harmonic 1-forms,
-trees, and
the Morgan-Shalen compactification
Abstract.
This paper studies the relationship between an analytic compactification of the moduli space of flat connections on a closed, oriented 3-manifold defined by Taubes, and the Morgan–Shalen compactification of the character variety of the fundamental group of . We exhibit an explicit correspondence between harmonic 1-forms, measured foliations, and equivariant harmonic maps to -trees, as initially proposed by Taubes. As an application, we prove that harmonic 1-forms exist on all Haken manifolds with respect to all Riemannian metrics. We also show that there exist manifolds that support singular harmonic 1-forms but have compact character varieties, which resolves a folklore conjecture.
Key words and phrases:
/2 harmonic forms, -trees, Morgan-Shalen compactification2020 Mathematics Subject Classification:
Primary: 58D27; Secondary: 14M35, 57K351. Introduction
Let be a closed, oriented Riemannian 3-manifold with fundamental group . Let denote the character variety of as defined in, for example, [CS83]. Note that as a set, the closed points of correspond to conjugacy classes of completely reducible representations, i.e. those that are either irreducible or direct sums of a –dimensional representation and its dual. The space has the structure of an affine algebraic variety which, if positive dimensional, admits an –invariant compactification , called the Morgan–Shalen compactification [CS83, MS84, MS88a, MS88b]. A boundary point in is given by the projective class of a length function for an isometric action of on an –tree.
In [Tau13b], Taubes introduced a compactification of the space of flat connections on –manifolds from an analytic perspective. Consider the set of solutions to the equations
(1) |
where is a connection on an bundle over , and is a section of , where is the adjoint bundle. Let denote the moduli space of solutions to (1) up to gauge transformations. We abuse notation and also use to denote the topological space of closed points of with the analytic topology. Then by the results in [Don87] and [Cor88, Thm. 3.3], is canonically homeomorphic to . The compactness results of Taubes [Tau13b] define a compactification of , where the boundary points are described by a class of objects called harmonic –forms. A harmonic –form can be regarded as a generalization to -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 –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 –trees and more generally to metric spaces of nonpositive curvature (NPC spaces) as developed by Wolf and Korevaar–Schoen [Wol95, KS93, KS97]. Let be the universal cover of . By the Corlette–Donaldson theorem, every completely reducible flat connection on defines a –equivariant harmonic map from to the –dimensional hyperbolic space form . 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 harmonic –forms appearing in the Taubes limit and the limiting harmonic maps to –trees. Such a relationship was proposed by Taubes in [Tau13a, pp. 12-14].
More precisely, let be the moduli space of harmonic 1-forms up to rescaling, equipped with the topology (see (6)), and let denote the Taubes compactification with boundary (see Definition 3.8). Let be the space of projective length functions of minimal –actions on complete –trees (see Section 2.2). In Section 6.1, we construct a map
(2) |
Using this, we define a map
(3) |
where is given by the Riemann–Hilbert correspondence on and by on . Then we have the following result.
Theorem 1.1.
The map is continuous and surjective.
Let be a harmonic form in the boundary of the compactified moduli space. In Section 4, we show that the leaf space of the pull-back of to naturally has the structure of an –tree, which we denote by . Let denote the projection from to . The following result describes the relationship between the leaf space and the Morgan–Shalen tree.
Theorem 1.2.
Let be a harmonic form in the boundary of the compactified moduli space, and assume satisfies . Suppose is the minimal tree with length function and is the equivariant harmonic map given by the Morgan–Shalen compactification. Assume is scaled so that its energy is equal to . Then
-
(1)
The map factors uniquely as the composition of and a continuous map .
-
(2)
The map is –Lipschitz.
-
(3)
The map is harmonic.
-
(4)
(see Section 5.1 for the definitions). Away from the zero set of , we have as two-valued differential forms.
The above results establish a bridge between the analytic and algebraic compactifications of the moduli space of flat connections on –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 harmonic forms by considering harmonic maps to –trees.
A non-zero harmonic 1-form is called non-trivial if has non-trivial holonomy; in other words, if it is not given by a single-valued –form. We say that is singular if there exists a point on where cannot be locally lifted to a single-valued form. Clearly, singular implies non-trivial, but the converse is false in general. Using the map constructed in Section 6.1, we prove the following theorem.
Theorem 1.3.
Suppose is a rational homology sphere, and suppose there exists a closed connected embedded surface such that is two-sided, –injective, and does not bound an embedded ball. Then for every Riemannian metric on , there exists a non-trivial harmonic 1-form on .
We remark that while all previous analytical constructions of non-trivial 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 satisfies the conditions in Theorem 1.3 if and only if is reducible or Haken. On the other hand, if a closed oriented 3-manifold has , then there always exist trivial harmonic forms that are given by usual harmonic –forms. Therefore, Theorem 1.3 has the following consequence.
Corollary 1.4.
Suppose a closed oriented 3-manifold is reducible or Haken. Then for every Riemannian metric on , there exist harmonic 1-forms on .
Since 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 harmonic spinors, which is a generalization of the concept of harmonic forms.
For the Seiberg–Witten equations with two spinors, Parker [Par24b] proved that all 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 harmonic 1-forms could not exists on any 3-manifold with compact 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 such that the character variety of is compact, and supports a singular 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 character varieties are zero-dimensional. Since character varieties are affine varieties, zero-dimensional character varieties are compact. There are infinitely many examples in [Mot88] such that are cyclic groups with odd orders. If is a rational homology sphere such that has no 2-torsion, every harmonic form on is singular. Hence the result is proved by Theorem 1.3. ∎
Corollary 1.5 gives the first example of harmonic 1-forms (and more generally, 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 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 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 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 harmonic –forms on manifolds that are diffeomorphic to , which is motivated by the relation of harmonic –forms and the representation variety. We refer to [HP24] for more discussion about this conjecture.
Conjecture 1.6.
Let be a Riemannian metric on . There exists no harmonic –form on .
Note that if a manifold has positive Ricci curvature, then the non-existence of harmonic –forms follows directly from the Weitzenböck formula (see Eqns. (7), (8)).
We approach this conjecture using the relationship between harmonic –forms and measured foliations, as developed in Sections 4 and 7 below. Given a harmonic 1-form on , the leaf space of the measured foliation defined by is an –tree (see Theorem 4.4). Under certain assumptions, we show that the projection map from 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 harmonic –form on satisfying both of the following conditions:
-
(1)
Every zero point of is cylindrical (see Definition 7.3).
-
(2)
For every arc in transverse to , where denotes the zero locus of , we have
Condition (1) is a local regularity requirement for the zero locus of ; it is weaker than the usual regularity assumption in the study of harmonic –forms as stated [Don21, He23, Par23b]. The notations and in Condition (2) refer to the transverse invariant measure and the distance function on the leaf space defined by (see (12) and (13)). We will show that Condition (2) always holds for harmonic –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 -trees
In this section, we briefly review relevant results about the Morgan–Shalen compactification and –trees from [CS83, MS84, MS88a, MS88b]. For a more comprehensive introduction, we refer the reader to the survey by Otal [Ota15].
