their embedded images in 𝕄\mathbb{M}. Moreover, Proposition 3.4 and Theorem 3.6 can be summarized into the following statement.

Corollary 3.7.

The space 𝕄\mathbb{M} is Hausdorff and compact.

Proof.

Since both SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} and /2\mathcal{M}_{\mathbb{Z}/2} are Hausdorff, the fact that 𝕄\mathbb{M} is Hausdorff follows straightforwardly from the definition of its topology. Proposition 3.4 and Theorem 3.6 imply that 𝕄\mathbb{M} is sequentially compact. It is also straightforward to verify that 𝕄\mathbb{M} is first countable. Hence 𝕄\mathbb{M} is compact. ∎

Definition 3.8.

Define ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} to be the closure of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} in 𝕄\mathbb{M}.

We call ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} the compactification of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})}.

Remark 3.9.

One can construct a more refined compactification of SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} from the analytical results in [Tau13b, HW15, WZ21, Par23a] by also considering the limits of the connection terms AiA_{i}. However, it is not clear to us how the convergence of the connection terms is related to the Morgan–Shalen compactification. See [OSWW20, HMNW23] for results in this direction in the two-dimensional case.

4. Measured foliations and /2\mathbb{Z}/2 harmonic 1-forms

In this section, we review a construction of measured foliations from /2\mathbb{Z}/2 harmonic 11–forms by Taubes [Tau13b, p. 14]. Then we prove that on simply connected manifolds, the leaf space of the measured foliation defined from a /2\mathbb{Z}/2 harmonic 11–form is always an \mathbb{R}–tree.

4.1. Measured foliations defined by /2\mathbb{Z}/2 harmonic 1-forms

The theory of measured foliations and quadratic differentials on Riemann surfaces has found significant applications in the geometry and topology of Riemann surfaces and 3-manifolds [Hub06, Hub16, Thu97]. For three- or four-dimensional Riemannian manifolds, concepts such as singular measured foliations, measured laminations, and weighted branched surfaces have been extensively developed in [HO96, GO89, Oer88]. We review the construction of a singular measured foliation from a /2\mathbb{Z}/2 harmonic 11–form, following Taubes [Tau13b, p. 14].

We first introduce the concept of a singular measured foliation.

Definition 4.1.

Let MM be a manifold with dimension nn. Let ZZ be a closed subset of Hausdorff codimension at least 22. A (codimension-one) singular foliation \mathcal{F} on MM with singular set ZZ is a smooth foliation on MZM\setminus Z with codimension-one leaves. A transverse measure μ\mu on \mathcal{F} is a measure for arcs in MZM\setminus Z such that the following conditions hold:

  1. (i)

    μ\mu is non-zero on transverse arcs.

  2. (ii)

    μ\mu vanishes on an arc if and only if the arc is tangent to a leaf.

  3. (iii)

    (Holonomy invariance) μ\mu is invariant along homotopies among transverse arcs that keep endpoints in the same leaves.

We will sometimes refer to a “singular foliation” as a “foliation” when there is no risk of confusion.

Now we associate a measured foliation to a /2\mathbb{Z}/2 harmonic 1-form. To provide some intuition, recall from Example 3.2 that, on a Riemann surface, a /2\mathbb{Z}/2 harmonic 1-form is the real part of the square root of a quadratic differential. Let us briefly review how a quadratic differential gives rise to a measured foliation.

Example 4.2.

Suppose MM is a closed Riemann surface. Given a nonzero holomorphic quadratic differential qq, the zeros Z:=q1(0)Z:=q^{-1}(0) form the singular set of a foliation. Away from ZZ, one can write qq locally as ω2\omega^{2} for some holomorphic 11–form ω\omega. The kernel of Reω\textrm{Re}\,\omega then defines a foliation, with a transverse measure given by |Reω||\textrm{Re}\,\omega|. This local construction can be combined to form a global measured foliation, often referred to as the vertical measured foliation of qq.

This construction can be generalized to /2\mathbb{Z}/2 harmonic 1-forms as in [Tau13a]. Consider a Riemannian manifold MM (not necessarily closed or complete), and let 𝐯\mathbf{v} be a /2\mathbb{Z}/2 harmonic 1-form on MM. We define the singular set as Z=|𝐯|1(0)Z=|\mathbf{v}|^{-1}(0). By Theorem 3.3, if the dimension of MM is no greater than 44, then ZZ meets the conditions described in Definition 4.1.

For a point xMZx\in M\setminus Z, let UxU_{x} be a small neighborhood around xx where 𝐯\mathbf{v} can be locally expressed as ±v\pm v, with vv a single-valued 1-form on UxU_{x}. By Definition 3.1, vv is a closed 1-form. The kernel of vv thus defines a codimension-one foliation on UxU_{x}, which is independent of the choice of sign for vv. Consequently, ker𝐯\ker\mathbf{v} defines a smooth foliation on MZM\setminus Z.

We define a transverse measure μ𝐯\mu_{\mathbf{v}} for this foliation as follows. For each 𝒞1\mathcal{C}^{1} arc γ:[0,1]MZ\gamma:[0,1]\to M\setminus Z, the transverse length of γ\gamma is given by

(12) μ𝐯(γ)=01|𝐯(γ˙(t))|𝑑t.\mu_{\mathbf{v}}(\gamma)=\int_{0}^{1}|\mathbf{v}(\dot{\gamma}(t))|dt.

It’s important to note that, similar to the cases of |𝐯||\mathbf{v}| and |𝐯||\nabla\mathbf{v}|, the value of |𝐯(γ˙(t))||\mathbf{v}(\dot{\gamma}(t))| is well-defined despite 𝐯\mathbf{v} being a two-valued section. Since 𝐯\mathbf{v} is locally given by a closed smooth form in MZM\setminus Z, the measure μ𝐯\mu_{\mathbf{v}} is holonomy invariant. As a result, the pair (ker𝐯,μ𝐯)(\ker\mathbf{v},\mu_{\mathbf{v}}) defines a measured foliation on MZM\setminus Z.

4.2. The leaf spaces of /2\mathbb{Z}/2 harmonic 1-forms

Next, we define the leaf space associated to a /2\mathbb{Z}/2 harmonic form. We prove that if the background manifold is simply connected, then the leaf space is an \mathbb{R}–tree.

Consider a connected Riemannian manifold MM (not necessarily closed or complete), and let 𝐯\mathbf{v} be a /2\mathbb{Z}/2 harmonic 11–form on MM. We define a pseudo-metric on MM by

(13) dM,𝐯(x,y)=infγ01|𝐯(γ˙(t))|𝑑t,d_{M,\mathbf{v}}(x,y)=\inf_{\gamma}\int_{0}^{1}|\mathbf{v}(\dot{\gamma}(t))|dt,

where γ\gamma ranges over all piecewise 𝒞1\mathcal{C}^{1} curves from xx to yy. Note that the definition of dM,𝐯d_{M,\mathbf{v}} depends globally on both MM and 𝐯\mathbf{v}: if UU is an open subset of MM, then it is not necessarily true that dU,𝐯d_{U,\mathbf{v}} equals the restriction of dM,𝐯d_{M,\mathbf{v}} on UU.

Definition 4.3.

The metric space 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} is the quotient space of MM with respect to the pseudo-metric dM,𝐯d_{M,\mathbf{v}}.

We call the space 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} the leaf space of ker𝐯\ker\mathbf{v} on MM, and we will use dM,𝐯d_{M,\mathbf{v}} to denote the metric on this quotient space as well.

Theorem 4.4.

Suppose MM is a simply connected manifold, 𝐯\mathbf{v} is a /2\mathbb{Z}/2 harmonic 11–form on MM with zero locus ZZ. If the dimension of MM is greater than 44, we assume in addition that 𝐯\mathbf{v} is continuous and ZZ is rectifiable with locally finite Hausdorff measure in codimension 22. Then the metric space (𝒯M,𝐯,dM,𝐯)(\mathcal{T}_{M,\mathbf{v}},d_{M,\mathbf{v}}) is an \mathbb{R}–tree.

The remainder of this section is devoted to the proof of Theorem 4.4. We first recall the following standard result about \mathbb{R}–trees.

Proposition 4.5 ([Py23, Prop. B.31]).

A metric space (X,d)(X,d) is an \mathbb{R}–tree if and only if it is path connected and the following inequality holds for all x,y,z,tXx,y,z,t\in X:

d(x,z)+d(y,t)max{d(x,y)+d(z,t),d(x,t)+d(y,z)}.d(x,z)+d(y,t)\leq\max\{d(x,y)+d(z,t),d(x,t)+d(y,z)\}.

Now we prove the following technical lemma.

Lemma 4.6.

Suppose DD is the closed unit disk in 2\mathbb{R}^{2}, and let int(D)int(D) denote the interior of DD. Suppose 𝐯\mathbf{v} is a continuous, two–valued 11–form defined on an open neighborhood of DD in 2\mathbb{R}^{2} with zero locus ZZ, such that away from its zero points 𝐯\mathbf{v} is locally given by ±v\pm v with vv being closed. We further assume that ZDZ\cap\partial D is finite, Zint(D)Z\cap int(D) is compact, and that for each δ>0\delta>0, there exists a smooth domain Ωδint(D)\Omega_{\delta}\subset int(D) such that Ωδ\Omega_{\delta} contains Zint(D)Z\cap int(D), and Ωδ\Omega_{\delta} is contained in the δ\delta–neighborhood of Zint(D)Z\cap int(D), and the total length (i.e. the Hausdorff measure in dimension 11) of Ωδ\partial\Omega_{\delta} is less than 11. Finally, assume that ker𝐯\ker\mathbf{v} is transverse to D\partial D at all but finitely many points.

Let dD,𝐯d_{D,\mathbf{v}} be the pseudo-distance function on DD associated with 𝐯\mathbf{v} given by (13). Let p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} be four distinct points appearing in cyclic order on D\partial D, the points p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} divide D\partial D into four arcs.

Then there exist two points aa and bb on a pair of opposite (closed) arcs divided by p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} such that dD,𝐯(a,b)=0d_{D,\mathbf{v}}(a,b)=0.

Proof.

We adapt an argument from [Lev93, Lem. III.4]. To simplify notation, we will write dD,𝐯d_{D,\mathbf{v}} as dd in the proof. Assume p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} are ordered counterclockwise. In the following, for x,yDx,y\in\partial D, we use xy¯\overline{xy} to denote the closed arc in D\partial D bounded by x,yx,y, where the arc goes from xx to yy in the counterclockwise direction. If A,BA,B are compact subsets of DD, we use d(A,B)d(A,B) to denote infaA,bBd(a,b)\inf_{a\in A,b\in B}d(a,b).

If d(p1p2¯,p3p4¯)=0d(\overline{p_{1}p_{2}},\overline{p_{3}p_{4}})=0, then the desired result already holds. In the following, assume d(p1p2¯,p3p4¯)>0d(\overline{p_{1}p_{2}},\overline{p_{3}p_{4}})>0. Let qp1p2¯q\in\overline{p_{1}p_{2}} be such that d(q,p4p1¯)=0d(q,\overline{p_{4}p_{1}})=0 and that qq is furthest away from p1p_{1} as a point on the arc p1p2¯\overline{p_{1}p_{2}} among all points satisfying this condition. Similarly, let rp1p2¯r\in\overline{p_{1}p_{2}} be such that d(r,p2p3¯)=0d(r,\overline{p_{2}p_{3}})=0 and that rr is furthest away from p2p_{2} among all points in p1p2¯\overline{p_{1}p_{2}} satisfying this condition.

We discuss three cases. If q=rq=r, then the above conditions imply that d(p4p1¯,p2p3¯)=0d(\overline{p_{4}p_{1}},\overline{p_{2}p_{3}})=0, and the desired result holds.

If rqr\neq q and rq¯p1p2¯\overline{rq}\subset\overline{p_{1}p_{2}} (that is, qq is on the counterclockwise side of rr), then by the above conditions, for every ϵ>0\epsilon>0, there exists a 𝒞1\mathcal{C}^{1} curve from rr to p2p3¯\overline{p_{2}p_{3}}, and a 𝒞1\mathcal{C}^{1} curve from qq to p4p1¯\overline{p_{4}p_{1}}, whose lengths with respect to dd are less than ϵ\epsilon. These two curves must intersect, so d(q,r)<2ϵd(q,r)<2\epsilon. Since this statement holds for all ϵ>0\epsilon>0, we conclude that d(q,r)=0d(q,r)=0, which implies d(p4p1¯,p2p3¯)=0d(\overline{p_{4}p_{1}},\overline{p_{2}p_{3}})=0, and the desired result holds.

If rqr\neq q and qr¯p1p2¯\overline{qr}\subset\overline{p_{1}p_{2}} (that is, rr is on the counterclockwise side of qq), we consider the singular foliation \mathcal{F} defined by ker𝐯\ker\mathbf{v} on DD. Since 𝐯\mathbf{v} is locally given by closed forms away from zero points, there is a transverse invariant measure μ𝐯\mu_{\mathbf{v}} on \mathcal{F} as given by (12).

For each point xqr¯{q,r}x\in\overline{qr}\setminus\{q,r\}, if \mathcal{F} is defined and is transverse to D\partial D at xx, we consider the leaf of \mathcal{F} passing through xx. View the leaf as a parametrized curve starting at xx. By the Poincaré–Bendixson theorem, one of the following holds:

  1. (1)

    The leaf intersects a boundary point of D\partial D other than xx.

  2. (2)

    The leaf converges to a zero point of 𝐯\mathbf{v}.

  3. (3)

    The leaf converges to a limit cycle.

Since \mathcal{F} admits a transverse measure, Case (3) cannot happen. By the definitions of q,rq,r, we know that d({x},rp3¯)>0d(\{x\},\overline{rp_{3}})>0, d({x},p4q¯)>0d(\{x\},\overline{p_{4}q})>0. By the assumption that d(p1p2¯,p3p4¯)>0d(\overline{p_{1}p_{2}},\overline{p_{3}p_{4}})>0, we know that d(x,p3p4¯)>0d(x,\overline{p_{3}p_{4}})>0. As a result, if Case (1) happens, then the intersection point of the leaf with D\partial D is in the interior of qr¯\overline{qr}.

Let Uqr¯{q,r}U\subset\overline{qr}\setminus\{q,r\} be the set of points xx such that

  1. (1)

    𝐯\mathbf{v} is non-zero at xx and =ker𝐯\mathcal{F}=\ker\mathbf{v} is transverse to D\partial D at xx.

  2. (2)

    The leaf of \mathcal{F} starting at xx transversely intersects a point of qr¯{q,r}\overline{qr}\setminus\{q,r\} other than xx.

Then UU is an open subset of qr¯{q,r}\overline{qr}\setminus\{q,r\}. We claim that

(14) μ𝐯(U)=μ𝐯(qr¯).\mu_{\mathbf{v}}(U)=\mu_{\mathbf{v}}(\overline{qr}).

This is because the leafs emanating from the zero points of 𝐯\mathbf{v} have zero measure with respect to μ𝐯\mu_{\mathbf{v}}. More precisely, recall that ZZ denotes the zero set of 𝐯\mathbf{v}. For each ϵ>0\epsilon>0, there exists δ\delta such that |𝐯|<ϵ|\mathbf{v}|<\epsilon on the δ\delta–neighborhood of ZZ. Let Ωδ\Omega_{\delta} be the smooth domain given by the assumptions on ZZ, then μ𝐯(Ωδ)<ϵ\mu_{\mathbf{v}}(\partial\Omega_{\delta})<\epsilon. Then the set of points xqr¯x\in\overline{qr} such that there is leaf of \mathcal{F} intersecting D\partial D transversely at xx and enters Ωδ\Omega_{\delta} has μ𝐯\mu_{\mathbf{v}}–measure less than ϵ\epsilon. Similarly, if pp is a zero point of 𝐯\mathbf{v} in D\partial D, then μ𝐯(Br(p)D)0\mu_{\mathbf{v}}(\partial B_{r}(p)\cap D)\to 0 as r0r\to 0, so the set of points xqr¯x\in\overline{qr} such that there is leaf of \mathcal{F} intersecting D\partial D transversely at xx and converges to pp on the other end has μ𝐯\mu_{\mathbf{v}}–measure zero. Recall that for every point xqr¯({q,r}U)x\in\overline{qr}\setminus(\{q,r\}\cup U), one of the following conditions holds:

  1. (1)

    xx is a zero point of 𝐯\mathbf{v},

  2. (2)

    xx is a point where \mathcal{F} is tangent to D\partial D,

  3. (3)

    xx is a point whose leaf passes through a tangent point of \mathcal{F} with D\partial D,

  4. (4)

    xx is a point whose leaf converges to a zero point of 𝐯\mathbf{v}.