2.1. –trees and length functions
An –tree is a metric space such that every pair of points is connected by a unique arc, and every arc is isometric to a closed interval in as a subspace of .
Let be a finitely generated group. A –tree is an –tree with an isometric action . A –tree is called minimal if there is no proper –invariant subtree. The length function is defined by
(4) |
By [CM87, Sec. (1.3)], every acts semisimply, i.e., the infimum is realized at some point in . By [MS84, Prop. II.2.15], is identically zero if and only if has a fixed point.
In general, depends only on the conjugacy class of and the conjugacy class of . So let be the set of conjugacy classes of , and let be the real projective space of nonzero functions on . If , then the class of in is called the projective length function.
A length function is called abelian if it is given by for some homomorphism . A ray in an –tree is the image of an isometric embedding of . A –tree is said to have a fixed end if there exists a ray such that for every , is also a ray. We will need the following result.
Theorem 2.1 ([CM87, Thm. 3.7]).
Assume is a minimal –tree with a non-trivial length function . Then is non-abelian if and only if acts without fixed ends. Moreover, if is non-abelian and is another minimal –tree with the same length function, then there exists a unique -equivariant isometry between and .
When is abelian, the –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 “–tree” as a “tree” when the –action is clear from the context.
2.2. The Morgan–Shalen compactification
The character of a representation is defined by . Clearly, depends only on up to overall conjugation. By definition, is the set of all possible characters. By invariant theory, has the structure of an affine complex algebraic variety. Locally, points are determined by finitely many characters from a sufficiently large generating set for . For more details, we refer to [CS83, Cor. 1.4.5].
The Morgan–Shalen compactification can be described as follows [MS84]: Let be the vector space of real-valued functions on with the weak topology, and the projective space of nonzero real-valued functions on with the quotient topology. Define a map by
Since , the element has an image in , which we denote by . Let denote the one-point compactification of with the inclusion map . The Morgan–Shalen compactification is then the closure of the embedded image of in by the map .
Note that the group is equal to the identity component of the isometry group of the 3–dimensional hyperbolic space form . For each representation , we define
It is straightforward to verify that for all (cf. [CS83]). Hence, if a sequence converges to a point in , then the limit of is the same as the limit of .
We record the following result for later reference.
Proof.
Since both and are Hausdorff, the space is Hausdorff. By [Ota15, Prop. 8], the projection of the closure of the image of to is compact. Since is compact, this implies that is compact.∎
The following result relates the limit of length functions with –trees.
Theorem 2.3 ([MS84, MS88a, MS88b]).
Given a sequence of representations , the following holds:
-
(i)
Suppose is bounded for each . Then, after passing to a subsequence if necessary, converges to an element in .
-
(ii)
Suppose for some . Then, after passing to a subsequence, there exists a minimal –tree , where the action is denoted by , such that its length function satisfies , and in .
Not every element can be realized as a length function of a tree (see [Chi76] and [CM87, p. 586]). If an element of is realized by the length function of a –tree, then it can be realized by a minimal –tree [CM87, Prop. 3.1]. Define to be the subspace of consisting of all elements that can be realized by the length functions of minimal –trees that are complete as metric spaces. For each , since is not identically zero, the –action on the corresponding tree has no fixed point.
3. Harmonic 1-forms and Taubes’ compactness results
In this section, we review the compactness results on stable flat connections due to Taubes [Tau13b, Tau13a]. We will use these results to define a compactification of the moduli space of flat connections using harmonic 1-forms.
3.1. harmonic 1-forms.
Let be a Riemannian manifold with metric . Here, is allowed to be non-compact or non-complete but we assume that has no boundary. The concept of a harmonic –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 harmonic –form on is given by a closed subset and a two-valued section of on the complement of , such that the following conditions hold:
-
(i)
For each , there exists an open neighborhood of such that the values of on have the form , where is a non-vanishing section of such that , .
-
(ii)
For every open subset of such that is compact, we have , and .
-
(iii)
Let be the dimension of . There exist constants and such that for every and every less than the injectivity radius of at , we have .
Here, and are defined pointwise as the values of and where satisfies Condition (i). Condition (iii) above is needed for the proofs of some key properties of harmonic –forms. The set is called the zero locus of .
Example 3.2 ([Tau13b, p. 9]).
When is a Riemann surface, a harmonic 1-form is given by the real part of the square root of a holomorphic quadratic differential. More explicitly, given a quadratic differential , where is the canonical bundle, defines a harmonic 1-form. Conversely, given a harmonic 1-form, let be the component of , then also defines a quadratic differential. This correspondence is a bijection. In particular, there exists no harmonic 1-form on .
We have the following analytic property for the zero loci of harmonic 1-forms.
Theorem 3.3 ([Tau14, Zha22]).
Assume is connected and the dimension of is . Then the zero locus of every harmonic –form on is –rectifiable with locally finite –Hausdorff measure.
In particular, the zero locus of every harmonic –form has Lebesgue measure zero. Therefore, we will sometimes write as when the integrand is defined on the complement of and extends to an function over .
It is convenient to regard a harmonic 1-form as a section of the bundle over , where the fiber of at each point is the quotient space . A distance function on the quotient space is defined by
(5) |
We define the moduli space of harmonic 1-forms as
(6) |
and endow with the topology induced from the distance function (5).
Proposition 3.4.
Assume is closed and has dimension no greater than 4. Then the space is compact.
Proof.
We first show that there exists a constant depending only on and the Riemannian metric, such that for each with zero locus , we have
By the Weitzenböck formula, near every , if we write as such that , then
(7) |
where is the Ricci curvature of .
We show that
(8) |
For each positive integer , let be a smooth function on such that when , when , and for all . Define by
then is a smooth function on .
On , we have
Hence the following inequality holds:
Also note that the support of is a subset of . Since , we conclude that
and hence converges to zero in as goes to infinity.
By integration by parts, we have
(9) |
By (7), the function is integrable on , and hence the limit of the left-hand side of (9) equals . Since and in as goes to infinity, the right-hand side of (9) converges to zero as goes to infinity. Hence (8) is proved. By (7), we conclude that for some constant depending only on .
The Sobolev space for multi-valued sections of a vector bundle was studied in [DLS11, Ch. 4]. By [Zha22, Lem. 2.1], the harmonic 1-form can be regarded as a –valued section with regularity over , and its –norm is equal to (up to constant multiplicative factors) . 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 is Hölder continuous.
Proposition 3.5.
Assume is closed. Then the topology and the topology coincide on .
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 convergence implies convergence; we show that the converse also holds. Namely, assuming is a sequence in that converges to in , we show that converges to in .
Let denote the injectivity radius of . By [Tau14, Lem. 2.3] and a rescaling argument, there exists a constant depending only on such that
(10) |
for all , , and .
Let be a sequence in that converges to in the topology. Let be a constant such that for every and . For each , let be the set consisting of all such that , where is the constant in (10). Let be the distance between and . Take sufficiently small so that .
Since converges to in , for sufficiently large, we have
for all . Let . By (10), this implies
So the –distance between and on is no greater than
On the other hand, converges to uniformly on by standard elliptic estimates. Hence the desired result is proved. ∎
3.2. The moduli space and its compactification
Let be a closed oriented 3-manifold. Let be a principal bundle over , and let be the associated bundle given by the adjoint action. Consider the system of equations
(11) |
where is a connection on and is a section of . We define to be the set of solutions to (11), modulo gauge transformations. The topology on is given by the –Sobolev norm for sufficiently large. By the standard elliptic bootstrapping argument, the topology on does not depend on the choice of when and is sufficiently large.
Note that if is a harmonic –form, then is a single-valued section of . The following result is a consequence of Taubes’ compactness theorems.
Theorem 3.6 ([Tau13b, WZ21, Par23a]).
Let be a sequence of solutions to (11), and let .
-
(i)
If is bounded, then there exists a subsequence of that converges in after gauge transformations.
-
(ii)
If , then there exists a subsequence (which we still denote by ) and such that converges to in .
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 , which we denote by the same notation, and a with zero locus , such that
-
(1)
converges to in .
-
(2)
For every compact set contained in an open ball in , there exists on such that, after a sequence of gauge transformations, converges to in the weak topology on .
-
(3)
The spinor in (ii) satisfies .
Statements (1) and (2) above are directly given by [WZ21, Thm. 1.28]. A stronger result was proved in [Par23a, Thm. 1.3], where was shown to converge to in the topology on . 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 , let be the set of such that . Then is a closed subset of that contains an open neighborhood of . By Statement (1) above, for sufficiently large, we have
Hence the distance between and is less than on when is sufficiently large. By Statement (2) above and the Sobolev embedding theorems, after passing to a subsequence if necessary, the sequence converges to on in the topology. Since is gauge invariant, the result is proved. ∎
Next, we define a compactification of using Theorem 3.6. Let be the disjoint union of and . We define a topology on as follows. The open sets on are generated by the following two collections of subsets:
-
(i)
Open subsets of .
-
(ii)
For , , , the subset of that contains all such that the distance between and is less than , and all such that
-
(a)
,
-
(b)
the distance between and is less than .
-
(a)
Note that for , we have if and only if , and that the topology on is the same as the pull-back topology from via the map . Therefore, and are homeomorphic to their embedded images in . Moreover, Proposition 3.4 and Theorem 3.6 can be summarized into the following statement.
Corollary 3.7.
The space is Hausdorff and compact.
Proof.
Definition 3.8.
Define to be the closure of in .
We call the compactification of .
Remark 3.9.
One can construct a more refined compactification of from the analytical results in [Tau13b, HW15, WZ21, Par23a] by also considering the limits of the connection terms . 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 harmonic 1-forms
In this section, we review a construction of measured foliations from harmonic –forms by Taubes [Tau13b, p. 14]. Then we prove that on simply connected manifolds, the leaf space of the measured foliation defined from a harmonic –form is always an –tree.
4.1. Measured foliations defined by 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 harmonic –form, following Taubes [Tau13b, p. 14].
We first introduce the concept of a singular measured foliation.
Definition 4.1.
Let be a manifold with dimension . Let be a closed subset of Hausdorff codimension at least . A (codimension-one) singular foliation on with singular set is a smooth foliation on with codimension-one leaves. A transverse measure on is a measure for arcs in such that the following conditions hold:
-
(i)
is non-zero on transverse arcs.
-
(ii)
vanishes on an arc if and only if the arc is tangent to a leaf.
-
(iii)
(Holonomy invariance) 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 harmonic 1-form. To provide some intuition, recall from Example 3.2 that, on a Riemann surface, a 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 is a closed Riemann surface. Given a nonzero holomorphic quadratic differential , the zeros form the singular set of a foliation. Away from , one can write locally as for some holomorphic –form . The kernel of then defines a foliation, with a transverse measure given by . This local construction can be combined to form a global measured foliation, often referred to as the vertical measured foliation of .
This construction can be generalized to harmonic 1-forms as in [Tau13a]. Consider a Riemannian manifold (not necessarily closed or complete), and let be a harmonic 1-form on . We define the singular set as . By Theorem 3.3, if the dimension of is no greater than , then meets the conditions described in Definition 4.1.
For a point , let be a small neighborhood around where can be locally expressed as , with a single-valued 1-form on . By Definition 3.1, is a closed 1-form. The kernel of thus defines a codimension-one foliation on , which is independent of the choice of sign for . Consequently, defines a smooth foliation on .
We define a transverse measure for this foliation as follows. For each arc , the transverse length of is given by
(12) |
It’s important to note that, similar to the cases of and , the value of is well-defined despite being a two-valued section. Since is locally given by a closed smooth form in , the measure is holonomy invariant. As a result, the pair defines a measured foliation on .
4.2. The leaf spaces of harmonic 1-forms
Next, we define the leaf space associated to a harmonic form. We prove that if the background manifold is simply connected, then the leaf space is an –tree.
Consider a connected Riemannian manifold (not necessarily closed or complete), and let be a harmonic –form on . We define a pseudo-metric on by
(13) |
where ranges over all piecewise curves from to . Note that the definition of depends globally on both and : if is an open subset of , then it is not necessarily true that equals the restriction of on .
Definition 4.3.
The metric space is the quotient space of with respect to the pseudo-metric .
We call the space the leaf space of on , and we will use to denote the metric on this quotient space as well.
Theorem 4.4.
Suppose is a simply connected manifold, is a harmonic –form on with zero locus . If the dimension of is greater than , we assume in addition that is continuous and is rectifiable with locally finite Hausdorff measure in codimension . Then the metric space is an –tree.
The remainder of this section is devoted to the proof of Theorem 4.4. We first recall the following standard result about –trees.
Proposition 4.5 ([Py23, Prop. B.31]).
A metric space is an –tree if and only if it is path connected and the following inequality holds for all :
Now we prove the following technical lemma.
Lemma 4.6.
Suppose is the closed unit disk in , and let denote the interior of . Suppose is a continuous, two–valued –form defined on an open neighborhood of in with zero locus , such that away from its zero points is locally given by with being closed. We further assume that is finite, is compact, and that for each , there exists a smooth domain such that contains , and is contained in the –neighborhood of , and the total length (i.e. the Hausdorff measure in dimension ) of is less than . Finally, assume that is transverse to at all but finitely many points.
Let be the pseudo-distance function on associated with given by (13). Let be four distinct points appearing in cyclic order on , the points divide into four arcs.
Then there exist two points and on a pair of opposite (closed) arcs divided by such that .
Proof.
We adapt an argument from [Lev93, Lem. III.4]. To simplify notation, we will write as in the proof. Assume are ordered counterclockwise. In the following, for , we use to denote the closed arc in bounded by , where the arc goes from to in the counterclockwise direction. If are compact subsets of , we use to denote .