The first three cases only contain finitely many points, and points in the last case have measure zero with respect to μ𝐯\mu_{\mathbf{v}} by the previous argument. Hence Equation (14) is proved. Since ker𝐯\ker\mathbf{v} is transverse to D\partial D at all but finitely many points, we conclude that UU has full measure in qr¯\overline{qr} with respect to the standard Lebesgue measure as well.

Define an involution ι:UU\iota:U\to U, such that for every xUx\in U, the image ι(x)U\iota(x)\in U is the other endpoint of the leaf of \mathcal{F} passing through xx. The map ι\iota is orientation-reversing.

Parameterize the arc qr¯\overline{qr} by the interval [0,1][0,1] via a smooth diffeomorphism: φ:[0,1]qr¯\varphi:[0,1]\to\overline{qr}. Define a function

ξ:[0,1]:td(q,φ(t)),\xi:[0,1]\to\mathbb{R}:t\mapsto d(q,\varphi(t)),

where we recall that dd denotes the pseudo-distance function dD,𝐯d_{D,\mathbf{v}}. Then ξ\xi a Lipschitz function on [0,1][0,1], so ξ\xi^{\prime} exists almost everywhere and

ξ(1)ξ(0)=01ξ.\xi(1)-\xi(0)=\int_{0}^{1}\xi^{\prime}.

For each open interval (s,t)φ1(U)(s,t)\in\varphi^{-1}(U), let s^=φ1ιφ(s)\hat{s}=\varphi^{-1}\circ\iota\circ\varphi(s) and t^=φ1ιφ(t)\hat{t}=\varphi^{-1}\circ\iota\circ\varphi(t), we have

stξ=d(q,φ(t))d(q,φ(s))=d(q,φ(t^))d(q,φ(s^))=t^s^ξ.\int_{s}^{t}\xi^{\prime}=d(q,\varphi(t))-d(q,\varphi(s))=d(q,\varphi(\hat{t}))-d(q,\varphi(\hat{s}))=-\int_{\hat{t}}^{\hat{s}}\xi^{\prime}.

Hence the involution φ1ιφ\varphi^{-1}\circ\iota\circ\varphi on φ1(U)\varphi^{-1}(U) reverses the orientation and preserves the differential form ξ(t)dt\xi^{\prime}(t)dt. This implies φ1(U)ξ=0\int_{\varphi^{-1}(U)}\xi^{\prime}=0. Since φ1(U)\varphi^{-1}(U) has full Lebesgue measure in [0,1][0,1], we conclude that 01ξ=0\int_{0}^{1}\xi^{\prime}=0, so d(q,r)=ξ(1)=0d(q,r)=\xi(1)=0. This contradicts the definitions of q,rq,r. ∎

We now prove Theorem 4.4.

Proof of Theorem 4.4.

Let p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} be distinct points in MM. By Lemma 4.6, we only need to show that

(15) dM,𝐯(p1,p3)+dM,𝐯(p2,p4)max{dM,𝐯(p1,p2)+dM,𝐯(p3,p4),dM,𝐯(p1,p4)+dM,𝐯(p2,p3)}.\begin{split}&d_{M,\mathbf{v}}(p_{1},p_{3})+d_{M,\mathbf{v}}(p_{2},p_{4})\\ \leq&\max\{d_{M,\mathbf{v}}(p_{1},p_{2})+d_{M,\mathbf{v}}(p_{3},p_{4}),d_{M,\mathbf{v}}(p_{1},p_{4})+d_{M,\mathbf{v}}(p_{2},p_{3})\}.\end{split}

Since (15) is a closed condition on (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}), we may assume without loss of generality that p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} are not zero points of 𝐯\mathbf{v}. For notational convenience, we will interpret the subscripts modulo 44. Fix ϵ>0\epsilon>0. For each ii, let γi\gamma_{i} be a smooth arc from pip_{i} to pi+1p_{i+1} such that

μ𝐯(γi)dM,𝐯(pi,pi+1)+ϵ.\mu_{\mathbf{v}}(\gamma_{i})\leq d_{M,\mathbf{v}}(p_{i},p_{i+1})+\epsilon.

After perturbation, we may assume that the union of γ1,,γ4\gamma_{1},\dots,\gamma_{4} forms a smoothly embedded circle in MM that is disjoint from the zero locus of 𝐯\mathbf{v}. Denote this circle by CC. After a further perturbation if necessary, we may assume that CC is transverse to ker𝐯\ker\mathbf{v} at all but finitely many points. Since MM is simply connected, there exists a smooth immersion f:DMf:D\to M, where DD is the unit disk in 2\mathbb{R}^{2}, such that f(D)=Cf(\partial D)=C. Since the zero locus of 𝐯\mathbf{v} has locally finite codimension two Hausdorff measure, we may perturb ff, with f|Df|_{\partial D} fixed, so that f(D)f(D) intersects the zero locus of 𝐯\mathbf{v} at finitely many points, and that the zero locus of f(𝐯)f^{*}(\mathbf{v}) is discrete on the complement of (f|𝐯|)1(0)(f\circ|\mathbf{v}|)^{-1}(0). As a result, the assumptions of Lemma 4.6 holds for f(𝐯)f^{*}(\mathbf{v}).

Applying Lemma 4.6 to f1(pi)f^{-1}(p_{i}) for i=1,2,3,4i=1,2,3,4 and to the 22–valued form f(𝐯)f^{*}(\mathbf{v}), we conclude that there exists a,bDa,b\in\partial D on opposite arcs such that dD,f(𝐯)(a,b)=0d_{D,f^{*}(\mathbf{v})}(a,b)=0. Since dM,𝐯(a,b)dD,f(𝐯)(a,b)d_{M,\mathbf{v}}(a,b)\leq d_{D,f^{*}(\mathbf{v})}(a,b), this implies dM,𝐯(a,b)=0d_{M,\mathbf{v}}(a,b)=0. Assume without loss of generality that aa is in the arc bounded by p1,p2p_{1},p_{2}, and bb is in the arc bounded by p3,p4p_{3},p_{4}. Then we have

dM,𝐯(p1,p3)dM,𝐯(p1,a)+dM,𝐯(b,p3),dM,𝐯(p2,p4)dM,𝐯(p2,a)+dM,𝐯(b,p4).d_{M,\mathbf{v}}(p_{1},p_{3})\leq d_{M,\mathbf{v}}(p_{1},a)+d_{M,\mathbf{v}}(b,p_{3}),\qquad d_{M,\mathbf{v}}(p_{2},p_{4})\leq d_{M,\mathbf{v}}(p_{2},a)+d_{M,\mathbf{v}}(b,p_{4}).

By the assumptions on f(D)f(\partial D), we have

dM,𝐯(p1,a)+dM,𝐯(p2,a)μ𝐯(γ1)dM,𝐯(p1,p2)+ϵ,d_{M,\mathbf{v}}(p_{1},a)+d_{M,\mathbf{v}}(p_{2},a)\leq\mu_{\mathbf{v}}(\gamma_{1})\leq d_{M,\mathbf{v}}(p_{1},p_{2})+\epsilon,

and similarly,

dM,𝐯(p3,b)+dM,𝐯(p4,b)dM,𝐯(p3,p4)+ϵ.d_{M,\mathbf{v}}(p_{3},b)+d_{M,\mathbf{v}}(p_{4},b)\leq d_{M,\mathbf{v}}(p_{3},p_{4})+\epsilon.

As a result,

dM,𝐯(p1,p3)+dM,𝐯(p2,p4)dM,𝐯(p1,p2)+dM,𝐯(p3,p4)+2ϵ.d_{M,\mathbf{v}}(p_{1},p_{3})+d_{M,\mathbf{v}}(p_{2},p_{4})\leq d_{M,\mathbf{v}}(p_{1},p_{2})+d_{M,\mathbf{v}}(p_{3},p_{4})+2\epsilon.

The desired result then follows by taking ϵ0\epsilon\to 0. ∎

5. Harmonic maps to \mathbb{R}–trees

Let MM be a closed 33–manifold and M~\widetilde{M} its universal cover. In this section, we show that if u:M~𝒯u:\widetilde{M}\to\mathcal{T} is a π1(M)\pi_{1}(M)–equivariant harmonic map to an \mathbb{R}–tree, then the gradient of uu defines a /2\mathbb{Z}/2 harmonic 1-form on MM. We first recall the formulation of Korevaar–Schoen which gives meaning to the gradient of uu as a Radon–Nikodym derivative. Then, we use regularity results of harmonic maps to construct the associated /2\mathbb{Z}/2 harmonic 11–form.

5.1. Energy density and directional derivatives

Here, we review some terminology from [KS93], specialized to the case of L2L^{2} norms. Following the notation of [KS93], let (Ω,g)(\Omega,g) be an nn-dimensional Riemannian manifold and (X,d)(X,d) a complete metric space. Unless otherwise specified, Ω\Omega is allowed to be non-compact or non-complete but we assume that Ω\Omega has no boundary. Later, we will take Ω\Omega to be the universal cover of a closed 33–manifold and (X,d)(X,d) to be an \mathbb{R}–tree.

For ϵ>0\epsilon>0, let Ωϵ\Omega_{\epsilon} be the set of all points xΩx\in\Omega such that the exponential map at xx is well-defined on the open ball with radius ϵ\epsilon. Let μ\mu denote the volume measure of Ω\Omega.

A map u:ΩXu:\Omega\to X is called locally L2L^{2} if for every compact subset KK in Ω\Omega and every pXp\in X, we have

Kd2(u(x),p)𝑑μ(x)<+.\int_{K}d^{2}(u(x),p)\,d\mu(x)<+\infty.

If uu is locally L2L^{2}, define the ϵ\epsilon–approximate energy function eϵ:Ωϵe_{\epsilon}:\Omega_{\epsilon}\to\mathbb{R} as

(16) eϵ(x):=1ωn1ϵn1Bϵ(x)d2(u(x),u(y))ϵ2𝑑σ,e_{\epsilon}(x):=\frac{1}{\omega_{n-1}\epsilon^{n-1}}\int_{\partial B_{\epsilon}(x)}\frac{d^{2}(u(x),u(y))}{\epsilon^{2}}\,d\sigma,

where ωn1\omega_{n-1} is the area of Sn1S^{n-1} in n\mathbb{R}^{n} (recall that nn is the dimension of Ω\Omega), and dσd\sigma is the area form on Bϵ(x)\partial B_{\epsilon}(x). The map uu is said to have finite energy, if

supφ𝒞0(M),0φ1lim supϵ0Ωeϵφ𝑑μ<+.\sup_{\varphi\in\mathcal{C}_{0}^{\infty}(M),0\leq\varphi\leq 1}\limsup_{\epsilon\to 0}\int_{\Omega}e_{\epsilon}\cdot\varphi\,d\mu<+\infty.

The space of maps from Ω\Omega to XX with finite energy is denoted by W1,2(Ω,X)W^{1,2}(\Omega,X). By [KS93, Thm.  1.10], if uW1,2(Ω,X)u\in W^{1,2}(\Omega,X), then the measures eϵ(x)μ(x)e_{\epsilon}(x)\mu(x) converge weakly to a limit e(x)μ(x)e(x)\mu(x) as ϵ0\epsilon\to 0, where e(x)L1(Ω)e(x)\in L^{1}(\Omega). The function e(x)e(x) is called the energy density function, and we define |u|(x):=e(x)1/2L2(Ω)|\nabla u|(x):=e(x)^{1/2}\in L^{2}(\Omega).

The directional derivatives are defined in a similar way. Assume ZZ is a Lipschitz vector field over Ω\Omega. For xΩx\in\Omega, let x+ϵZx+\epsilon Z denote the image of xx after flowing along ZZ for time ϵ\epsilon. Define eϵZ(x)=d2(u(x),u(x+ϵZ))/ϵ2{}^{Z}e_{\epsilon}(x)=d^{2}(u(x),u(x+\epsilon Z))/\epsilon^{2}. By [KS93], if uW1,2(Ω,X)u\in W^{1,2}(\Omega,X), then there exists a unique non-negative function |u(Z)|L2(Ω)|u_{*}(Z)|\in L^{2}(\Omega) such that the measures eϵZ(x)μ(x){}^{Z}e_{\epsilon}(x)\mu(x) converge weakly to |u(Z)|2(x)μ(x)|u_{*}(Z)|^{2}(x)\mu(x).

A map uWloc1,2(Ω,)u\in W^{1,2}_{loc}(\Omega,\mathbb{R}) is called harmonic, if it is a critical point of the Dirichlet functional (we refer the reader to [KS93, Sec. 2.2] for more details). If the target (X,d)(X,d) is an \mathbb{R}–tree, then harmonic maps to XX minimizes the energy with respect to compactly supported perturbations.

Now assume MM is a closed manifold and let M~\widetilde{M} be its universal cover. Assume (𝒯,d𝒯)(\mathcal{T},d_{\mathcal{T}}) is a complete \mathbb{R}–tree with a fixed isometric action by π1(M)\pi_{1}(M). The following results established the existence and uniqueness of equivariant harmonic maps. The existence result follows from the work of Korevaar–Schoen [KS93, KS97], and the uniqueness follows from [Mes02]. See also [DM21] for generalizations.

Theorem 5.1 ([KS93, KS97, Mes02, DM21]).

Suppose the π1(M)\pi_{1}(M) action on a complete \mathbb{R} tree (𝒯,d𝒯)(\mathcal{T},d_{\mathcal{T}}) has no fixed ends (see the definitions above Theorem 2.1). Then there exists a π1(M)\pi_{1}(M)–equivariant harmonic map u:M~𝒯u:\widetilde{M}\to\mathcal{T}. Moreover, if u0,u1:M~𝒯u_{0},u_{1}:\widetilde{M}\to\mathcal{T} are two π1(M)\pi_{1}(M)–equivariant harmonic maps such that u0u1u_{0}\neq u_{1}, then either u0(M~)u_{0}(\widetilde{M}) or u1(M~)u_{1}(\widetilde{M}) is contained in a geodesic.

Mese [Mes02] showed that distinct harmonic maps always have the same directional derivatives.

Proposition 5.2 ([Mes02, Cor. 13]).

Fix a π1(M)\pi_{1}(M) action on an \mathbb{R}–tree (𝒯,d𝒯)(\mathcal{T},d_{\mathcal{T}}). Let u0,u1:M~𝒯u_{0},u_{1}:\widetilde{M}\to\mathcal{T} be two equivariant harmonic maps to 𝒯\mathcal{T}, and let ZZ be a Lipschitz vector field on M~\widetilde{M}. Then we have |(u0)Z|=|(u1)Z||(u_{0})_{*}Z|=|(u_{1})_{*}Z| (almost everywhere).

5.2. Regularity of harmonic maps to \mathbb{R}–trees

Now we review some regularity results on harmonic maps to \mathbb{R}–trees from the literature.

Definition 5.3.

Let u:Ω𝒯u:\Omega\to\mathcal{T} be a harmonic map from a manifold to a tree. We say that pΩp\in\Omega is a regular point of uu, if there exists an open ball Br(p)B_{r}(p) centered at pp such that u(Br(p))u(B_{r}(p)) is contained in a geodesic in 𝒯\mathcal{T}. If pp is not a regular point, then it is called a singular point. We denote the set of regular points by (u)\mathcal{R}(u) and the set of singular points by 𝒮(u)\mathcal{S}(u).

If pp is a regular point of uu, then locally uu is equal to the composition of a harmonic map to a closed interval II\subset\mathbb{R}, and an isometric embedding of II in 𝒯\mathcal{T}. It is clear from the definition that (u)\mathcal{R}(u) is open and 𝒮(u)\mathcal{S}(u) is closed.

Theorem 5.4 ([Sun03, Thm. 1.4]).

Let u:Ω𝒯u:\Omega\to\mathcal{T} be a harmonic map from a manifold to an \mathbb{R}–tree. Then, for every point pΩp\in\Omega, there exists an open ball Br(p)B_{r}(p) centered at pp such that u(Br(p))u(B_{r}(p)) lies in an embedded locally finite subtree.

Theorem 5.4 allows one to reduce the regularity problem of harmonic maps into \mathbb{R}–trees to the case where the tree is locally finite. Before introducing the next results, we review the definition of the order function from [GS92, Sec. 2].

Definition 5.5.

Assume u:Ω𝒯u:\Omega\to\mathcal{T} is harmonic. For pΩp\in\Omega, the order of uu at pp is defined to be the limit

(17) ord(u;p):=limr0rBr(p)|u(x)|2μ(x)Br(p)d2(u(p),u(x))σ(x),\mathrm{ord}(u;p):=\lim_{r\to 0}\frac{r\int_{B_{r}(p)}|\nabla u(x)|^{2}\,\mu(x)}{\int_{\partial B_{r}(p)}d^{2}(u(p),u(x))\,\sigma(x)},

where μ(x)\mu(x) denotes the volume measure of Ω\Omega and σ(x)\sigma(x) denotes the area form.

It was proved in [GS92, Sec. 2] that the limit in (17) always exists. When 𝒯=\mathcal{T}=\mathbb{R}, the value of ord(u;p)\mathrm{ord}(u;p) equals the vanishing order of uu(p)u-u(p) at pp. Moreover, the order function ord(u;p)\mathrm{ord}(u;p) is upper semi-continuous with respect to pp ([GS92, statement above Thm. 2.3]).

We collect the following regularity results from the literature, which will be used later. In the following, u:Ω𝒯u:\Omega\to\mathcal{T} denotes a harmonic map, and dd denotes the metric on the \mathbb{R}–tree 𝒯\mathcal{T}.

Theorem 5.6 ([GS92, Thm. 6.3], [Sun03, Thm. 1.1]).

There is a constant ϵ\epsilon depending only on the dimension of the domain, such that for each point pΩp\in\Omega, either ord(u;p)=1\mathrm{ord}(u;p)=1, or ord(u;p)1+ϵ\mathrm{ord}(u;p)\geq 1+\epsilon. Moreover, if ord(u;p)=1\mathrm{ord}(u;p)=1, then pp is a regular point.

Theorem 5.7 ([KS93, Thm. 2.4.6]).

The map uu is locally Lipschitz, and the Lipschitz constants only depend on the curvature of Ω\Omega.

Theorem 5.8 ([Sun03, Thm. 1.3], [GS92, Thm. 6.4]).

The singular set 𝒮(u)\mathcal{S}(u) has Hausdorff codimension at least 22.

A stronger version of this result was given by Dees [Dee22].

Theorem 5.9 ([Dee22, Thms. 1.1 and 1.2]).

Assume Ω\Omega has dimension nn, then 𝒮(u)\mathcal{S}(u) is (n2)(n-2)–rectifiable. Moreover, for each compact set KΩK\subset\Omega, there exists a constant CC depending on KK, uu, and Ω\Omega, such that

Vol(Br(𝒮(u)K))Cr2\mathrm{Vol}(B_{r}(\mathcal{S}(u)\cap K))\leq C\cdot r^{2}

for r<1r<1, where Br(𝒮(u)K)B_{r}(\mathcal{S}(u)\cap K) denotes rr–neighborhood of 𝒮(u)K\mathcal{S}(u)\cap K.

Theorem 5.10 ([GS92, proof of Thm. 2.3]).

Assume Ω\Omega as a manifold is the open unit ball B1(0)B_{1}(0) in n\mathbb{R}^{n}, but the metric is not necessarily Euclidean. Suppose ord(u;0)=α\mathrm{ord}(u;0)=\alpha. Then there exists a constant CC depending only on the curvature, such that for all λ[0,1]\lambda\in[0,1] and xB1/2(0)x\in B_{1/2}(0), we have

(18) d(u(λx),u(0))Cλαd(u(x),u(0)).d(u(\lambda x),u(0))\leq C\lambda^{\alpha}d(u(x),u(0)).
Corollary 5.11.

Assume ord(u;p)=α\mathrm{ord}(u;p)=\alpha. Let nn be the dimension of Ω\Omega. Let r0r_{0} be the minimum of 11 and the injectivity radius of Ω\Omega at pp. Assume r<r0/2r<r_{0}/2. Then there exists a constant CC depending only on the curvature of Ω\Omega in Br0(p)B_{r_{0}}(p), such that

Br(p)|u|2Crn2+2αr022α.\int_{B_{r}(p)}|\nabla u|^{2}\leq Cr^{n-2+2\alpha}\cdot r_{0}^{2-2\alpha}.
Proof.

By [GS92, Sec. 2], there exists a constant C1C_{1} depending only on the curvature, such that

eC1r2rBr(p)|u|2Br(p)d2(u(p),u)e^{C_{1}r^{2}}\frac{r\int_{B_{r}(p)}|\nabla u|^{2}}{\int_{\partial B_{r}(p)}d^{2}(u(p),u)}

is increasing with respect to rr for all r<r0r<r_{0}. Hence

eC1r2rBr(p)|u|2Br(p)d2(u(p),u)eC1r02r0Br0(p)|u|2Br0(p)d2(u(p),u).e^{C_{1}r^{2}}\frac{r\int_{B_{r}(p)}|\nabla u|^{2}}{\int_{\partial B_{r}(p)}d^{2}(u(p),u)}\leq e^{C_{1}r_{0}^{2}}\frac{r_{0}\int_{B_{r_{0}}(p)}|\nabla u|^{2}}{\int_{\partial B_{r_{0}}(p)}d^{2}(u(p),u)}.

By Theorem 5.7, we have r0Br0(p)|u|2C2r0n+1r_{0}\int_{B_{r_{0}}(p)}|\nabla u|^{2}\leq C_{2}{r_{0}}^{n+1} for a constant C2C_{2}. By Theorem 5.10, we have

Br(p)d2(u(p),u)Br0(p)d2(u(p),u)(rr0)n1(rr0)2α,\frac{\int_{\partial B_{r}(p)}d^{2}(u(p),u)}{\int_{\partial B_{r_{0}}(p)}d^{2}(u(p),u)}\leq\left(\frac{r}{r_{0}}\right)^{n-1}\left(\frac{r}{r_{0}}\right)^{2\alpha},

where the first factor comes from the comparison of area forms. The desired result then follows from a straightforward computation. ∎

We also need the following estimate.

Theorem 5.12 ([GS92, Thm. 2.4]).

Assume pΩp\in\Omega and rr is less than the injectivity radius of Ω\Omega at pp. Then there exists a constant CC, depending only on the curvature of Ω\Omega on Br(p)B_{r}(p), such that

uL(Br/2(p))2CBr(p)|u|2.\|\nabla u\|_{L^{\infty}(B_{r/2}(p))}^{2}\leq C\fint_{B_{r}(p)}|\nabla u|^{2}.
Remark 5.13.

Note that the definitions in Section 5.1 only defined |u||\nabla u| almost everywhere on Ω\Omega. On (u)\mathcal{R}(u), the function |u||\nabla u| is represented by a smooth function. By Theorem 5.12 and Corollary 5.11, we see that the function |u||\nabla u| is given by a continuous function on Ω\Omega, where |u|=0|\nabla u|=0 on 𝒮(u)\mathcal{S}(u).

We now prove the following estimate.

Proposition 5.14.

Assume KK is a compact subset of Ω\Omega. Then

K𝒮(u)|u|2<+.\int_{K\setminus\mathcal{S}(u)}|\nabla\nabla u|^{2}<+\infty.
Proof.

By Theorem 5.8, 𝒮(u)\mathcal{S}(u) has zero Lebesgue measure. Hence we may write Ω𝒮(u)\int_{\Omega\setminus\mathcal{S}(u)} as Ω\int_{\Omega}, if the integrand is defined on (u)\mathcal{R}(u) and extends to an integrable function on Ω\Omega.

By shrinking Ω\Omega to a smaller domain containing KK, we may assume that the curvature of Ω\Omega is uniformly bounded and Ω\Omega has finite volume. By Theorem 5.7, this implies |u||\nabla u| is bounded and Ω|u|2<+\int_{\Omega}|\nabla u|^{2}<+\infty.

By [GS92, eq. (6.2)], there is a constant CC depending only on the curvature of Ω\Omega such that

(19) 12Δ|u|2|u|2C|u|2\frac{1}{2}\Delta|\nabla u|^{2}\geq|\nabla\nabla u|^{2}-C|\nabla u|^{2}

pointwise on (u)\mathcal{R}(u).

Let ρ\rho be a Lipschitz function that is compactly supported in (u)\mathcal{R}(u) such that 0ρ10\leq\rho\leq 1. By (19),

12ΩΔ|u|2ρ2Ω|u|2ρ2CΩ|u|2ρ2,\frac{1}{2}\int_{\Omega}\Delta|\nabla u|^{2}\rho^{2}\geq\int_{\Omega}|\nabla\nabla u|^{2}\rho^{2}-C\int_{\Omega}|\nabla u|^{2}\rho^{2},

so

Ω|u|2ρ2\displaystyle\int_{\Omega}|\nabla\nabla u|^{2}\rho^{2} CΩ|u|2ρ2+12ΩΔ|u|2ρ2\displaystyle\leq C\int_{\Omega}|\nabla u|^{2}\rho^{2}+\frac{1}{2}\int_{\Omega}\Delta|\nabla u|^{2}\rho^{2}
=CΩ|u|2ρ212Ω|u|2,(ρ2)\displaystyle=C\int_{\Omega}|\nabla u|^{2}\rho^{2}-\frac{1}{2}\int_{\Omega}\langle\nabla|\nabla u|^{2},\nabla(\rho^{2})\rangle
CΩ|u|2ρ2+2Ω||u|||u||ρ|ρ\displaystyle\leq C\int_{\Omega}|\nabla u|^{2}\rho^{2}+2\int_{\Omega}|\nabla|\nabla u||\cdot|\nabla u|\cdot|\nabla\rho|\cdot\rho
CΩ|u|2ρ2+2Ω|u||u||ρ|ρ\displaystyle\leq C\int_{\Omega}|\nabla u|^{2}\rho^{2}+2\int_{\Omega}|\nabla\nabla u|\cdot|\nabla u|\cdot|\nabla\rho|\cdot\rho
CΩ|u|2ρ2+12Ω|u|2ρ2+2Ω|u|2|ρ|2.\displaystyle\leq C\int_{\Omega}|\nabla u|^{2}\rho^{2}+\frac{1}{2}\int_{\Omega}|\nabla\nabla u|^{2}\rho^{2}+2\int_{\Omega}|\nabla u|^{2}|\nabla\rho|^{2}.

As a result,

(20) Ω|u|2ρ22CΩ|u|2ρ2+4Ω|u|2|ρ|2.\int_{\Omega}|\nabla\nabla u|^{2}\rho^{2}\leq 2C\int_{\Omega}|\nabla u|^{2}\rho^{2}+4\int_{\Omega}|\nabla u|^{2}|\nabla\rho|^{2}.

By the assumptions on Ω\Omega, we have

2CΩ|u|2ρ22CuL2Vol(Ω).2C\int_{\Omega}|\nabla u|^{2}\rho^{2}\leq 2C\|\nabla u\|_{L^{\infty}}^{2}\,\mathrm{Vol}(\Omega).

For every positive integer ii, let νi:[0,+)[0,1]\nu_{i}:[0,+\infty)\to[0,1] be the Lipschitz function such that ν(x)=1\nu(x)=1 for x2/ix\geq 2/i, ν(x)=0\nu(x)=0 for x1/ix\leq 1/i, and ν(x)=i\nu^{\prime}(x)=i for x[1/i,2/i]x\in[1/i,2/i]. Let χ\chi be a smooth function that is compactly support in Ω\Omega, such that 0χ10\leq\chi\leq 1, and χ=1\chi=1 on the given compact set KK. Define ρi\rho_{i} by

ρi(x)=[νidΩ(K𝒮(u),x)]χ,\rho_{i}(x)=[\nu_{i}\circ d_{\Omega}(K\cap\mathcal{S}(u),x)]\cdot\chi,

where dΩd_{\Omega} is the distance function on Ω\Omega. Theorem 5.9 implies that Ω|ρi|2\int_{\Omega}|\nabla\rho_{i}|^{2} is bounded as ii\to\infty. Hence by (20), the integral Ω|u|2ρi2\int_{\Omega}|\nabla\nabla u|^{2}\rho_{i}^{2} is bounded as ii\to\infty. Since ρi\rho_{i} is increasing with respect to ii and converges to χ\chi as ii\to\infty, this implies Ωχ|u|2\int_{\Omega}\chi\cdot|\nabla u|^{2} is finite, so the desired result is proved. ∎

5.3. From harmonic maps to /2\mathbb{Z}/2 harmonic 1-forms

Now assume MM is a closed oriented 33–manifold and let M~\widetilde{M} be its universal cover. Let π:M~M\pi:\widetilde{M}\to M be the covering map. Assume (𝒯,d𝒯)(\mathcal{T},d_{\mathcal{T}}) is an \mathbb{R}–tree with an isometric π1(M)\pi_{1}(M)–action, and assume u:M~𝒯u:\widetilde{M}\to\mathcal{T} is a π1(M)\pi_{1}(M)–equivariant harmonic map.

Then u\nabla u defines a two-valued 11–form on (u)M~\mathcal{R}(u)\subset\widetilde{M}. It is two-valued because the geodesics on 𝒯\mathcal{T} do not have canonical orientations. Locally, u\nabla u is given by ±v\pm v for some (single-valued) harmonic 11–form vv. Since uu is π1(M)\pi_{1}(M) equivariant, u\nabla u defines a two-valued 11–form on the image of (u)\mathcal{R}(u) in MM.

Theorem 5.15.

There exists a unique /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v} on MM, such that π(𝐯)=u\pi^{*}(\mathbf{v})=\nabla u on (u)\mathcal{R}(u).

Proof.

The uniqueness is clear since 𝒮(u)\mathcal{S}(u) has Hausdorff codimension 22. We prove the existence of such 𝐯\mathbf{v}. Recall that |u||\nabla u| is continuous (see Remark 5.13). Let Z~=|u|1(0)\widetilde{Z}=|\nabla u|^{-1}(0). Then 𝒮(u)Z~\mathcal{S}(u)\subset\widetilde{Z}, so u\nabla u defines a non-vanishing two-valued 11–form 𝐯~\tilde{\mathbf{v}} on M~Z~\widetilde{M}\setminus\widetilde{Z}. Since uu is equivariant, the set Z~\widetilde{Z} is invariant under the π1(M)\pi_{1}(M)–action. Let ZZ be the image of Z~\widetilde{Z} in MM. Then 𝐯~\tilde{\mathbf{v}} is the pull-back of a two valued 11–form 𝐯\mathbf{v} on MZM\setminus Z.

We show that 𝐯\mathbf{v} is a /2\mathbb{Z}/2 harmonic 11–form with respect to the zero locus ZZ. Since u\nabla u is harmonic in (u)\mathcal{R}(u), we know that 𝐯\mathbf{v} is locally given by non-vanishing harmonic 11–forms on MZM\setminus Z.