If , then the desired result already holds. In the following, assume . Let be such that and that is furthest away from as a point on the arc among all points satisfying this condition. Similarly, let be such that and that is furthest away from among all points in satisfying this condition.
We discuss three cases. If , then the above conditions imply that , and the desired result holds.
If and (that is, is on the counterclockwise side of ), then by the above conditions, for every , there exists a curve from to , and a curve from to , whose lengths with respect to are less than . These two curves must intersect, so . Since this statement holds for all , we conclude that , which implies , and the desired result holds.
If and (that is, is on the counterclockwise side of ), we consider the singular foliation defined by on . Since is locally given by closed forms away from zero points, there is a transverse invariant measure on as given by (12).
For each point , if is defined and is transverse to at , we consider the leaf of passing through . View the leaf as a parametrized curve starting at . By the Poincaré–Bendixson theorem, one of the following holds:
-
(1)
The leaf intersects a boundary point of other than .
-
(2)
The leaf converges to a zero point of .
-
(3)
The leaf converges to a limit cycle.
Since admits a transverse measure, Case (3) cannot happen. By the definitions of , we know that , . By the assumption that , we know that . As a result, if Case (1) happens, then the intersection point of the leaf with is in the interior of .
Let be the set of points such that
-
(1)
is non-zero at and is transverse to at .
-
(2)
The leaf of starting at transversely intersects a point of other than .
Then is an open subset of . We claim that
(14) |
This is because the leafs emanating from the zero points of have zero measure with respect to . More precisely, recall that denotes the zero set of . For each , there exists such that on the –neighborhood of . Let be the smooth domain given by the assumptions on , then . Then the set of points such that there is leaf of intersecting transversely at and enters has –measure less than . Similarly, if is a zero point of in , then as , so the set of points such that there is leaf of intersecting transversely at and converges to on the other end has –measure zero. Recall that for every point , one of the following conditions holds:
-
(1)
is a zero point of ,
-
(2)
is a point where is tangent to ,
-
(3)
is a point whose leaf passes through a tangent point of with ,
-
(4)
is a point whose leaf converges to a zero point of .
The first three cases only contain finitely many points, and points in the last case have measure zero with respect to by the previous argument. Hence Equation (14) is proved. Since is transverse to at all but finitely many points, we conclude that has full measure in with respect to the standard Lebesgue measure as well.
Define an involution , such that for every , the image is the other endpoint of the leaf of passing through . The map is orientation-reversing.
Parameterize the arc by the interval via a smooth diffeomorphism: . Define a function
where we recall that denotes the pseudo-distance function . Then a Lipschitz function on , so exists almost everywhere and
For each open interval , let and , we have
Hence the involution on reverses the orientation and preserves the differential form . This implies . Since has full Lebesgue measure in , we conclude that , so . This contradicts the definitions of . ∎
We now prove Theorem 4.4.
Proof of Theorem 4.4.
Let be distinct points in . By Lemma 4.6, we only need to show that
(15) |
Since (15) is a closed condition on , we may assume without loss of generality that are not zero points of . For notational convenience, we will interpret the subscripts modulo . Fix . For each , let be a smooth arc from to such that
After perturbation, we may assume that the union of forms a smoothly embedded circle in that is disjoint from the zero locus of . Denote this circle by . After a further perturbation if necessary, we may assume that is transverse to at all but finitely many points. Since is simply connected, there exists a smooth immersion , where is the unit disk in , such that . Since the zero locus of has locally finite codimension two Hausdorff measure, we may perturb , with fixed, so that intersects the zero locus of at finitely many points, and that the zero locus of is discrete on the complement of . As a result, the assumptions of Lemma 4.6 holds for .
Applying Lemma 4.6 to for and to the –valued form , we conclude that there exists on opposite arcs such that . Since , this implies . Assume without loss of generality that is in the arc bounded by , and is in the arc bounded by . Then we have
By the assumptions on , we have
and similarly,
As a result,
The desired result then follows by taking . ∎
5. Harmonic maps to –trees
Let be a closed –manifold and its universal cover. In this section, we show that if is a –equivariant harmonic map to an –tree, then the gradient of defines a harmonic 1-form on . We first recall the formulation of Korevaar–Schoen which gives meaning to the gradient of as a Radon–Nikodym derivative. Then, we use regularity results of harmonic maps to construct the associated harmonic –form.
5.1. Energy density and directional derivatives
Here, we review some terminology from [KS93], specialized to the case of norms. Following the notation of [KS93], let be an -dimensional Riemannian manifold and a complete metric space. Unless otherwise specified, is allowed to be non-compact or non-complete but we assume that has no boundary. Later, we will take to be the universal cover of a closed –manifold and to be an –tree.
For , let be the set of all points such that the exponential map at is well-defined on the open ball with radius . Let denote the volume measure of .
A map is called locally if for every compact subset in and every , we have
If is locally , define the –approximate energy function as
(16) |
where is the area of in (recall that is the dimension of ), and is the area form on . The map is said to have finite energy, if
The space of maps from to with finite energy is denoted by . By [KS93, Thm. 1.10], if , then the measures converge weakly to a limit as , where . The function is called the energy density function, and we define .
The directional derivatives are defined in a similar way. Assume is a Lipschitz vector field over . For , let denote the image of after flowing along for time . Define . By [KS93], if , then there exists a unique non-negative function such that the measures converge weakly to .
A map 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 is an –tree, then harmonic maps to minimizes the energy with respect to compactly supported perturbations.
Now assume is a closed manifold and let be its universal cover. Assume is a complete –tree with a fixed isometric action by . 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 action on a complete tree has no fixed ends (see the definitions above Theorem 2.1). Then there exists a –equivariant harmonic map . Moreover, if are two –equivariant harmonic maps such that , then either or 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 action on an –tree . Let be two equivariant harmonic maps to , and let be a Lipschitz vector field on . Then we have (almost everywhere).
5.2. Regularity of harmonic maps to –trees
Now we review some regularity results on harmonic maps to –trees from the literature.
Definition 5.3.
Let be a harmonic map from a manifold to a tree. We say that is a regular point of , if there exists an open ball centered at such that is contained in a geodesic in . If is not a regular point, then it is called a singular point. We denote the set of regular points by and the set of singular points by .
If is a regular point of , then locally is equal to the composition of a harmonic map to a closed interval , and an isometric embedding of in . It is clear from the definition that is open and is closed.
Theorem 5.4 ([Sun03, Thm. 1.4]).
Let be a harmonic map from a manifold to an –tree. Then, for every point , there exists an open ball centered at such that lies in an embedded locally finite subtree.
Theorem 5.4 allows one to reduce the regularity problem of harmonic maps into –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 is harmonic. For , the order of at is defined to be the limit
(17) |
where denotes the volume measure of and denotes the area form.