Note that if p(u)p\in\mathcal{R}(u) and |u(p)|=0|\nabla u(p)|=0, then ord(u;p)2\mathrm{ord}(u;p)\geq 2. As a result, by Theorem 5.6, there exists ϵ>0\epsilon>0 such that ord(u;p)1+ϵ\mathrm{ord}(u;p)\geq 1+\epsilon for all pZ~p\in\widetilde{Z}. By Corollary 5.11, we have

Br(p)|u|2Cr3+2ϵr022ord(u;p)\int_{B_{r}(p)}|\nabla u|^{2}\leq Cr^{3+2\epsilon}\cdot r_{0}^{2-2\mathrm{ord}(u;p)}

for all pZ~p\in\widetilde{Z} and r<r0/2r<r_{0}/2, where r0r_{0} is the minimum of 11 and the injectivity radius of MM at pp. Since ord(u;p)\mathrm{ord}(u;p) is upper semi-continuous and MM is closed, it is bounded from above on M~\widetilde{M}. Since MM is closed, the value of r0r_{0} has a positive lower bound. Therefore, the value of r022ord(u;p)r_{0}^{2-2\mathrm{ord}(u;p)} has an upper bound (which may depend on uu). This verifies Condition (iii) of Definition 3.1.

Condition (ii) of Definition 3.1 follows immediately from the fact that |u||\nabla u| is bounded and Proposition 5.14. ∎

Remark 5.16.

From the above construction, we also see that 𝒮(u)Z~\mathcal{S}(u)\subset\widetilde{Z}. In other words, every non-vanishing point of the /2\mathbb{Z}/2 harmonic 11–form corresponds to regular points of uu on M~\widetilde{M}.

Definition 5.17.

We call the /2\mathbb{Z}/2 harmonic form 𝐯\mathbf{v} obtained by Theorem 5.15 the /2\mathbb{Z}/2 harmonic 1-form associated with uu.

5.4. Maps between trees

Let MM, M~\widetilde{M}, (𝒯,d)(\mathcal{T},d), uu, be as in Section 5.3. Let 𝐯\mathbf{v} be the /2\mathbb{Z}/2 harmonic 11–form on MM associated with uu. Let 𝐯~\tilde{\mathbf{v}} be the pull back of 𝐯\mathbf{v} to M~\widetilde{M}. By Theorem 4.4, the leaf space (𝒯M~,𝐯~,dM~,𝐯~)(\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}},d_{\widetilde{M},\tilde{\mathbf{v}}}) is also an \mathbb{R}–tree. In this subsection, we study the relationship between 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}} and 𝒯\mathcal{T}.

Let μ𝐯~\mu_{\tilde{\mathbf{v}}} be the transverse measure defined from 𝐯~\tilde{\mathbf{v}} by (12). Let ZMZ\subset M, Z~M~\widetilde{Z}\subset\widetilde{M} denote the zero loci of 𝐯\mathbf{v} and 𝐯~\tilde{\mathbf{v}}.

Lemma 5.18.

Suppose γ:[0,1]M~\gamma:[0,1]\to\widetilde{M} is a 𝒞1\mathcal{C}^{1} arc. Let p=γ(0)p=\gamma(0), q=γ(1)q=\gamma(1). Then

(21) μ𝐯~(γ)d(u(p),u(q)).\mu_{\tilde{\mathbf{v}}}(\gamma)\geq d(u(p),u(q)).

Moreover, equality holds if γ\gamma is transverse to ker𝐯~\ker\tilde{\mathbf{v}}.

Proof.

We first prove (21). Both sides of (21) are continuous with respect to γ\gamma in the 𝒞1\mathcal{C}^{1} topology, so the inequality is a closed condition. Recall that (u)M~\mathcal{R}(u)\subset\widetilde{M} denotes the regular set of uu. Since (u)\mathcal{R}(u) has Hausdorff codimension at least 22, after perturbing γ\gamma if necessary, we may assume without loss of generality that the image of γ\gamma is contained in (u)\mathcal{R}(u).

Then we have

(22) ddtd(u(p),u(γ(t)))|u(γ(t)),γ˙(t)|=ddtμ𝐯~(γ|[0,t]).\frac{d}{dt}d\bigg{(}u(p),u\big{(}\gamma(t)\big{)}\bigg{)}\leq\big{|}\big{\langle}\nabla u(\gamma(t)),\dot{\gamma}(t)\big{\rangle}\big{|}\\ =\frac{d}{dt}\mu_{\tilde{\mathbf{v}}}(\gamma|_{[0,t]}).

Hence (21) is proved by integrating (22) for tt in [0,1][0,1].

If γ\gamma is transverse to ker𝐯~\ker\tilde{\mathbf{v}}, then the image of γ\gamma is contained in the complement of Z~\widetilde{Z}. By Remark 5.16, the image of γ\gamma is contained in (u)\mathcal{R}(u). Then for t[0,1]t\in[0,1], the value of uγ(t)𝒯u\circ\gamma(t)\in\mathcal{T} locally moves along geodesics at non-vanishing velocities. Since 𝒯\mathcal{T} is an \mathbb{R}–tree, this implies that the image of uγ(t)u\circ\gamma(t) for t[0,1]t\in[0,1] is a geodesic segment where the point uγ(t)u\circ\gamma(t) moves at non-vanishing velocities with respect to tt. Hence the inequality in (22) achieves equality for all t[0,1]t\in[0,1], and the desired result is proved. ∎

Let π𝐯~\pi_{\tilde{\mathbf{v}}} denote the quotient map from M~\widetilde{M} to 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}}.

Theorem 5.19.

There exists a unique continuous map f:𝒯M~,𝐯~𝒯f:\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}}\to\mathcal{T} such that u=fπ𝐯~u=f\circ\pi_{\tilde{\mathbf{v}}}. Moreover, we have the following properties:

  1. (1)

    ff is 11–Lipschitz.

  2. (2)

    The map π𝐯~\pi_{\tilde{\mathbf{v}}} is harmonic.

  3. (3)

    |π𝐯~|=|u||\nabla\pi_{\tilde{\mathbf{v}}}|=|\nabla u|.

  4. (4)

    If |u(p)|0|\nabla u(p)|\neq 0, then π𝐯~(p)=u(p)\nabla\pi_{\tilde{\mathbf{v}}}(p)=\nabla u(p) as two-valued harmonic 1-forms.

Proof.

By Lemma 5.18, we have

dM~,𝐯~(π𝐯~(p),π𝐯~(q))d(u(p),u(q))d_{\widetilde{M},\tilde{\mathbf{v}}}(\pi_{\tilde{\mathbf{v}}}(p),\pi_{\tilde{\mathbf{v}}}(q))\geq d(u(p),u(q))

for all p,qM~p,q\in\widetilde{M}. Hence the map uu factorizes uniquely as u=fπ𝐯~u=f\circ\pi_{\tilde{\mathbf{v}}} for a continuous map ff, and ff is 11–Lipschitz. As a result, we have |u||π𝐯~||\nabla u|\leq|\nabla\pi_{\tilde{\mathbf{v}}}|.

By the definition of dM~,𝐯~d_{\widetilde{M},\tilde{\mathbf{v}}}, the map π𝐯~\pi_{\tilde{\mathbf{v}}} is Lipschitz with Lipschitz constant 𝐯𝒞0\|\mathbf{v}\|_{\mathcal{C}^{0}}, so the map π𝐯~\pi_{\tilde{\mathbf{v}}} is in Wloc1,2W^{1,2}_{loc}.

For pM~Z~p\in\widetilde{M}\setminus\widetilde{Z}, let γ:[0,1]M~\gamma:[0,1]\to\widetilde{M} be an arc that is transverse to ker𝐯~\ker\tilde{\mathbf{v}}. Then

dM~,𝐯~(π𝐯~γ(0),π𝐯~γ(1))d(u(γ(0)),u(γ(1)))=μ𝐯~(γ)dM~,𝐯~(π𝐯~γ(0),π𝐯~γ(1)).d_{\widetilde{M},\tilde{\mathbf{v}}}\big{(}\pi_{\tilde{\mathbf{v}}}\circ\gamma(0),\pi_{\tilde{\mathbf{v}}}\circ\gamma(1)\big{)}\geq d(u(\gamma(0)),u(\gamma(1)))=\mu_{\tilde{\mathbf{v}}}(\gamma)\geq d_{\widetilde{M},\tilde{\mathbf{v}}}\big{(}\pi_{\tilde{\mathbf{v}}}\circ\gamma(0),\pi_{\tilde{\mathbf{v}}}\circ\gamma(1)\big{)}.

Here, the first inequality follows from the fact that ff is 11–Lipschitz, the second equality follows from Lemma 5.18, and the third inequality follows from the definition of dM~,𝐯~d_{\widetilde{M},\tilde{\mathbf{v}}}. Hence dM~,𝐯~(π𝐯~γ(0),π𝐯~γ(1))=d(u(γ(0)),u(γ(1)))d_{\widetilde{M},\tilde{\mathbf{v}}}\big{(}\pi_{\tilde{\mathbf{v}}}\circ\gamma(0),\pi_{\tilde{\mathbf{v}}}\circ\gamma(1)\big{)}=d(u(\gamma(0)),u(\gamma(1))). On the other hand, if γ\gamma is tangent to ker𝐯~\ker\tilde{\mathbf{v}} and is contained in the complement of Z~\widetilde{Z}, then both dM~,𝐯~(π𝐯~γ(0),π𝐯~γ(1))d_{\widetilde{M},\tilde{\mathbf{v}}}\big{(}\pi_{\tilde{\mathbf{v}}}\circ\gamma(0),\pi_{\tilde{\mathbf{v}}}\circ\gamma(1)\big{)} and d(u(γ(0),u(γ(1)))d(u(\gamma(0),u(\gamma(1))) are zero. In conclusion, if pp is in the complement of Z~\widetilde{Z}, then there exists an open ball Br(p)B_{r}(p) centered at pp such that

dM~,𝐯~(π𝐯~(p),π𝐯~(q))=d(u(p),u(q))d_{\widetilde{M},\tilde{\mathbf{v}}}\big{(}\pi_{\tilde{\mathbf{v}}}(p),\pi_{\tilde{\mathbf{v}}}(q)\big{)}=d(u(p),u(q))

for all qBr(p)q\in B_{r}(p). This implies that the directional derivatives and energy densities of π𝐯~\pi_{\tilde{\mathbf{v}}} and uu are the same on M~Z~\widetilde{M}\setminus\widetilde{Z}. Since Z~\widetilde{Z} has Lebesgue measure zero, they also define the same measure density functions on M~\widetilde{M}.

It remains to show that π𝐯~\pi_{\tilde{\mathbf{v}}} is harmonic. Let M0M~M_{0}\subset\widetilde{M} be a fundamental domain of the π1\pi_{1}–action. An equivariant map from M~\widetilde{M} to a tree is harmonic if and only if it minimizes the energy on M0M_{0} among all locally W1,2W^{1,2} equivariant maps. We show that π𝐯~\pi_{\tilde{\mathbf{v}}} minimizes the energy on M0M_{0}. Assume there exists equivariant map g:M~𝒯M~,𝐯~g:\widetilde{M}\to\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}} such that

M0|g|2<M0|π𝐯~|2,\int_{M_{0}}|\nabla g|^{2}<\int_{M_{0}}|\nabla\pi_{\tilde{\mathbf{v}}}|^{2},

then we have

M0|(fg)|2M0|g|2<M0|π𝐯~|2=M0|u|2,\int_{M_{0}}|\nabla(f\circ g)|^{2}\leq\int_{M_{0}}|\nabla g|^{2}<\int_{M_{0}}|\nabla\pi_{\tilde{\mathbf{v}}}|^{2}=\int_{M_{0}}|\nabla u|^{2},

where the first inequality follows from the fact that ff is 11–Lipschitz, and the last equation follows from the fact that |π𝐯~|=|u||\nabla\pi_{\tilde{\mathbf{v}}}|=|\nabla u| on M~\widetilde{M}. Since ff is π1(M)\pi_{1}(M)–equivariant, this contradicts the assumption that uu is energy minimizing. Hence the theorem is proved. ∎

Remark 5.20.

In general, if (M,𝐯)(M,\mathbf{v}) satisfies the conditions of Theorem 4.4, the quotient map from MM to the leaf tree 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} may not necessarily be harmonic. See Section 7.2 for a counterexample.

6. Proofs of the main theorems

In this section, we set up the relevant constructions and prove Theorem 1.1. We will also show that Theorem 1.2 follows directly from Theorem 5.19, once the map \mathcal{H} in (2) is constructed. Following our previous convention, let MM denote an oriented closed Riemannian 3-manifold, and let M~\widetilde{M} be its universal cover. Let Γ=π1(M)\Gamma=\pi_{1}(M).

6.1. Definition of the boundary map \mathcal{H}

Recall that 𝒫(Γ)\mathcal{PL}(\Gamma) denotes the space of projective length functions of all complete minimal Γ\Gamma–trees, and /2\mathcal{M}_{\mathbb{Z}/2} denotes the space of all /2\mathbb{Z}/2 harmonic 11–forms on MM with unit L2L^{2}–norm. In this subsection, we define a map

(23) :𝒫(Γ)/2.\mathcal{H}:\mathcal{PL}(\Gamma)\to\mathcal{M}_{\mathbb{Z}/2}.

By Theorem 2.1, if []𝒫(Γ)[\ell]\in\mathcal{PL}(\Gamma) is non-abelian, then there is a unique (up to equivariant isometry) minimal Γ\Gamma–tree 𝒯\mathcal{T} with length function \ell and no fixed ends. By Theorem 5.1 there is a π1(M)\pi_{1}(M) equivariant harmonic map u:M~𝒯u:\widetilde{M}\to\mathcal{T}. By Proposition 5.2, (see also Definition 5.17), the /2\mathbb{Z}/2 harmonic 11–form associated with uu is independent of the choice of \ell or uu. Let 𝐯\mathbf{v} be the /2\mathbb{Z}/2 harmonic 11–form associated with uu. Define

([])=𝐯/𝐯L2(M).\mathcal{H}([\ell])=\mathbf{v}/\|\mathbf{v}\|_{L^{2}(M)}.

Since 𝐯/𝐯L2(M)\mathbf{v}/\|\mathbf{v}\|_{L^{2}(M)} only depends on the projective class of \ell in 𝒫(Γ)\mathcal{PL}(\Gamma), the value of ([])\mathcal{H}([\ell]) is well-defined.

Next, we define the map \mathcal{H} on abelian length functions. Assume \ell is an abelian length function. By definition, there exists a homomorphism μ:Γ\mu:\Gamma\to\mathbb{R} so that (x)=|μ(x)|\ell(x)=|\mu(x)| for all xΓx\in\Gamma. It follows from straightforward algebra that if μ1,μ2\mu_{1},\mu_{2} are two homomorphisms Γ\Gamma\to\mathbb{R} such that |μ1(x)|=|μ2(x)||\mu_{1}(x)|=|\mu_{2}(x)| for all xΓx\in\Gamma, then μ1=±μ2\mu_{1}=\pm\mu_{2}. In other words, μ\mu is determined by \ell up to an overall sign.

Define vμv_{\mu} to be the (single-valued) harmonic 11–form on MM such that the periods of vμv_{\mu} on π1(M)\pi_{1}(M) are given by μ\mu. Let 𝐯\mathbf{v} be the /2\mathbb{Z}/2 harmonic 11–form given by ±vμ\pm v_{\mu}. Define

([])=𝐯/𝐯L2(M).\mathcal{H}([\ell])=\mathbf{v}/\|\mathbf{v}\|_{L^{2}(M)}.

In conclusion, we have constructed a well-defined map \mathcal{H} from 𝒫(Γ)\mathcal{PL}(\Gamma) to /2\mathcal{M}_{\mathbb{Z}/2}.

Lemma 6.1.

Suppose 𝒯\mathcal{T} is a minimal Γ\Gamma tree whose length function \ell is abelian and not identically zero, and u:M~𝒯u:\widetilde{M}\to\mathcal{T} is a Γ\Gamma–equivariant harmonic map. Then the tree 𝒯\mathcal{T} must be \mathbb{R}, and the Γ\Gamma action on 𝒯\mathcal{T} is given by translations.