It was proved in [GS92, Sec. 2] that the limit in (17) always exists. When , the value of equals the vanishing order of at . Moreover, the order function is upper semi-continuous with respect to ([GS92, statement above Thm. 2.3]).
We collect the following regularity results from the literature, which will be used later. In the following, denotes a harmonic map, and denotes the metric on the –tree .
Theorem 5.6 ([GS92, Thm. 6.3], [Sun03, Thm. 1.1]).
There is a constant depending only on the dimension of the domain, such that for each point , either , or . Moreover, if , then is a regular point.
Theorem 5.7 ([KS93, Thm. 2.4.6]).
The map is locally Lipschitz, and the Lipschitz constants only depend on the curvature of .
Theorem 5.8 ([Sun03, Thm. 1.3], [GS92, Thm. 6.4]).
The singular set has Hausdorff codimension at least .
A stronger version of this result was given by Dees [Dee22].
Theorem 5.9 ([Dee22, Thms. 1.1 and 1.2]).
Assume has dimension , then is –rectifiable. Moreover, for each compact set , there exists a constant depending on , , and , such that
for , where denotes –neighborhood of .
Theorem 5.10 ([GS92, proof of Thm. 2.3]).
Assume as a manifold is the open unit ball in , but the metric is not necessarily Euclidean. Suppose . Then there exists a constant depending only on the curvature, such that for all and , we have
(18) |
Corollary 5.11.
Assume . Let be the dimension of . Let be the minimum of and the injectivity radius of at . Assume . Then there exists a constant depending only on the curvature of in , such that
Proof.
By [GS92, Sec. 2], there exists a constant depending only on the curvature, such that
is increasing with respect to for all . Hence
By Theorem 5.7, we have for a constant . By Theorem 5.10, we have
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 and is less than the injectivity radius of at . Then there exists a constant , depending only on the curvature of on , such that
Remark 5.13.
We now prove the following estimate.
Proposition 5.14.
Assume is a compact subset of . Then
Proof.
By Theorem 5.8, has zero Lebesgue measure. Hence we may write as , if the integrand is defined on and extends to an integrable function on .
By shrinking to a smaller domain containing , we may assume that the curvature of is uniformly bounded and has finite volume. By Theorem 5.7, this implies is bounded and .
By [GS92, eq. (6.2)], there is a constant depending only on the curvature of such that
(19) |
pointwise on .
Let be a Lipschitz function that is compactly supported in such that . By (19),
so
As a result,
(20) |
By the assumptions on , we have
For every positive integer , let be the Lipschitz function such that for , for , and for . Let be a smooth function that is compactly support in , such that , and on the given compact set . Define by
where is the distance function on . Theorem 5.9 implies that is bounded as . Hence by (20), the integral is bounded as . Since is increasing with respect to and converges to as , this implies is finite, so the desired result is proved. ∎
5.3. From harmonic maps to harmonic 1-forms
Now assume is a closed oriented –manifold and let be its universal cover. Let be the covering map. Assume is an –tree with an isometric –action, and assume is a –equivariant harmonic map.
Then defines a two-valued –form on . It is two-valued because the geodesics on do not have canonical orientations. Locally, is given by for some (single-valued) harmonic –form . Since is equivariant, defines a two-valued –form on the image of in .
Theorem 5.15.
There exists a unique harmonic –form on , such that on .
Proof.
The uniqueness is clear since has Hausdorff codimension . We prove the existence of such . Recall that is continuous (see Remark 5.13). Let . Then , so defines a non-vanishing two-valued –form on . Since is equivariant, the set is invariant under the –action. Let be the image of in . Then is the pull-back of a two valued –form on .
We show that is a harmonic –form with respect to the zero locus . Since is harmonic in , we know that is locally given by non-vanishing harmonic –forms on .
Note that if and , then . As a result, by Theorem 5.6, there exists such that for all . By Corollary 5.11, we have
for all and , where is the minimum of and the injectivity radius of at . Since is upper semi-continuous and is closed, it is bounded from above on . Since is closed, the value of has a positive lower bound. Therefore, the value of has an upper bound (which may depend on ). This verifies Condition (iii) of Definition 3.1.
Remark 5.16.
From the above construction, we also see that . In other words, every non-vanishing point of the harmonic –form corresponds to regular points of on .
Definition 5.17.
We call the harmonic form obtained by Theorem 5.15 the harmonic 1-form associated with .
5.4. Maps between trees
Let , , , , be as in Section 5.3. Let be the harmonic –form on associated with . Let be the pull back of to . By Theorem 4.4, the leaf space is also an –tree. In this subsection, we study the relationship between and .
Let be the transverse measure defined from by (12). Let , denote the zero loci of and .
Lemma 5.18.
Suppose is a arc. Let , . Then
(21) |
Moreover, equality holds if is transverse to .
Proof.
We first prove (21). Both sides of (21) are continuous with respect to in the topology, so the inequality is a closed condition. Recall that denotes the regular set of . Since has Hausdorff codimension at least , after perturbing if necessary, we may assume without loss of generality that the image of is contained in .
If is transverse to , then the image of is contained in the complement of . By Remark 5.16, the image of is contained in . Then for , the value of locally moves along geodesics at non-vanishing velocities. Since is an –tree, this implies that the image of for is a geodesic segment where the point moves at non-vanishing velocities with respect to . Hence the inequality in (22) achieves equality for all , and the desired result is proved. ∎
Let denote the quotient map from to .
Theorem 5.19.
There exists a unique continuous map such that . Moreover, we have the following properties:
-
(1)
is –Lipschitz.
-
(2)
The map is harmonic.
-
(3)
.
-
(4)
If , then as two-valued harmonic 1-forms.
Proof.
By Lemma 5.18, we have
for all . Hence the map factorizes uniquely as for a continuous map , and is –Lipschitz. As a result, we have .
By the definition of , the map is Lipschitz with Lipschitz constant , so the map is in .
For , let be an arc that is transverse to . Then
Here, the first inequality follows from the fact that is –Lipschitz, the second equality follows from Lemma 5.18, and the third inequality follows from the definition of . Hence . On the other hand, if is tangent to and is contained in the complement of , then both and are zero. In conclusion, if is in the complement of , then there exists an open ball centered at such that
for all . This implies that the directional derivatives and energy densities of and are the same on . Since has Lebesgue measure zero, they also define the same measure density functions on .
It remains to show that is harmonic. Let be a fundamental domain of the –action. An equivariant map from to a tree is harmonic if and only if it minimizes the energy on among all locally equivariant maps. We show that minimizes the energy on . Assume there exists equivariant map such that
then we have
where the first inequality follows from the fact that is –Lipschitz, and the last equation follows from the fact that on . Since is –equivariant, this contradicts the assumption that is energy minimizing. Hence the theorem is proved. ∎
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 in (2) is constructed. Following our previous convention, let denote an oriented closed Riemannian 3-manifold, and let be its universal cover. Let .