Proof.

Suppose that Γ\Gamma has a minimal action on a tree 𝒯\mathcal{T} not isometric to \mathbb{R}, but with an abelian length function. By [CM87, p. 573], the action of Γ\Gamma must have a fixed end. Combining [DDW98, Thm. 5.3] and [Mes02, Thm. 1.2], we see that there can be no Γ\Gamma-equivariant harmonic map to 𝒯\mathcal{T}.

Now that we have proved 𝒯\mathcal{T}\cong\mathbb{R}, we show that the Γ\Gamma action must be given by translations. Assume xΓx\in\Gamma acts on 𝒯\mathcal{T}\cong\mathbb{R} by an orientation-reversing isometry, then (x)=0\ell(x)=0. As a result, for every yΓy\in\Gamma, we must have (xy)=(y)\ell(xy)=\ell(y). This is only possible when the action of yy equals the identity or the action of xx. Hence the length function of the Γ\Gamma–action is identically zero, contradicting the assumptions. ∎

Corollary 6.2.

Assume 𝒯\mathcal{T} is a complete minimal Γ\Gamma–tree and u:M~𝒯u:\widetilde{M}\to\mathcal{T} is a Γ\Gamma–equivariant harmonic map. Let \ell be the length function of 𝒯\mathcal{T}, and assume \ell is not identically zero. Then the /2\mathbb{Z}/2 harmonic 11–form associated with uu is equal to ([])\mathcal{H}([\ell]) up to a non-zero constant multiplication.

Proof.

If \ell is non-abelian, the statement follows from the definition of \mathcal{H}. If \ell is abelian, then Lemma 6.1 implies that 𝒯\mathcal{T}\cong\mathbb{R} and the action of Γ\Gamma on 𝒯\mathcal{T} are translations. Therefore, uu is a harmonic function and the associated /2\mathbb{Z}/2 harmonic 11–form is given by ±du\pm du. Since the periods of dudu coincide with the translation distances of the Γ\Gamma–action on 𝒯\mathcal{T}, the desired statement is proved. ∎

6.2. Comparison of the compactification limits

Recall that 𝒳(Γ)\mathcal{X}(\Gamma) denotes the character variety of Γ\Gamma, and that SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} denotes the moduli space of solutions to (1). Also recall that 𝒳(Γ)\mathcal{X}(\Gamma) and SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} are canonically identified by the Riemann–Hilbert correspondence. In this subsection, we prove the following proposition.

Proposition 6.3.

Assume {pi}\{p_{i}\} is a sequence of points in 𝒳(Γ)\mathcal{X}(\Gamma), and let {pi}\{p_{i}^{\prime}\} be the corresponding sequence in SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})}. Assume that pip_{i} converges in the Morgan–Shalen compactification to the projective length function [][\ell], and that pip_{i}^{\prime} converges in the Taubes compactification ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} to a /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v}. Then ([])=𝐯\mathcal{H}([\ell])=\mathbf{v}.

The proof of Proposition 6.3 amounts to careful book-keeping using results from the earlier sections. We start by reviewing the characterizations of the Morgan–Shalen compactification and the Taubes compactification.

6.2.1. Characterization of the Morgan–Shalen limit

By [DDW98], the Morgan–Shalen limit is given by the limit of harmonic maps from M~\widetilde{M} to 3\mathbb{H}^{3}. More precisely, 3\mathbb{H}^{3} denotes the three-dimensional hyperbolic space form. Its orientation-preserving isometry group is isomorphic to PSL2()\mathrm{PSL}_{2}(\mathbb{C}). Every piχ(Γ)p_{i}\in\chi(\Gamma) is given by a representation of Γ\Gamma in SL2()\mathrm{SL}_{2}(\mathbb{C}), which defines a Γ\Gamma–action on 3\mathbb{H}^{3}. Let uiu_{i} be a Γ\Gamma–equivariant harmonic map from M~\widetilde{M} to 3\mathbb{H}^{3} with respect to the above Γ\Gamma–action. Let E(ui)E(u_{i}) be the energy of uiu_{i} on a fundamental domain, and let did_{i} denote the pseudo-metric on M~\widetilde{M} defined by

(24) di(p,q)=d3(ui(p),ui(q))E(ui)1/2.d_{i}(p,q)=\frac{d_{\mathbb{H}^{3}}(u_{i}(p),u_{i}(q))}{E(u_{i})^{1/2}}.

If {pi}\{p_{i}\} converges to a projective length function [][\ell] on the boundary of the Morgan–Shalen compactification, then E(ui)E(u_{i})\to\infty, and did_{i} converges locally uniformly to a limit dd_{\infty}. The quotient metric space of M~\widetilde{M} with respect to dd_{\infty} is a minimal Γ\Gamma–tree 𝒯\mathcal{T}, the length function of 𝒯\mathcal{T} is in the projective class [][\ell], and the quotient map from M~\widetilde{M} to 𝒯\mathcal{T} is harmonic.

6.2.2. Characterization of the Taubes limit

If {pi}\{p_{i}^{\prime}\} converges to a /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v}, assume each pip_{i}^{\prime} is given by the pair (Ai,ϕi)(A_{i},\phi_{i}). Let PiP_{i} be the principal SU(2)\mathrm{SU}(2) bundle over MM where AiA_{i} is defined. Let 𝔤Pi\mathfrak{g}_{P_{i}} be the associated 𝔰𝔲(2)\mathfrak{su}(2) bundle with respect to the adjoint action.

Let ZZ be the zero locus of 𝐯\mathbf{v}. By the constructions in Section 3, on every compact set KK that is contained in an open ball in MZM\setminus Z, there exists ϕ\phi on KK, such that after a sequence of gauge transformations, ϕi/ϕiL2(M)\phi_{i}/\|\phi_{i}\|_{L^{2}(M)} converge to ϕ\phi in the weak W2,2W^{2,2} topology on KK. By Sobolev embedding theorems, ϕi/ϕiL2(M)\phi_{i}/\|\phi_{i}\|_{L^{2}(M)} converge to ϕ\phi in the C0C^{0} topology.

By part (3) of the second bullet point of [Tau13b, Thm. 1.1a], the limit spinor ϕ\phi locally has the form σv\sigma\otimes v, where vv is a harmonic 11–form, σ\sigma is a section of 𝔤Pi\mathfrak{g}_{P_{i}}, and |σ|=1|\sigma|=1 pointwise. In this case, the /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v} is locally equal to ±v\pm v.

6.2.3. Relationship between the two constructions

The relationship of the maps uiu_{i} and ϕi\phi_{i} is given as follows (see [DDW98, Sec. 2C]). Let P~i\widetilde{P}_{i} and 𝔤~Pi\tilde{\mathfrak{g}}_{P_{i}} denote the pull-backs of PiP_{i} and 𝔤Pi\mathfrak{g}_{P_{i}} to M~\widetilde{M}. The orthonormal frame bundle of 3\mathbb{H}^{3} is an SO(3)\mathrm{SO}(3) bundle. Its Lie algebra bundle is an 𝔰𝔬(3)\mathfrak{so}(3) bundle, and the pull-back of this bundle to M~\widetilde{M} is isomorphic to 𝔤~Pi\tilde{\mathfrak{g}}_{P_{i}}. Then there exists an equivariant isomorphism such that 12ui=ϕi-\frac{1}{2}\nabla u_{i}=\phi_{i} under the identification of Lie algebra bundles, where \nabla is given by the Levi-Civita connections of M~\widetilde{M} and \mathbb{H}. Different choices of equivariant isomorphisms of Lie algebra bundles correspond to gauge transformations on solutions to (1).

As a consequence, we have

(25) 12|ui(X)|=|ϕi(X)|\frac{1}{2}|\nabla u_{i}(X)|=|\phi_{i}(X)|

for every tangent vector XX of M~\widetilde{M}. Note that this equation is independent of the choice of the isomorphism between Lie algebra bundles, or gauge transformations.

6.2.4. Proof of Proposition 6.3

Let u:M~𝒯u:\widetilde{M}\to\mathcal{T} denote the quotient map from M~\widetilde{M} to 𝒯\mathcal{T}. Recall that 𝐯\mathbf{v} denotes the /2\mathbb{Z}/2 harmonic form in the Taubes limit. Let 𝐯~\tilde{\mathbf{v}} denote the pull-back of 𝐯\mathbf{v} to M~\widetilde{M}. Let 𝒰\mathcal{U} be the set of xM~x\in\widetilde{M} such that xx is not on the zero locus of 𝐯~\tilde{\mathbf{v}} or the zero set of |u||\nabla u|. Then 𝒰\mathcal{U} is a Γ\Gamma–equivariant open and dense subset of M~\widetilde{M}. The set 𝒰\mathcal{U} is contained in the regular set of uu, so u\nabla u is a non-vanishing /2\mathbb{Z}/2 harmonic 11–form on 𝒰\mathcal{U}.

As a result, ker𝐯\ker\mathbf{v} and keru\ker\nabla u are two smooth foliations on 𝒰\mathcal{U}. We first show that these two foliation are the same.

For each p𝒰p\in\mathcal{U}, let Br(p)B_{r}(p) be an open ball centered at pp such that

  1. (1)

    Br(p)𝒰B_{r}(p)\subset\mathcal{U},

  2. (2)

    the metrics did_{i} (defined in Equation (24)) converge to dd_{\infty} uniformly on Br(p)B_{r}(p),

  3. (3)

    ϕi\phi_{i} uniformly converge to a limit ϕ\phi on Br(p)B_{r}(p) after gauge transformations.

Let γ:[0,1]Br(p)\gamma:[0,1]\to B_{r}(p) be a smooth arc that is tangent to ker𝐯\ker\mathbf{v}. Write x=γ(0)x=\gamma(0), y=γ(1)y=\gamma(1). By (25), we have

d(x,y)\displaystyle d_{\infty}(x,y) =limidi(x,y)lim supi01|ui(γ˙(t))|𝑑t\displaystyle=\lim_{i\to\infty}d_{i}(x,y)\leq\limsup_{i\to\infty}\int_{0}^{1}|\nabla u_{i}(\dot{\gamma}(t))|\,dt
=2lim supi01|ϕi,γ˙(t)|𝑑t\displaystyle=2\limsup_{i\to\infty}\int_{0}^{1}|\langle\phi_{i},\dot{\gamma}(t)\rangle|\,dt
=201|ϕ,γ˙(t)|𝑑t.\displaystyle=2\int_{0}^{1}|\langle\phi,\dot{\gamma}(t)\rangle|\,dt.

Recall that ϕ\phi is locally given by σv\sigma\otimes v, where vv is a harmonic 11–form such that locally 𝐯=±v\mathbf{v}=\pm v. Since γ˙\dot{\gamma} is tangent to ker𝐯\ker\mathbf{v}, we have |ϕ,γ˙(t)|=0|\langle\phi,\dot{\gamma}(t)\rangle|=0 for all t[0,1]t\in[0,1]. This implies d(x,y)=0d_{\infty}(x,y)=0. As a result, every leaf of ker𝐯\ker\mathbf{v} is contained in a leaf of keru\ker\nabla u as foliations on 𝒰\mathcal{U}. Therefore, ker𝐯=keru\ker\mathbf{v}=\ker\nabla u on 𝒰\mathcal{U}.

Next, we show that u\nabla u and 𝐯\mathbf{v} differ only by a locally constant multiplication on 𝒰\mathcal{U}. Let Br(p)B_{r}(p) be as above. Since Br(p)B_{r}(p) is simply connected, both u\nabla u and 𝐯\mathbf{v} can be lifted to single-valued harmonic 11–forms. Locally, write 𝐯\mathbf{v} as ±v\pm v and u\nabla u as ±w\pm w. Both vv and ww are non-vanishing. Since ker𝐯=keru\ker\mathbf{v}=\ker\nabla u, there exists a non-zero function gg such that v=gwv=g\cdot w. Then we have

0=dv=d(gw)=dgw+gdw=dgw,0=dv=d(gw)=dg\wedge w+gdw=dg\wedge w,
0=d(v)=d(gw)=dg(w)+gd(w)=dg(w).0=d(*v)=d(*gw)=dg\wedge(*w)+gd(*w)=dg\wedge(*w).

As a result, dgw=0dg\wedge w=0, dg(w)=0dg\wedge(*w)=0, so dg=0dg=0, thus gg is a constant function. In conclusion, u\nabla u and 𝐯\mathbf{v} differ only by a locally constant multiplication on 𝒰\mathcal{U}.

Since the zero loci of 𝐯\mathbf{v} and |u||\nabla u| have Hausdorff codimension 22, the above result implies that 𝐯\mathbf{v} is the /2\mathbb{Z}/2 harmonic 11–form associated with uu up to a constant multiplication. By Corollary 6.2, we conclude that 𝐯=([])\mathbf{v}=\mathcal{H}([\ell]). ∎

6.3. Proofs of Theorems 1.1 and 1.2

Define

Ξ¯:𝒳(Γ)¯¯SL2()\overline{\Xi}:\overline{\mathcal{X}(\Gamma)}\to\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}

such that Ξ¯\overline{\Xi} is given by the (inverse of the) Riemann–Hilbert map on 𝒳(Γ)\mathcal{X}(\Gamma) and by \mathcal{H} on 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)}.

Lemma 6.4.

Both 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} and ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} are metrizable.

Proof.

By Corollary 3.7, the space ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} is compact and Hausdorff, and hence it is regular. The space /2\mathcal{M}_{\mathbb{Z}/2} defined in Section 3.1 is separable with respect to the 𝒞0\mathcal{C}^{0}–topology, because it is a closed subset of the space of continuous 22–valued sections of TMT^{*}M, which is a separable metric space. Hence the topology on 𝕄\mathbb{M} defined in Section 3.2 is second countable. By the Urysohn metrization theorem, we conclude that ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} is metrizable.

Similarly, by Lemma 2.2, the space 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is compact and Hausdorff. Since CC is countable, the space C\mathbb{P}^{C} is second countable. This implies 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is second countable. Hence 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} is metrizable. ∎

The following statement is the first part of Theorem 1.1:

Theorem 6.5.

The map Ξ¯\overline{\Xi} is continuous.

Proof.

The result follows from Proposition 6.3 and a formal argument. Since both 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} and ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} are metrizable, we only need to show that if {pi}\{p_{i}\} is a sequence that converges to pp in 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)}, then Ξ¯(pi)\overline{\Xi}(p_{i}) converges to Ξ¯(p)\overline{\Xi}(p). Since 𝒳(Γ)\mathcal{X}(\Gamma) is an open subset of 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} and the map Ξ¯\overline{\Xi} is already known to be continuous on 𝒳(Γ)\mathcal{X}(\Gamma), we only need to prove the statement when p𝒳(Γ)¯p\in\partial\overline{\mathcal{X}(\Gamma)}. We only need to consider two cases: (1) all pip_{i}’s are in 𝒳(Γ)\mathcal{X}(\Gamma); (2) all pip_{i}’s are in 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)}.

If all pip_{i}’s are in 𝒳(Γ)\mathcal{X}(\Gamma), then by the compactness of ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}, every subsequence of Ξ¯(pi)\overline{\Xi}(p_{i}) has a convergent subsequence. By Proposition 6.3, this subsequence must converge to Ξ¯(p)\overline{\Xi}(p). Hence Ξ¯(pi)\overline{\Xi}(p_{i}) converge to Ξ¯(p)\overline{\Xi}(p).

If all pip_{i}’s are in 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)}, we use an argument by contradiction. Let d1d_{1} be a metric for the topology of 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)} and let d2d_{2} be a metric for ¯SL2()\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}. Assume there exists ϵ>0\epsilon>0 such that there is a subsequence of {pi}\{p_{i}\}, which we will still denote by {pi}\{p_{i}\}, such that d2(Ξ¯(pi),Ξ¯(p))ϵd_{2}(\overline{\Xi}(p_{i}),\overline{\Xi}(p))\geq\epsilon for all ii. For each ii, since pi𝒳(Γ)¯p_{i}\in\partial\overline{\mathcal{X}(\Gamma)}, there exists a sequence {qi,j}j1\{q_{i,j}\}_{j\geq 1} in 𝒳(Γ)\mathcal{X}(\Gamma), such that {qi,j}\{q_{i,j}\} converges to pip_{i} as jj\to\infty. By Proposition 6.3, we have

limjΞ¯(qi,j)=Ξ¯(pi).\lim_{j\to\infty}\overline{\Xi}(q_{i,j})=\overline{\Xi}(p_{i}).

As a result, we may find a pip_{i}^{\prime} in the sequence {qi,j}j1\{q_{i,j}\}_{j\geq 1} such that pi𝒳(Γ)p_{i}^{\prime}\in\mathcal{X}(\Gamma), d1(pi,pi)<1/id_{1}(p_{i},p_{i}^{\prime})<1/i, and d2(Ξ¯(pi),Ξ¯(pi))<ϵ/2d_{2}(\overline{\Xi}(p_{i}^{\prime}),\overline{\Xi}(p_{i}))<\epsilon/2. The sequence {pi}i1\{p_{i}^{\prime}\}_{i\geq 1} then satisfies limipi=p\lim_{i\to\infty}p_{i}^{\prime}=p and

d2(Ξ¯(pi),Ξ¯(p))ϵ/2d_{2}(\overline{\Xi}(p_{i}^{\prime}),\overline{\Xi}(p))\geq\epsilon/2

for all ii. This contradicts Proposition 6.3. ∎

Now we prove the second part of Theorem 1.1:

Theorem 6.6.

The map Ξ¯\overline{\Xi} is surjective.

Proof.

We need to show that every point q¯SL2()q\in\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} is in the image of Ξ¯\overline{\Xi}. Let qiq_{i} be a sequence of points in SL2()\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} such that qiqq_{i}\to q. Let pip_{i} be the preimage of qiq_{i} in 𝒳(Γ)\mathcal{X}(\Gamma). By the compactness of 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)}, there is a subsequence of pip_{i}, which we still denote by {pi}i1\{p_{i}\}_{i\geq 1}, such that pip_{i} converges to a point pp in 𝒳(Γ)¯\overline{\mathcal{X}(\Gamma)}. Since Ξ¯\overline{\Xi} is continuous, we have Ξ¯(p)=q\overline{\Xi}(p)=q. ∎

Proof of Theorem 1.2.

By Corollary 6.2, the /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v} is associated with the harmonic map uu up to a non-zero constant multiplication. So Theorem 1.2 follows from Theorem 5.19. ∎

6.4. Non-injectivity of the map Ξ¯\overline{\Xi}

In general, the map Ξ¯\overline{\Xi} may not be injective. This is certainly well-known, though we have been unable to find an exact reference. For completeness we provide the details of an example. In the following, we give a counterexample for the 22–dimensional analogue. Multiplying the spaces with S1S^{1} will yield a counterexample for the 33–dimensional case. Specifically, we construct distinct length functions appearing in 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)} with the same image under =Ξ¯\mathcal{H}=\partial\overline{\Xi}.

Let Σg\Sigma_{g} be a Riemann surface of genus g3g\geq 3. Express the surface as a connect sum: Σg=Σ1Σg1\Sigma_{g}=\Sigma_{1}\sharp\Sigma_{g-1} of surfaces of genus 11 and g1g-1, respectively. This gives an amalgamated product expression for the fundamental group: Γ=π1(Σg)=𝔽2𝔽2g4\Gamma=\pi_{1}(\Sigma_{g})=\mathbb{F}_{2}*_{\mathbb{Z}}\mathbb{F}_{2g-4}, where 𝔽k\mathbb{F}_{k} denotes the free group on kk generators.

Let ρ^i:π1(Σg1)SL2()\hat{\rho}_{i}:\pi_{1}(\Sigma_{g-1})\to\mathrm{SL}_{2}(\mathbb{C}) be a divergent family of completely reducible representations converging in the Morgan–Shalen limit to a nonabelian projective length function ^\hat{\ell}. This is possible because g12g-1\geq 2, so we may take ρ^i\hat{\rho}_{i} to be discrete and faithful, for example.

Now by the product expression, the ρ^i\hat{\rho}_{i} extend to representations ρi\rho_{i} of Γ\Gamma, and they clearly converge to a Morgan–Shalen limit \ell, which is the extension of ^\hat{\ell}. Let Σ~g\widetilde{\Sigma}_{g} be the universal cover of Σg\Sigma_{g}, and let u:Σ~g𝒯u:\widetilde{\Sigma}_{g}\to\mathcal{T}_{\ell} be an equivariant harmonic map with associated /2\mathbb{Z}/2 harmonic 1-form 𝐯\mathbf{v}. We denote by 𝒯𝐯\mathcal{T}_{\mathbf{v}} the associated leaf tree defined by 𝐯\mathbf{v}. On the other hand, each ρi\rho_{i} contains the free group 𝔽2\mathbb{F}_{2} in its kernel, and therefore the edge stabilizers of 𝒯\mathcal{T}_{\ell} all contain free groups. By Skora’s Theorem [Sko96], 𝒯\mathcal{T}_{\ell} cannot be dual to a measured foliation on Σg\Sigma_{g}. It follows that 𝒯𝐯\mathcal{T}_{\mathbf{v}} and 𝒯\mathcal{T}_{\ell} will necessarily have different length functions.

Finally, the length function 𝐯\ell_{\mathbf{v}} of 𝒯𝐯\mathcal{T}_{\mathbf{v}} also appears in the boundary. Indeed, as in Example 3.2, s:=𝐯𝐯H0(KΣg2)s:=\mathbf{v}\otimes\mathbf{v}\in H^{0}(K^{2}_{\Sigma_{g}}) is a holomorphic quadratic differential on Σg\Sigma_{g}. The length function 𝐯\ell_{\mathbf{v}} is associated to the dual tree of the vertical measured foliation of ss, and this appears in the Thurston boundary of (P)SL2(){\rm(P)SL}_{2}(\mathbb{R}) representations of Γ\Gamma.

Therefore, we have constructed two length functions [][𝐯][\ell]\neq[\ell_{\mathbf{v}}] such that ([])=([𝐯])\mathcal{H}([\ell])=\mathcal{H}([\ell_{\mathbf{v}}]), which implies Ξ¯\overline{\Xi} is not injective.

7. Applications

In this section, we give several applications of Theorems 1.1 and 1.2 and discuss some related results.

Section 7.1 proves Theorem 1.3, which is an existence result for singular /2\mathbb{Z}/2 harmonic forms on a large class of rational homology spheres. In Section 7.2, we show that the projection map from a simply-connected manifold to the leaf space of a /2\mathbb{Z}/2 harmonic 11–form may not be harmonic. In Section 7.3, we discuss a folklore conjecture on the non-existence of /2\mathbb{Z}/2 harmonic 11–forms on S3S^{3}, and we prove a result about /2\mathbb{Z}/2 harmonic forms on simply connected closed manifolds. In Section 7.4, we prove a 𝒞\mathcal{C}^{\infty} convergence result for Korevaar–Schoen limits using results from gauge theory. In Section 7.5, we discuss the relationship of Morgan–Shalen limits, singular measured foliations, and the Hubbard–Masur construction.

7.1. Existence of singular /2\mathbb{Z}/2 harmonic 1-forms

We prove Theorem 1.3 using Culler–Shalen’s construction of dual trees associated with essential surfaces.

Let MM be a closed oriented 3-manifold with Γ:=π1(M)\Gamma:=\pi_{1}(M). Assume SMS\subset M is an embedded two-sided surface. We briefly review the construction of the dual tree of SS as in [Sha02, Sec. 1.4]. Let p:M~Mp:\widetilde{M}\to M be the universal cover and S~:=p1(S)\widetilde{S}:=p^{-1}(S) the preimage of SS. There is a canonical 1-dimensional simplicial complex 𝒯S\mathcal{T}_{S} associated with SS, defined as follows: the vertices are the connected components of M~S~\widetilde{M}\setminus\widetilde{S}, and the edges are the connected components of S~\widetilde{S}. An edge is incidental to a vertex if the corresponding component of S~\widetilde{S} is contained in the boundary of the corresponding component of M~S~\widetilde{M}\setminus\widetilde{S}. The space 𝒯S\mathcal{T}_{S} is a simplicial tree [Sha02, Sec. 1.4], and the fundamental group Γ\Gamma acts on 𝒯S\mathcal{T}_{S} by simplicial homeomorphisms. In the following lemma, we show that under the assumptions in Theorem 1.3, the action of Γ\Gamma on 𝒯S\mathcal{T}_{S} has no fixed points (see also [Sha02, Proposition 1.5.2]).

Lemma 7.1.

Suppose MM is a rational homology sphere. Suppose SS is a closed connected embedded surface in MM that is two-sided, π1\pi_{1}–injective, and does not bound an embedded ball in MM. Then the action of Γ\Gamma on 𝒯S\mathcal{T}_{S} has no global fixed points.

Proof.

Since MM is a rational sphere and SS is two-sided, MSM\setminus S must be disconnected. In fact, if MSM\setminus S is connected, then there is a circle in MM that intersects SS transversely at one point, which implies that the homology class of SS must be non-torsion, contradicting the assumptions. Since SS is connected, MSM\setminus S has two connected components.

The action of Γ\Gamma on 𝒯S\mathcal{T}_{S} has a globally fixed vertex if and only if there is a component CC of MSM\setminus S such that π1(C)π1(M)\pi_{1}(C)\to\pi_{1}(M) is surjective. By the Seifert–van Kampen theorem, this is equivalent to the surjectivity of π1(S)π1(MC)\pi_{1}(S)\to\pi_{1}(M\setminus C). Since SS is π1\pi_{1}–injective, this condition implies π1(S)π1(MC)\pi_{1}(S)\to\pi_{1}(M\setminus C) is an isomorphism. If SS is a sphere, then MCM\setminus C must be a 33–ball, which contradicts the assumptions. If the genus of SS is positive, by a standard result in 3-manifold topology (see, for example, [Hat07, Lemma 3.5]), the rank of the kernel of H1(S)H1(MC)H_{1}(S)\to H_{1}(M\setminus C) equals the genus of SS, so π1(S)π1(MC)\pi_{1}(S)\to\pi_{1}(M\setminus C) cannot be isomorphic.

The action of Γ\Gamma on 𝒯S\mathcal{T}_{S} has a globally fixed edge if and only if π1(S)π1(M)\pi_{1}(S)\to\pi_{1}(M) is surjective. By the Seifert–van Kampen theorem and the assumption that SS is π1\pi_{1}–injective, this implies π1(S)π1(MC)\pi_{1}(S)\to\pi_{1}(M\setminus C) is an isomorphism for each component CC of MSM\setminus S. Hence we get a contradiction by the same argument as before. ∎

Proof of Theorem 1.3.

Let Γ=π1(M)\Gamma=\pi_{1}(M). Since 𝒯S\mathcal{T}_{S} is a simplicial tree, it is complete as a metric space. By Lemma 7.1, the set of projective length functions 𝒫(Γ)\mathcal{PL}(\Gamma) is non-empty. Hence the construction of the map :𝒫(Γ)/2\mathcal{H}:\mathcal{PL}(\Gamma)\to\mathcal{M}_{\mathbb{Z}/2} in Section 6.1 implies that /2\mathcal{M}_{\mathbb{Z}/2} is non-empty. Since MM is a rational homology sphere, every non-zero /2\mathbb{Z}/2 harmonic form on MM must be non-trivial. ∎

We remark that essential surfaces also played an important role in the recent construction of new Casson-type invariants by Dunfield and Rasmussen [DR24]; see also [DGR22] for results on the counting of essential surfaces.

7.2. The quotient map to the leaf space

Let MM be a closed manifold, 𝐯\mathbf{v} be a /2\mathbb{Z}/2 harmonic 1-form over MM, let π:M~M\pi:\widetilde{M}\to M be the universal covering map with 𝐯~:=π𝐯\tilde{\mathbf{v}}:=\pi^{*}\mathbf{v}, then by Theorem 4.4, the leaf space (𝒯M~,𝐯~,dM~,𝐯~)(\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}},d_{\widetilde{M},\tilde{\mathbf{v}}}) is an \mathbb{R}–tree. When MM is 2-dimensional, it is proved in [Wol95, Prop. 3.1] that the quotient map from M~\widetilde{M} to 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}} is harmonic. It turns out that this statement does not hold generally in higher dimensions. The following is a counterexample when MM is four dimensional.

Example 7.2.

In [BDO11, Sec. 3, example before Thm. 3.2], an example of a compact, simply-connected, smooth projective variety MM is constructed, such that there is a non-trivial holomorphic section ss of ΩM1ΩM1\Omega^{1}_{M}\otimes\Omega^{1}_{M} that locally has the form fμ2f\mu^{2} with ff a holomorphic function and μ\mu a holomorphic 11–form. Therefore, the real part of the square-root of ss defines a /2\mathbb{Z}/2 harmonic form on MM, which we denote by 𝐯\mathbf{v}. The universal cover of MM is the same as MM. Let 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} be the leaf space of 𝐯\mathbf{v} on MM as given by Theorem 4.4.

One can choose the parameters in the construction so that the leaf space of ker𝐯\ker\mathbf{v} contains infinitely many points. To explain this, we need to recall the construction of MM from [BDO11]. Let A3A^{3} be a 33–dimensional abelian variety and let M^\hat{M} be a smooth hypersurface in A3A^{3}. Let θ=id\theta=-\mathrm{id} be the natural involution on A3A^{3}, and assume M^\hat{M} passes through exactly one fixed point pp of θ\theta. Then MM is defined to be the minimal resolution of M^/θ\hat{M}/\theta. Up to multiplication by constants, there is a unique holomorphic 11–form ww on A3A^{3} such that w|Tp(M^)=0w|_{T_{p}(\hat{M})}=0. The section ss is defined as an extension of the push-forward of w2w^{2} to (M^{p})/θ(\hat{M}\setminus\{p\})/\theta.

As a result, the /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v} is equal to the push-forward of the real part of ±w\pm w on (M^{p})/θ(\hat{M}\setminus\{p\})/\theta. Identifying A3A^{3} with a quotient of 3\mathbb{C}^{3} by a discrete group of translations, one may choose MM so that the space TpMT_{p}M is given by a linear equation with coefficients in [1]\mathbb{Q}[\sqrt{-1}]. In this case, the kernel of the real part of ω\omega defines a foliation on A3A^{3} such that every leaf is closed. If L1,L2A3L_{1},L_{2}\subset A^{3} are two closed leaves whose images in A3/θA^{3}/\theta are disjoint, are both disjoint from pp, and both intersect M^\hat{M} non-trivially, then the images of LiM^L_{i}\cap\hat{M} in 𝒯M,𝐯\mathcal{T}_{{M},\mathbf{v}} are two distinct points. As a result, 𝒯M,𝐯\mathcal{T}_{{M},\mathbf{v}} contains infinitely many distinct points.

Since M{M} is compact, by the maximum principle for harmonic maps, every harmonic map from M{M} to an \mathbb{R}–tree must be constant. Therefore, the quotient map from M{M} to 𝒯M,𝐯\mathcal{T}_{{M},\mathbf{v}} cannot be harmonic. ∎

7.3. /2\mathbb{Z}/2 harmonic forms on closed simply connected manifolds

In this subsection, we prove Theorem 1.7, which states that /2\mathbb{Z}/2 harmonic 11–forms cannot exist on (S3,g)(S^{3},g) under certain additional conditions. In fact, we prove a more general result that holds for closed simply connected manifolds in all dimensions, as stated in Theorem 7.4 below.