6.1. Definition of the boundary map
Recall that denotes the space of projective length functions of all complete minimal –trees, and denotes the space of all harmonic –forms on with unit –norm. In this subsection, we define a map
(23) |
By Theorem 2.1, if is non-abelian, then there is a unique (up to equivariant isometry) minimal –tree with length function and no fixed ends. By Theorem 5.1 there is a equivariant harmonic map . By Proposition 5.2, (see also Definition 5.17), the harmonic –form associated with is independent of the choice of or . Let be the harmonic –form associated with . Define
Since only depends on the projective class of in , the value of is well-defined.
Next, we define the map on abelian length functions. Assume is an abelian length function. By definition, there exists a homomorphism so that for all . It follows from straightforward algebra that if are two homomorphisms such that for all , then . In other words, is determined by up to an overall sign.
Define to be the (single-valued) harmonic –form on such that the periods of on are given by . Let be the harmonic –form given by . Define
In conclusion, we have constructed a well-defined map from to .
Lemma 6.1.
Suppose is a minimal tree whose length function is abelian and not identically zero, and is a –equivariant harmonic map. Then the tree must be , and the action on is given by translations.
Proof.
Suppose that has a minimal action on a tree not isometric to , but with an abelian length function. By [CM87, p. 573], the action of must have a fixed end. Combining [DDW98, Thm. 5.3] and [Mes02, Thm. 1.2], we see that there can be no -equivariant harmonic map to .
Now that we have proved , we show that the action must be given by translations. Assume acts on by an orientation-reversing isometry, then . As a result, for every , we must have . This is only possible when the action of equals the identity or the action of . Hence the length function of the –action is identically zero, contradicting the assumptions. ∎
Corollary 6.2.
Assume is a complete minimal –tree and is a –equivariant harmonic map. Let be the length function of , and assume is not identically zero. Then the harmonic –form associated with is equal to up to a non-zero constant multiplication.
Proof.
If is non-abelian, the statement follows from the definition of . If is abelian, then Lemma 6.1 implies that and the action of on are translations. Therefore, is a harmonic function and the associated harmonic –form is given by . Since the periods of coincide with the translation distances of the –action on , the desired statement is proved. ∎
6.2. Comparison of the compactification limits
Recall that denotes the character variety of , and that denotes the moduli space of solutions to (1). Also recall that and are canonically identified by the Riemann–Hilbert correspondence. In this subsection, we prove the following proposition.
Proposition 6.3.
Assume is a sequence of points in , and let be the corresponding sequence in . Assume that converges in the Morgan–Shalen compactification to the projective length function , and that converges in the Taubes compactification to a harmonic –form . Then .
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 to . More precisely, denotes the three-dimensional hyperbolic space form. Its orientation-preserving isometry group is isomorphic to . Every is given by a representation of in , which defines a –action on . Let be a –equivariant harmonic map from to with respect to the above –action. Let be the energy of on a fundamental domain, and let denote the pseudo-metric on defined by
(24) |
If converges to a projective length function on the boundary of the Morgan–Shalen compactification, then , and converges locally uniformly to a limit . The quotient metric space of with respect to is a minimal –tree , the length function of is in the projective class , and the quotient map from to is harmonic.
6.2.2. Characterization of the Taubes limit
If converges to a harmonic –form , assume each is given by the pair . Let be the principal bundle over where is defined. Let be the associated bundle with respect to the adjoint action.
Let be the zero locus of . By the constructions in Section 3, on every compact set that is contained in an open ball in , there exists on , such that after a sequence of gauge transformations, converge to in the weak topology on . By Sobolev embedding theorems, converge to in the topology.
By part (3) of the second bullet point of [Tau13b, Thm. 1.1a], the limit spinor locally has the form , where is a harmonic –form, is a section of , and pointwise. In this case, the harmonic –form is locally equal to .
6.2.3. Relationship between the two constructions
The relationship of the maps and is given as follows (see [DDW98, Sec. 2C]). Let and denote the pull-backs of and to . The orthonormal frame bundle of is an bundle. Its Lie algebra bundle is an bundle, and the pull-back of this bundle to is isomorphic to . Then there exists an equivariant isomorphism such that under the identification of Lie algebra bundles, where is given by the Levi-Civita connections of and . Different choices of equivariant isomorphisms of Lie algebra bundles correspond to gauge transformations on solutions to (1).
As a consequence, we have
(25) |
for every tangent vector of . 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 denote the quotient map from to . Recall that denotes the harmonic form in the Taubes limit. Let denote the pull-back of to . Let be the set of such that is not on the zero locus of or the zero set of . Then is a –equivariant open and dense subset of . The set is contained in the regular set of , so is a non-vanishing harmonic –form on .
As a result, and are two smooth foliations on . We first show that these two foliation are the same.
For each , let be an open ball centered at such that
-
(1)
,
-
(2)
the metrics (defined in Equation (24)) converge to uniformly on ,
-
(3)
uniformly converge to a limit on after gauge transformations.
Let be a smooth arc that is tangent to . Write , . By (25), we have
Recall that is locally given by , where is a harmonic –form such that locally . Since is tangent to , we have for all . This implies . As a result, every leaf of is contained in a leaf of as foliations on . Therefore, on .
Next, we show that and differ only by a locally constant multiplication on . Let be as above. Since is simply connected, both and can be lifted to single-valued harmonic –forms. Locally, write as and as . Both and are non-vanishing. Since , there exists a non-zero function such that . Then we have
As a result, , , so , thus is a constant function. In conclusion, and differ only by a locally constant multiplication on .
Since the zero loci of and have Hausdorff codimension , the above result implies that is the harmonic –form associated with up to a constant multiplication. By Corollary 6.2, we conclude that . ∎
6.3. Proofs of Theorems 1.1 and 1.2
Define
such that is given by the (inverse of the) Riemann–Hilbert map on and by on .
Lemma 6.4.
Both and are metrizable.
Proof.
By Corollary 3.7, the space is compact and Hausdorff, and hence it is regular. The space defined in Section 3.1 is separable with respect to the –topology, because it is a closed subset of the space of continuous –valued sections of , which is a separable metric space. Hence the topology on defined in Section 3.2 is second countable. By the Urysohn metrization theorem, we conclude that is metrizable.
Similarly, by Lemma 2.2, the space is compact and Hausdorff. Since is countable, the space is second countable. This implies is second countable. Hence is metrizable. ∎
The following statement is the first part of Theorem 1.1:
Theorem 6.5.
The map is continuous.
Proof.
The result follows from Proposition 6.3 and a formal argument. Since both and are metrizable, we only need to show that if is a sequence that converges to in , then converges to . Since is an open subset of and the map is already known to be continuous on , we only need to prove the statement when . We only need to consider two cases: (1) all ’s are in ; (2) all ’s are in .
If all ’s are in , then by the compactness of , every subsequence of has a convergent subsequence. By Proposition 6.3, this subsequence must converge to . Hence converge to .
If all ’s are in , we use an argument by contradiction. Let be a metric for the topology of and let be a metric for . Assume there exists such that there is a subsequence of , which we will still denote by , such that for all . For each , since , there exists a sequence in , such that converges to as . By Proposition 6.3, we have
As a result, we may find a in the sequence such that , , and . The sequence then satisfies and
for all . This contradicts Proposition 6.3. ∎
Now we prove the second part of Theorem 1.1:
Theorem 6.6.
The map is surjective.
Proof.
We need to show that every point is in the image of . Let be a sequence of points in such that . Let be the preimage of in . By the compactness of , there is a subsequence of , which we still denote by , such that converges to a point in . Since is continuous, we have . ∎
6.4. Non-injectivity of the map
In general, the map 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 –dimensional analogue. Multiplying the spaces with will yield a counterexample for the –dimensional case. Specifically, we construct distinct length functions appearing in with the same image under .
Let be a Riemann surface of genus . Express the surface as a connect sum: of surfaces of genus and , respectively. This gives an amalgamated product expression for the fundamental group: , where denotes the free group on generators.
Let be a divergent family of completely reducible representations converging in the Morgan–Shalen limit to a nonabelian projective length function . This is possible because , so we may take to be discrete and faithful, for example.
Now by the product expression, the extend to representations of , and they clearly converge to a Morgan–Shalen limit , which is the extension of . Let be the universal cover of , and let be an equivariant harmonic map with associated harmonic 1-form . We denote by the associated leaf tree defined by . On the other hand, each contains the free group in its kernel, and therefore the edge stabilizers of all contain free groups. By Skora’s Theorem [Sko96], cannot be dual to a measured foliation on . It follows that and will necessarily have different length functions.
Finally, the length function of also appears in the boundary. Indeed, as in Example 3.2, is a holomorphic quadratic differential on . The length function is associated to the dual tree of the vertical measured foliation of , and this appears in the Thurston boundary of representations of .
Therefore, we have constructed two length functions such that , which implies 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 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 harmonic –form may not be harmonic. In Section 7.3, we discuss a folklore conjecture on the non-existence of harmonic –forms on , and we prove a result about harmonic forms on simply connected closed manifolds. In Section 7.4, we prove a 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 harmonic 1-forms
We prove Theorem 1.3 using Culler–Shalen’s construction of dual trees associated with essential surfaces.
Let be a closed oriented 3-manifold with . Assume is an embedded two-sided surface. We briefly review the construction of the dual tree of as in [Sha02, Sec. 1.4]. Let be the universal cover and the preimage of . There is a canonical 1-dimensional simplicial complex associated with , defined as follows: the vertices are the connected components of , and the edges are the connected components of . An edge is incidental to a vertex if the corresponding component of is contained in the boundary of the corresponding component of . The space is a simplicial tree [Sha02, Sec. 1.4], and the fundamental group acts on by simplicial homeomorphisms. In the following lemma, we show that under the assumptions in Theorem 1.3, the action of on has no fixed points (see also [Sha02, Proposition 1.5.2]).
Lemma 7.1.
Suppose is a rational homology sphere. Suppose is a closed connected embedded surface in that is two-sided, –injective, and does not bound an embedded ball in . Then the action of on has no global fixed points.
Proof.
Since is a rational sphere and is two-sided, must be disconnected. In fact, if is connected, then there is a circle in that intersects transversely at one point, which implies that the homology class of must be non-torsion, contradicting the assumptions. Since is connected, has two connected components.
The action of on has a globally fixed vertex if and only if there is a component of such that is surjective. By the Seifert–van Kampen theorem, this is equivalent to the surjectivity of . Since is –injective, this condition implies is an isomorphism. If is a sphere, then must be a –ball, which contradicts the assumptions. If the genus of is positive, by a standard result in 3-manifold topology (see, for example, [Hat07, Lemma 3.5]), the rank of the kernel of equals the genus of , so cannot be isomorphic.
The action of on has a globally fixed edge if and only if is surjective. By the Seifert–van Kampen theorem and the assumption that is –injective, this implies is an isomorphism for each component of . Hence we get a contradiction by the same argument as before. ∎
Proof of Theorem 1.3.
Let . Since is a simplicial tree, it is complete as a metric space. By Lemma 7.1, the set of projective length functions is non-empty. Hence the construction of the map in Section 6.1 implies that is non-empty. Since is a rational homology sphere, every non-zero harmonic form on must be non-trivial. ∎
7.2. The quotient map to the leaf space
Let be a closed manifold, be a harmonic 1-form over , let be the universal covering map with , then by Theorem 4.4, the leaf space is an –tree. When is 2-dimensional, it is proved in [Wol95, Prop. 3.1] that the quotient map from to is harmonic. It turns out that this statement does not hold generally in higher dimensions. The following is a counterexample when 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 is constructed, such that there is a non-trivial holomorphic section of that locally has the form with a holomorphic function and a holomorphic –form. Therefore, the real part of the square-root of defines a harmonic form on , which we denote by . The universal cover of is the same as . Let be the leaf space of on as given by Theorem 4.4.
One can choose the parameters in the construction so that the leaf space of contains infinitely many points. To explain this, we need to recall the construction of from [BDO11]. Let be a –dimensional abelian variety and let be a smooth hypersurface in . Let be the natural involution on , and assume passes through exactly one fixed point of . Then is defined to be the minimal resolution of . Up to multiplication by constants, there is a unique holomorphic –form on such that . The section is defined as an extension of the push-forward of to .
As a result, the harmonic –form is equal to the push-forward of the real part of on . Identifying with a quotient of by a discrete group of translations, one may choose so that the space is given by a linear equation with coefficients in . In this case, the kernel of the real part of defines a foliation on such that every leaf is closed. If are two closed leaves whose images in are disjoint, are both disjoint from , and both intersect non-trivially, then the images of in are two distinct points. As a result, contains infinitely many distinct points.
Since is compact, by the maximum principle for harmonic maps, every harmonic map from to an –tree must be constant. Therefore, the quotient map from to cannot be harmonic. ∎
7.3. harmonic forms on closed simply connected manifolds
In this subsection, we prove Theorem 1.7, which states that harmonic –forms cannot exist on 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 is a simply connected closed Riemannian manifold (not necessarily in dimension ), and is a harmonic –form on with zero locus . Let be the transverse measure on defined by (see (12)), and let be the corresponding pseudo-metric on (see (13)); we also use to denote the induced metric on the leaf space .
Definition 7.3.
We say that a zero point of is cylindrical, if there exists a local coordinate chart on a neighborhood of such that for some integer , where is a smooth and non-vanishing function, and .
The following theorem is a generalization of Theorem 1.7.
Theorem 7.4.
Suppose is a closed, simply-connected, Riemannian manifold. Then there is no harmonic –form on such that the following conditions hold at the same time.
-
(1)
Every point of is cylindrical.
-
(2)
For every arc in transverse to , where denotes the zero locus of , we have
Remark 7.5.
Theorem 7.4 will be a straightforward consequence of the following lemma.
Lemma 7.6.