Suppose MM is a simply connected closed Riemannian manifold (not necessarily in dimension 33), and 𝐯\mathbf{v} is a /2\mathbb{Z}/2 harmonic 11–form on MM with zero locus ZZ. Let μ𝐯\mu_{\mathbf{v}} be the transverse measure on MZM\setminus Z defined by 𝐯\mathbf{v} (see (12)), and let dM,𝐯d_{M,\mathbf{v}} be the corresponding pseudo-metric on MM (see (13)); we also use dM,𝐯d_{M,\mathbf{v}} to denote the induced metric on the leaf space 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}}.

Definition 7.3.

We say that a zero point pZp\in Z of 𝐯\mathbf{v} is cylindrical, if there exists a local coordinate chart (x1,,xn)(x_{1},\dots,x_{n}) on a neighborhood of pp such that 𝐯=φRe(dzk)\mathbf{v}=\varphi\cdot Re(d\sqrt{z^{k}}) for some integer k3k\geq 3, where φ\varphi is a smooth and non-vanishing function, and z=x1+x21z=x_{1}+x_{2}\sqrt{-1}\in\mathbb{C}.

The following theorem is a generalization of Theorem 1.7.

Theorem 7.4.

Suppose MM is a closed, simply-connected, Riemannian manifold. Then there is no /2\mathbb{Z}/2 harmonic 11–form 𝐯\mathbf{v} on MM such that the following conditions hold at the same time.

  1. (1)

    Every point pp of 𝐯\mathbf{v} is cylindrical.

  2. (2)

    For every arc γ\gamma in MZM\setminus Z transverse to ker𝐯\ker\mathbf{v}, where ZZ denotes the zero locus of 𝐯\mathbf{v}, we have

    μ𝐯(γ)=dM,𝐯(x,y).\mu_{\mathbf{v}}(\gamma)=d_{M,\mathbf{v}}(x,y).
Remark 7.5.

Condition (1) implies that the hypotheses of Theorem 4.4 hold. By Lemma 5.18, Condition (2) always holds if the /2\mathbb{Z}/2 harmonic form is associated with an equivariant harmonic map from a universal cover of a closed 33–manifold to a tree.

Theorem 7.4 will be a straightforward consequence of the following lemma.

Lemma 7.6.

Assume MM is a simply connected (but not necessarily closed) manifold and 𝐯\mathbf{v} satisfies the conditions in Theorem 7.4. Let π𝐯\pi_{\mathbf{v}} be the projection map from MM to the leaf tree 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}}. Then germs of convex functions on 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} pull back to germs of subharmonic functions on MM.

Here, a function ff on an \mathbb{R}–tree 𝒯\mathcal{T} is called convex, if for every geodesic γ:[a,b]𝒯\gamma:[a,b]\to\mathcal{T} parametrized by arc length and every λ[0,1]\lambda\in[0,1], we have

λfγ(a)+(1λ)fγ(b)fγ(λa+(1λ)b).\lambda f\circ\gamma(a)+(1-\lambda)f\circ\gamma(b)\geq f\circ\gamma(\lambda a+(1-\lambda)b).
Remark 7.7.

The condition that germs of convex functions pull back to germs of subharmonic functions is a local condition that can be verified near every point. It is also well-known that for maps between manifolds, this condition is equivalent to the map being harmonic (see, for example, [Ish79, Thm. 3.4]).

Proof.

By Condition (2) in Theorem 7.4, the map π𝐯\pi_{\mathbf{v}} is regular and harmonic near every non-zero point of 𝐯\mathbf{v}, so it satisfies the desired properties on the complement of the zero locus of 𝐯\mathbf{v}.

Let pp be a zero point of 𝐯\mathbf{v}. Take a coordinate chart (x1,x2,x3,,xn)(x_{1},x_{2},x_{3},\dots,x_{n}) containing pp such that Definition 7.3 holds. Let kk be the integer in Condition (1). Let UϵU_{\epsilon} be the open neighborhood of pp given by |xi|<ϵ|x_{i}|<\epsilon (i=1,,n)(i=1,\dots,n) in this chart.

Let 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} denote the leaf space of 𝐯\mathbf{v} on UϵU_{\epsilon}, then 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} is given by the union of kk segments with one end point identified. The identified end point is the image of pp, which we denote by p^\hat{p}. In the following, we will call each segment in 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} a branch. The kk branches admit a natural cyclic order.

Let 𝒯ϵ\mathcal{T}_{\epsilon} denote the image of 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} in 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}}. We will abuse notation and use p^\hat{p} to also denote the image of pp in 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}}.

By Condition (2), for ϵ\epsilon sufficiently small, the map of each branch of 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} into 𝒯ϵ\mathcal{T}_{\epsilon} is an isometric embedding. Therefore, the map from 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} to 𝒯ϵ\mathcal{T}_{\epsilon} is a quotient map that identifies pairs or groups of branches. Moreover, Condition (2) also implies that neighboring branches of 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} cannot be identified in 𝒯ϵ\mathcal{T}_{\epsilon}.

Let ff be a convex function defined on a neighborhood of the closure of 𝒯ϵ\mathcal{T}_{\epsilon}, we show that ff pulls back to a subharmonic function near pp.

We first consider the case that f(x)=dist𝒯ϵ(x,p^)f(x)=-\mathrm{dist}_{\mathcal{T}_{\epsilon}}(x,\hat{p}) on one of the branches of 𝒯ϵ\mathcal{T}_{\epsilon} (which we denote by II), and f(x)=dist𝒯ϵ(x,p^)f(x)=\mathrm{dist}_{\mathcal{T}_{\epsilon}}(x,\hat{p}) on 𝒯ϵI\mathcal{T}_{\epsilon}\setminus I.

Let f~\tilde{f} be the pull-back of ff to UϵU_{\epsilon}. Then the pre-image of p^𝒯Uϵ,𝐯\hat{p}\in\mathcal{T}_{U_{\epsilon},\mathbf{v}} in MM cuts UϵU_{\epsilon} into kk domains, which we denote by Ω1,,Ωk\Omega_{1},\dots,\Omega_{k} in cyclic order. Then there exists a set S{1,,k}S\subset\{1,\dots,k\}, such that

(26) f~(x)={dUϵ,𝐯(x,p) if xΩi,iSdUϵ,𝐯(x,p) if xΩi,iS.\tilde{f}(x)=\begin{cases}-d_{U_{\epsilon},\mathbf{v}}(x,p)\quad\text{ if }x\in\Omega_{i},\,\,i\in S\\ d_{U_{\epsilon},\mathbf{v}}(x,p)\quad\text{ if }x\in\Omega_{i},\,\,i\notin S.\end{cases}

In the following, the subscripts will always be interpreted modulo kk. Since neighboring branches of 𝒯Uϵ,𝐯\mathcal{T}_{U_{\epsilon},\mathbf{v}} have distinct images in 𝒯ϵ\mathcal{T}_{\epsilon}, we have {i,i+1}S\{i,i+1\}\not\subseteq S for every ii.

By [Ish95], we only need to show that

(27) Uϵf~,n0,\int_{\partial U_{\epsilon}}\langle\nabla\tilde{f},n\rangle\geq 0,

for all ϵ\epsilon, where nn denotes the outward unit normal vector field on Uϵ\partial U_{\epsilon}. For each Ωi\Omega_{i}, we decompose Ωi\partial\Omega_{i} into Ωi=FiFiFi′′\partial\Omega_{i}=F_{i}\cup F_{i}^{\prime}\cup F_{i}^{\prime\prime}, where Fi=ΩiUϵF_{i}=\partial\Omega_{i}\cap\partial U_{\epsilon}, Fi=ΩiΩi1F_{i}^{\prime}=\Omega_{i}\cap\Omega_{i-1}, Fi′′=ΩiΩi+1F_{i}^{\prime\prime}=\Omega_{i}\cap\Omega_{i+1}. It is clear that Fi=Fi1′′F_{i}^{\prime}=F_{i-1}^{\prime\prime} for all ii.

We have

Ωf~,n=i=1nFif~,n.\int_{\partial\Omega}\langle\nabla\tilde{f},n\rangle=\sum_{i=1}^{n}\int_{F_{i}}\langle\nabla\tilde{f},n\rangle.

Define

Ei=Fi|𝐯,n|,E_{i}=\int_{F_{i}^{\prime}}|\langle\mathbf{v},n\rangle|,

where nn is a unit normal vector field of FiF_{i}^{\prime}. By (26) and the fact that dUϵ,𝐯(x,p)d_{U_{\epsilon},\mathbf{v}}(x,p) is a harmonic function on xx in the interior of each Ωi\Omega_{i}, we have

Fif~,n={EiEi+1 if iS,Ei+Ei+1 if iS.\int_{F_{i}}\langle\nabla\tilde{f},n\rangle=\begin{cases}-E_{i}-E_{i+1}\quad\text{ if }i\in S,\\ E_{i}+E_{i+1}\quad\text{ if }i\notin S.\\ \end{cases}

Hence (27) follows from the fact that {i,i+1}S\{i,i+1\}\not\subseteq S for all ii.

This proves that the pull-back of ff is subharmonic near pp when ff has the special form given as above. In general, assume gg is an arbitrary convex function that is defined on a neighborhood of the closure of 𝒯ϵ\mathcal{T}_{\epsilon}. Then there exist constants c10c_{1}\geq 0, c2c_{2}, and a function ff having the form above, such that gc1f+c2g\geq c_{1}f+c_{2} on 𝒯ϵ\mathcal{T}_{\epsilon} and g(p^)=c1f(p^)+c2g(\hat{p})=c_{1}f(\hat{p})+c_{2}. Let g~\tilde{g} denote the pull back of gg to UϵU_{\epsilon}. Let hgh_{g} denote the harmonic function on UϵU_{\epsilon} such that hg=g~h_{g}=\tilde{g} on Uϵ\partial U_{\epsilon}, let hfh_{f} denote the harmonic function on UϵU_{\epsilon} such that hf=f~h_{f}=\tilde{f} on Uϵ\partial U_{\epsilon}. Then we have

g~(p)=c1f~(p)+c2c1hf(p)+c2hg(p).\tilde{g}(p)=c_{1}\tilde{f}(p)+c_{2}\leq c_{1}h_{f}(p)+c_{2}\leq h_{g}(p).

Since the above inequality holds for all p|𝐯|1(0)p\in|\mathbf{v}|^{-1}(0) and ϵ\epsilon, we conclude that g~\tilde{g} is subharmonic, and the lemma is proved. ∎

Proof of Theorem 7.4.

Assume 𝐯\mathbf{v} is a /2\mathbb{Z}/2 harmonic form satisfying the conditions of Theorem 7.4. Then for every q𝒯M,𝐯q\in\mathcal{T}_{M,\mathbf{v}}, the distance to qq is a convex function on 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}}. By Lemma 7.6 and the maximum principal, the distance function to qq must be constant on the image of π𝐯\pi_{\mathbf{v}}. Since this holds for all qq, the map π𝐯\pi_{\mathbf{v}} must be constant and hence 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} contains only one point. By Condition (2) in Theorem 7.4, the leaf space 𝒯M,𝐯\mathcal{T}_{M,\mathbf{v}} contains infinitely many points, which yields a contradiction. ∎

Theorem 1.7 then follows as a special case of Theorem 7.4.

7.4. 𝒞\mathcal{C}^{\infty}–convergence for Korevaar–Schoen limits

Theorem 1.1 establishes a bridge between the analytic and algebraic compactifications of the moduli space of flat SL2()\mathrm{SL}_{2}(\mathbb{C}) connections on 33–manifolds. This connection allows us to prove new results on one side using results from the other. Let MM be a closed 33–manifold, let Γ=π1(M)\Gamma=\pi_{1}(M), and let {pi}\{p_{i}\} be a sequence of points in 𝒳(Γ)\mathcal{X}(\Gamma) that converges to a point in 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)}. Let ui:M~3u_{i}:\widetilde{M}\to\mathbb{H}^{3} be the corresponding equivariant harmonic maps (see the discussion in Section 6.2.1). Let E(ui)E(u_{i}) be the energy of uiu_{i} on a fundamental domain. By [DDW98, Thm. 2.2] (see also Section 6.2.1), after rescaling, the limit of uiu_{i} defines an equivariant harmonic map u:M~𝒯u:\widetilde{M}\to\mathcal{T} to a tree. By [KS97, Thm. 3.9], for every smooth vector field WW on M~\widetilde{M}, the sequence |ui(W)|/E(ui)1/2|\nabla u_{i}(W)|/E(u_{i})^{1/2} converges weakly to |u(W)||\nabla u(W)| in L2L^{2}. We use a result of Parker [Par23a, Thm.  1.3] to show that, in fact, |ui(W)|/E(ui)1/2|\nabla u_{i}(W)|/E(u_{i})^{1/2} also converges to |u(W)||\nabla u(W)| in CC^{\infty} on compact subsets of the complement of the zero locus of |u||\nabla u|.

Corollary 7.8.

Let ui,uu_{i},u be as above. Let 𝒰\mathcal{U} be the open subset of M~\widetilde{M} where |u|0|\nabla u|\neq 0. Then for every smooth vector field WW, we have |ui(W)|/E(ui)1/2|\nabla u_{i}(W)|/E(u_{i})^{1/2} converges in 𝒞\mathcal{C}^{\infty} to |u(W)||\nabla u(W)| on compact subsets of 𝒰\mathcal{U}.

Proof.

Without loss of generality, assume KK is a compact set contained in a fundamental domain of M~\widetilde{M}, and assume that WW is π1(M)\pi_{1}(M) invariant. Then |ui(W)||\nabla u_{i}(W)| and |u(W)||\nabla u(W)| reduce to functions on MM, and WW reduces to a vector field on MM. Assume (Ai,ϕi)(A_{i},\phi_{i}) is the solution to (1) corresponding to uiu_{i}, and let 𝐯SL2()\mathbf{v}\in\mathcal{M}_{\mathrm{SL}_{2}(\mathbb{C})} be the limit of the sequence (Ai,ϕi)(A_{i},\phi_{i}) in ¯SL2()\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}. By Theorem 1.1 and Corollary 6.2, we have |𝐯(W)|=|u(W)|/uL2(M)|\mathbf{v}(W)|=|\nabla u(W)|/\|\nabla u\|_{L^{2}(M)}.

We abuse notation and also use KK to denote the image of KK in MM. By [Par23a, Thm. 1.3], we know that every subsequence of |ϕi(W)|/ϕiL2(M)|\nabla\phi_{i}(W)|/\|\phi_{i}\|_{L^{2}(M)} has a subsequence that converges in 𝒞\mathcal{C}^{\infty} on KK. Hence, by (25), every subsequence of |ui(W)|/E(ui)1/2|\nabla u_{i}(W)|/E(u_{i})^{1/2} has a subsequence that converges in 𝒞\mathcal{C}^{\infty} on KK. Since |ui(W)|/E(ui)1/2|\nabla u_{i}(W)|/E(u_{i})^{1/2} weakly converges to |u(W)||\nabla u(W)|, the 𝒞\mathcal{C}^{\infty} limit of the subsequence must also be |u(W)||\nabla u(W)|. Therefore, |ui(W)|/E(ui)1/2|\nabla u_{i}(W)|/E(u_{i})^{1/2} converges to |u(W)||\nabla u(W)| in 𝒞\mathcal{C}^{\infty} on KK. ∎

7.5. Properties of the boundary of the analytic compactification ¯SL2()\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}

In this subsection, we discuss several properties of the /2\mathbb{Z}/2 harmonic 1-forms that actually appear in ¯SL2()\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})}. We compare them with classical results of measured foliations and the Hubbard-Masur map [HM79] over Riemann surfaces.

7.5.1. Leaf spaces of the analytic boundary

Observe that by Theorem 1.2, if a /2\mathbb{Z}/2 harmonic form 𝐯\mathbf{v} arises as a limit in Taubes’ compactification, then the leaf space of 𝐯\mathbf{v} has similar properties to the leaf spaces of quadratic differentials.

Corollary 7.9.