Assume is a simply connected (but not necessarily closed) manifold and satisfies the conditions in Theorem 7.4. Let be the projection map from to the leaf tree . Then germs of convex functions on pull back to germs of subharmonic functions on .
Here, a function on an –tree is called convex, if for every geodesic parametrized by arc length and every , we have
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 is regular and harmonic near every non-zero point of , so it satisfies the desired properties on the complement of the zero locus of .
Let be a zero point of . Take a coordinate chart containing such that Definition 7.3 holds. Let be the integer in Condition (1). Let be the open neighborhood of given by in this chart.
Let denote the leaf space of on , then is given by the union of segments with one end point identified. The identified end point is the image of , which we denote by . In the following, we will call each segment in a branch. The branches admit a natural cyclic order.
Let denote the image of in . We will abuse notation and use to also denote the image of in .
By Condition (2), for sufficiently small, the map of each branch of into is an isometric embedding. Therefore, the map from to is a quotient map that identifies pairs or groups of branches. Moreover, Condition (2) also implies that neighboring branches of cannot be identified in .
Let be a convex function defined on a neighborhood of the closure of , we show that pulls back to a subharmonic function near .
We first consider the case that on one of the branches of (which we denote by ), and on .
Let be the pull-back of to . Then the pre-image of in cuts into domains, which we denote by in cyclic order. Then there exists a set , such that
(26) |
In the following, the subscripts will always be interpreted modulo . Since neighboring branches of have distinct images in , we have for every .
By [Ish95], we only need to show that
(27) |
for all , where denotes the outward unit normal vector field on . For each , we decompose into , where , , . It is clear that for all .
We have
Define
where is a unit normal vector field of . By (26) and the fact that is a harmonic function on in the interior of each , we have
Hence (27) follows from the fact that for all .
This proves that the pull-back of is subharmonic near when has the special form given as above. In general, assume is an arbitrary convex function that is defined on a neighborhood of the closure of . Then there exist constants , , and a function having the form above, such that on and . Let denote the pull back of to . Let denote the harmonic function on such that on , let denote the harmonic function on such that on . Then we have
Since the above inequality holds for all and , we conclude that is subharmonic, and the lemma is proved. ∎
Proof of Theorem 7.4.
Assume is a harmonic form satisfying the conditions of Theorem 7.4. Then for every , the distance to is a convex function on . By Lemma 7.6 and the maximum principal, the distance function to must be constant on the image of . Since this holds for all , the map must be constant and hence contains only one point. By Condition (2) in Theorem 7.4, the leaf space contains infinitely many points, which yields a contradiction. ∎
7.4. –convergence for Korevaar–Schoen limits
Theorem 1.1 establishes a bridge between the analytic and algebraic compactifications of the moduli space of flat connections on –manifolds. This connection allows us to prove new results on one side using results from the other. Let be a closed –manifold, let , and let be a sequence of points in that converges to a point in . Let be the corresponding equivariant harmonic maps (see the discussion in Section 6.2.1). Let be the energy of on a fundamental domain. By [DDW98, Thm. 2.2] (see also Section 6.2.1), after rescaling, the limit of defines an equivariant harmonic map to a tree. By [KS97, Thm. 3.9], for every smooth vector field on , the sequence converges weakly to in . We use a result of Parker [Par23a, Thm. 1.3] to show that, in fact, also converges to in on compact subsets of the complement of the zero locus of .
Corollary 7.8.
Let be as above. Let be the open subset of where . Then for every smooth vector field , we have converges in to on compact subsets of .
Proof.
Without loss of generality, assume is a compact set contained in a fundamental domain of , and assume that is invariant. Then and reduce to functions on , and reduces to a vector field on . Assume is the solution to (1) corresponding to , and let be the limit of the sequence in . By Theorem 1.1 and Corollary 6.2, we have .
We abuse notation and also use to denote the image of in . By [Par23a, Thm. 1.3], we know that every subsequence of has a subsequence that converges in on . Hence, by (25), every subsequence of has a subsequence that converges in on . Since weakly converges to , the limit of the subsequence must also be . Therefore, converges to in on . ∎
7.5. Properties of the boundary of the analytic compactification
In this subsection, we discuss several properties of the harmonic 1-forms that actually appear in . 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 harmonic form arises as a limit in Taubes’ compactification, then the leaf space of has similar properties to the leaf spaces of quadratic differentials.
Corollary 7.9.
7.5.2. Canonical measured foliations and the Hubbard–Masur map
We give a geometric interpretation for the restriction of to the boundary , which equals . In Morgan–Shalen’s theory [MS84, MS88a, MS88b], if is in the limit with the corresponding –tree , the transverse equivalence map construction in [MS88b, Sec. I.2] defines a measured lamination from the harmonic map.
By comparison, for each , the corresponding harmonic form defines a (singular) measured foliation on as in Section 4.1. Therefore, the map constructed in Section 6.1 canonically associates a (singular) measured foliation with every Morgan–Shalen limit point.
In fact, the boundary map 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 . As shown in Corollary 6.2, the map generalizes this concept to three dimensions, where harmonic maps to –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 -trees, Proc. London Math. Soc. (3) 55 (1987), no. 3, 571–604. MR 907233
- [Cor88] Kevin Corlette, Flat -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 -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 -trees, Math. Res. Lett. 5 (1998), no. 4, 523–533. MR 1653328
- [DDW00] by same author, On the Morgan-Shalen compactification of the 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 , 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 spinors, J. Differential Geom. 117 (2021), no. 3, 395–449. MR 4255067
- [GO89] David Gabai and Ulrich Oertel, Essential laminations in -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 -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 harmonic 1-forms via 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 -manifolds, Math. Proc. Cambridge Philos. Soc. 119 (1996), no. 1, 73–82. MR 1356159
- [HP24] Siqi He and Greg Parker, -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 -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 , 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 -manifolds, Ann. of Math. (2) 127 (1988), no. 2, 403–456. MR 932305
- [MS88b] by same author, Degenerations of hyperbolic structures. III. Actions of -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 -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 -harmonic spinors on 3-manifolds, arXiv preprint arXiv:2301.06245 (2023).
- [Par24a] Gregory J. Parker, A gauge-theoretic compactification for generic ends of the representation variety on 3-manifolds, in preparation.
- [Par24b] Gregory J Parker, Gluing -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 generalizations of the 4-dimensional anti-self dual equations, arXiv preprint arXiv:1307.6447 (2013).
- [Tau13b] by same author, connections on 3-manifolds with bounds on curvature, Camb. J. Math. 1 (2013), no. 2, 239–397. MR 3272050
- [Tau14] by same author, The zero loci of /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 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 eigenfunctions for the Laplacian on , J. Differential Geom. 128 (2024), no. 1, 379–462. MR 4773187
- [Wol95] Michael Wolf, Harmonic maps from surfaces to -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 harmonic spinors in dimension 4, Comm. Anal. Geom. 30 (2022), no. 7, 1633–1681. MR 4596625