Suppose 𝐯¯SL2()\mathbf{v}\in\partial\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} and let ZZ be the zero locus of 𝐯\mathbf{v}. Let M~\widetilde{M} be the universal cover of MM and let 𝐯~\tilde{\mathbf{v}} denote the pull-back of 𝐯\mathbf{v} to M~\widetilde{M}. Let μ𝐯~\mu_{\tilde{\mathbf{v}}}, dM~,𝐯~d_{\widetilde{M},\tilde{\mathbf{v}}} be defined as in (12), (13), and let 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}} be the leaf space given by Theorem 4.4. Then

  1. (1)

    the quotient map from M~\widetilde{M} to 𝒯M~,𝐯~\mathcal{T}_{\widetilde{M},\tilde{\mathbf{v}}} is harmonic; and

  2. (2)

    if γ\gamma is an arc in M~\widetilde{M} from xx to yy that is transverse to ker𝐯~\ker\tilde{\mathbf{v}}, then μ𝐯~(γ)=dM~,𝐯~(x,y)\mu_{\tilde{\mathbf{v}}}(\gamma)=d_{\widetilde{M},\tilde{\mathbf{v}}}(x,y).

Proof.

Statement (1) follows from Theorem 1.2. By Theorem 1.1, 𝐯\mathbf{v} is in the image of \mathcal{H}. Statement (2) then follows from Corollary 6.2 and Lemma 5.18. ∎

7.5.2. Canonical measured foliations and the Hubbard–Masur map

We give a geometric interpretation for the restriction of Ξ¯:𝒳(Γ)¯¯SL2()\overline{\Xi}:\overline{\mathcal{X}(\Gamma)}\to\overline{\mathcal{M}}_{\mathrm{SL}_{2}(\mathbb{C})} to the boundary 𝒳(Γ)¯\partial\overline{\mathcal{X}(\Gamma)}, which equals |𝒳(Γ)¯\mathcal{H}|_{\partial\overline{\mathcal{X}(\Gamma)}}. In Morgan–Shalen’s theory [MS84, MS88a, MS88b], if []𝒳(Γ)¯[\ell]\in\partial\overline{\mathcal{X}(\Gamma)} is in the limit with the corresponding Γ\Gamma–tree 𝒯\mathcal{T}_{\ell}, the transverse equivalence map construction in [MS88b, Sec. I.2] defines a measured lamination from the harmonic map.

By comparison, for each []𝒳(Γ)¯[\ell]\in\partial\overline{\mathcal{X}(\Gamma)}, the corresponding /2\mathbb{Z}/2 harmonic form ([])\mathcal{H}([\ell]) defines a (singular) measured foliation on M~\widetilde{M} as in Section 4.1. Therefore, the map \mathcal{H} constructed in Section 6.1 canonically associates a (singular) measured foliation with every Morgan–Shalen limit point.

In fact, the boundary map \mathcal{H} can be viewed as a generalization of the Hubbard–Masur map. The original Hubbard–Masur map [HM79] establishes a homeomorphism between the space of projective classes of measured foliations (up to equivalence) and the space of quadratic differentials on a Riemann surface Σ\Sigma. As shown in Corollary 6.2, the map \mathcal{H} generalizes this concept to three dimensions, where harmonic maps to \mathbb{R}–trees take the place of measured foliations.

References

  • [BDO11] Fedor Bogomolov and Bruno De Oliveira, Symmetric differentials of rank 1 and holomorphic maps, Pure Appl. Math. Q. 7 (2011), no. 4, Special Issue: In memory of Eckart Viehweg, 1085–1103. MR 2918155
  • [BIPP23] Marc Burger, Alessandra Iozzi, Anne Parreau, and Maria Pozzetti, The real spectrum compactification of character varieties, arXiv preprint arXiv:2311.01892 (2023).
  • [BZ98] S. Boyer and X. Zhang, On Culler-Shalen seminorms and Dehn filling, Ann. of Math. (2) 148 (1998), no. 3, 737–801. MR 1670053
  • [CH24] Jiahuang Chen and Siqi He, On the existence and rigidity of critical Z2 eigenvalues, arXiv preprint arXiv:2404.05387 (2024).
  • [Chi76] I. M. Chiswell, Abstract length functions in groups, Math. Proc. Cambridge Philos. Soc. 80 (1976), no. 3, 451–463. MR 427480
  • [CM87] Marc Culler and John W. Morgan, Group actions on 𝐑{\bf R}-trees, Proc. London Math. Soc. (3) 55 (1987), no. 3, 571–604. MR 907233
  • [Cor88] Kevin Corlette, Flat GG-bundles with canonical metrics, J. Differential Geom. 28 (1988), no. 3, 361–382. MR 965220
  • [CS83] Marc Culler and Peter B. Shalen, Varieties of group representations and splittings of 33-manifolds, Ann. of Math. (2) 117 (1983), no. 1, 109–146. MR 683804
  • [DDW98] G. Daskalopoulos, S. Dostoglou, and R. Wentworth, Character varieties and harmonic maps to {\mathbb{R}}-trees, Math. Res. Lett. 5 (1998), no. 4, 523–533. MR 1653328
  • [DDW00] by same author, On the Morgan-Shalen compactification of the SL(2,𝐂){\rm SL}(2,{\bf C}) character varieties of surface groups, Duke Math. J. 101 (2000), no. 2, 189–207. MR 1738182
  • [Dee22] Ben K. Dees, Rectifiability of the singular set of harmonic maps into buildings, J. Geom. Anal. 32 (2022), no. 7, Paper No. 205, 57. MR 4425367
  • [DGR22] Nathan M. Dunfield, Stavros Garoufalidis, and J. Hyam Rubinstein, Counting essential surfaces in 3-manifolds, Invent. Math. 228 (2022), no. 2, 717–775. MR 4411731
  • [DLS11] Camillo De Lellis and Emanuele Spadaro, Q-valued functions revisited, vol. 211, American Mathematical Society, 2011.
  • [DM21] Georgios Daskalopoulos and Chikako Mese, Uniqueness of equivariant harmonic maps to symmetric spaces and buildings, 2021, arXiv:2111.11422.
  • [Don87] Simon Kirwan Donaldson, Twisted harmonic maps and the self-duality equations, Proc. London Math. Soc. (3) 55 (1987), no. 1, 127–131. MR 887285
  • [Don21] Simon Donaldson, Deformations of multivalued harmonic functions, Q. J. Math. 72 (2021), no. 1-2, 199–235. MR 4271385
  • [DR24] Nathan M. Dunfield and Jacob Rasmussen, A unified casson-lin invariant for the real forms of SL(2)\mathrm{SL}(2), Geometry and Topology, to appear, arXiv:2407.10922 (2024).
  • [DW20] Aleksander Doan and Thomas Walpuski, Deformation theory of the blown-up Seiberg–Witten equation in dimension three, Selecta Mathematica 26 (2020), 1–48.
  • [DW21] by same author, On the existence of harmonic Z2\rm Z_{2} spinors, J. Differential Geom. 117 (2021), no. 3, 395–449. MR 4255067
  • [GO89] David Gabai and Ulrich Oertel, Essential laminations in 33-manifolds, Ann. of Math. (2) 130 (1989), no. 1, 41–73. MR 1005607
  • [GS92] Michael Gromov and Richard Schoen, Harmonic maps into singular spaces and pp-adic superrigidity for lattices in groups of rank one, Publications Mathématiques de l’IHÉS 76 (1992), 165–246.
  • [Hat07] Allen Hatcher, Notes on basic 3-manifold topology, 2007.
  • [He22] Siqi He, Existence of nondegenerate 2\mathbb{Z}_{2} harmonic 1-forms via 3\mathbb{Z}_{3} symmetry, arXiv preprint arXiv:2202.12283 (2022).
  • [He23] by same author, The branched deformations of the special Lagrangian submanifolds, Geom. Funct. Anal. 33 (2023), no. 5, 1266–1321. MR 4646409
  • [HM79] John Hubbard and Howard Masur, Quadratic differentials and foliations, Acta Math. 142 (1979), no. 3-4, 221–274. MR 523212
  • [HMNW23] Siqi He, Rafe Mazzeo, Xuesen Na, and Richard Wentworth, The algebraic and analytic compactifications of the Hitchin moduli space, to appear in Moduli (2023), arXiv:2304.08198.
  • [HO96] A. Hatcher and U. Oertel, Full laminations in 33-manifolds, Math. Proc. Cambridge Philos. Soc. 119 (1996), no. 1, 73–82. MR 1356159
  • [HP24] Siqi He and Greg Parker, 2\mathbb{Z}_{2}-harmonic spinors and 1-forms on connected sums and torus sums of 3-manifoldss, arXiv preprint arXiv:2407.10922 (2024).
  • [Hub06] John Hamal Hubbard, Teichmüller theory and applications to geometry, topology, and dynamics. Vol. 1, Matrix Editions, Ithaca, NY, 2006, Teichmüller theory, With contributions by Adrien Douady, William Dunbar, Roland Roeder, Sylvain Bonnot, David Brown, Allen Hatcher, Chris Hruska and Sudeb Mitra, With forewords by William Thurston and Clifford Earle. MR 2245223
  • [Hub16] by same author, Teichmüller theory and applications to geometry, topology, and dynamics. Vol. 2, Matrix Editions, Ithaca, NY, 2016, Surface homeomorphisms and rational functions. MR 3675959
  • [HW15] Andriy Haydys and Thomas Walpuski, A compactness theorem for the Seiberg-Witten equation with multiple spinors in dimension three, Geom. Funct. Anal. 25 (2015), no. 6, 1799–1821. MR 3432158
  • [Ish79] Tôru Ishihara, A mapping of Riemannian manifolds which preserves harmonic functions, Journal of Mathematics of Kyoto University 19 (1979), no. 2, 215–229.
  • [Ish95] Hitoshi Ishii, On the equivalence of two notions of weak solutions, viscosity solutions and distribution solutions, Funkcial. Ekvac. 38 (1995), no. 1, 101–120. MR 1341739
  • [KNPS15] Ludmil Katzarkov, Alexander Noll, Pranav Pandit, and Carlos Simpson, Harmonic maps to buildings and singular perturbation theory, Communications in Mathematical Physics 336 (2015), 853–903.
  • [KS93] Nicholas J. Korevaar and Richard M. Schoen, Sobolev spaces and harmonic maps for metric space targets, Comm. Anal. Geom. 1 (1993), no. 3-4, 561–659. MR 1266480
  • [KS97] by same author, Global existence theorems for harmonic maps to non-locally compact spaces, Comm. Anal. Geom. 5 (1997), no. 2, 333–387. MR 1483983
  • [Lev93] Gilbert Levitt, Constructing free actions on \mathbb{R}-trees, Duke Math. J. 69 (1993), no. 3, 615–633. MR 1208813
  • [LTW22] John Loftin, Andrea Tamburelli, and Mochael Wolf, Limits of cubic differentials and buildings, arXiv preprint arXiv:2208.07532 (2022).
  • [Mes02] Chikako Mese, Uniqueness theorems for harmonic maps into metric spaces, Commun. Contemp. Math. 4 (2002), no. 4, 725–750. MR 1938491
  • [Mot88] Kimihiko Motegi, Haken manifolds and representations of their fundamental groups in SL(2,𝐂){\rm SL}(2,{\bf C}), Topology Appl. 29 (1988), no. 3, 207–212. MR 953952
  • [MS84] John W. Morgan and Peter B. Shalen, Valuations, trees, and degenerations of hyperbolic structures. I, Ann. of Math. (2) 120 (1984), no. 3, 401–476. MR 769158
  • [MS88a] by same author, Degenerations of hyperbolic structures. II. Measured laminations in 33-manifolds, Ann. of Math. (2) 127 (1988), no. 2, 403–456. MR 932305
  • [MS88b] by same author, Degenerations of hyperbolic structures. III. Actions of 33-manifold groups on trees and Thurston’s compactness theorem, Ann. of Math. (2) 127 (1988), no. 3, 457–519. MR 942518
  • [Oer88] Ulrich Oertel, Measured laminations in 33-manifolds, Trans. Amer. Math. Soc. 305 (1988), no. 2, 531–573. MR 924769
  • [OSWW20] Andreas Ott, Jan Swoboda, Richard Wentworth, and Michael Wolf, Higgs bundles, harmonic maps, and pleated surfaces, to appear in Geometry and Topology (2020), arXiv:2004.06071.
  • [Ota15] Jean-Pierre Otal, Compactification of spaces of representations after Culler, Morgan and Shalen, Berkovich spaces and applications, Lecture Notes in Math., vol. 2119, Springer, Cham, 2015, pp. 367–413. MR 3330768
  • [Par23a] Gregory J Parker, Concentrating Dirac operators and generalized Seiberg-Witten equations, arXiv preprint arXiv:2307.00694 (2023).
  • [Par23b] by same author, Deformations of 2\mathbb{Z}_{2}-harmonic spinors on 3-manifolds, arXiv preprint arXiv:2301.06245 (2023).
  • [Par24a] Gregory J. Parker, A gauge-theoretic compactification for generic ends of the SL(2;)\mathrm{SL}(2;\mathbb{C}) representation variety on 3-manifolds, in preparation.
  • [Par24b] Gregory J Parker, Gluing 2\mathbb{Z}_{2}-harmonic spinors and Seiberg-Witten monopoles on 3-manifolds, arXiv preprint arXiv:2402.03682 (2024).
  • [Py23] Pierre Py, Lectures on Kähler groups, 2023.
  • [Sha02] Peter B. Shalen, Representations of 3-manifold groups, Handbook of geometric topology, North-Holland, Amsterdam, 2002, pp. 955–1044. MR 1886685
  • [Sko96] Richard K. Skora, Splittings of surfaces, J. Amer. Math. Soc. 9 (1996), no. 2, 605–616. MR 1339846
  • [Sun03] Xiaofeng Sun, Regularity of harmonic maps to trees, Amer. J. Math. 125 (2003), no. 4, 737–771. MR 1993740
  • [Tau13a] Clifford Henry Taubes, Compactness theorems for SL(2;)\mathrm{SL}(2;\mathbb{C}) generalizations of the 4-dimensional anti-self dual equations, arXiv preprint arXiv:1307.6447 (2013).
  • [Tau13b] by same author, PSL(2;){\rm PSL}(2;\mathbb{C}) connections on 3-manifolds with L2L^{2} bounds on curvature, Camb. J. Math. 1 (2013), no. 2, 239–397. MR 3272050
  • [Tau14] by same author, The zero loci of \mathbb{Z}/2 harmonic spinors in dimension 2, 3 and 4, arXiv preprint arXiv:1407.6206 (2014).
  • [Thu97] William P. Thurston, Three-dimensional geometry and topology. Vol. 1, Princeton Mathematical Series, vol. 35, Princeton University Press, Princeton, NJ, 1997. MR 1435975
  • [TW20] C. H. Taubes and Y. Wu, Examples of singularity models for /2\mathbb{Z}/2 harmonic 1-forms and spinors in dimensional three, Proceedings of the Gökova Geometry-Topology Conferences 2018/2019, Int. Press, Somerville, MA, 2020, pp. 37–66. MR 4251085
  • [TW24] by same author, Topological aspects of /2\mathbb{Z}/2\mathbb{Z} eigenfunctions for the Laplacian on S2S^{2}, J. Differential Geom. 128 (2024), no. 1, 379–462. MR 4773187
  • [Wol95] Michael Wolf, Harmonic maps from surfaces to \mathbb{R}-trees, Math. Z. 218 (1995), no. 4, 577–593. MR 1326987
  • [WZ21] Thomas Walpuski and Boyu Zhang, On the compactness problem for a family of generalized Seiberg-Witten equations in dimension 3, Duke Math. J. 170 (2021), no. 17, 3891–3934. MR 4340726
  • [Zha22] Boyu Zhang, Rectifiability and Minkowski bounds for the zero loci of /2\mathbb{Z}/2 harmonic spinors in dimension 4, Comm. Anal. Geom. 30 (2022), no. 7, 1633–1681. MR 4596625