On the equidistribution of closed geodesics and geodesic nets
Abstract.
We show that given a closed -manifold , for a Baire-generic set of Riemannian metrics on there exists a sequence of closed geodesics that are equidistributed in if ; and an equidistributed sequence of embedded stationary geodesic nets if . One of the main tools that we use is the Weyl Law for the volume spectrum for -cycles, proved in [14] for and in [11] for . We show that our proof of the equidistribution of stationary geodesic nets can be generalized for any dimension provided the Weyl Law for -cycles in -manifolds holds.
1. Introduction
Marques, Neves and Song proved in [18] that for a generic set of Riemannian metrics in a closed manifold , there exists a sequence of closed, embedded, connected minimal hypersurfaces which is equidistributed in . In this paper, we study the equidistribution of closed geodesics and stationary geodesic nets (which are -dimensional analogs of minimal hypersurfaces) on a Riemannian manifold , . We prove the following two results, for dimensions and of the ambient manifold respectively:
Theorem 1.1.
Let be a closed -manifold. For a Baire-generic set of Riemannian metrics on , there exists a set of closed geodesics that is equidistributed in . Specifically, for every in the generic set, there exists a sequence of closed geodesics in , such that for every function we have
Theorem 1.2.
Let be a closed -manifold. For a Baire-generic set of Riemannian metrics on , there exists a set of connected embedded stationary geodesic nets that is equidistributed in . Specifically, for every in the generic set , there exists a sequence of connected embedded stationary geodesic nets in , such that for every function we have
Remark 1.3.
We have an equivalent notion of equidistribution for a sequence of closed geodesics or geodesic nets: we say that is equidistributed in if for every open subset it holds
Theorem 1.2 is, as far as the authors know, the first result on equidistribution of -stationary varifolds in Riemannian -manifolds for (i.e. in codimension greater than ). Regarding Theorem 1.1, similar equidistribution results for closed geodesics have been proved for compact hyperbolic manifolds in [4] in 1972 and for compact surfaces with constant negative curvature in [22] in 1985. More recently, those results were extended to non-compact manifolds with negative curvature in [23] and to surfaces without conjugate points in [8]. The four previous works have in common that they approach the problem from the dynamical systems point of view. In the present paper, we approach it using Almgren-Pitts min-max theory (as it was done in [18] for minimal hypersurfaces). Additionally, Theorem 1.1 is the first equidistribution result for closed geodesics on closed surfaces that is proved for generic metrics, without any restriction regarding the curvature of the metric or the presence of conjugate points.
Our proof is inspired by the ideas in [18]. There are two key results used in [18] to prove equidistribution of minimal hypersurfaces for generic metrics: the Bumpy Metrics Theorem of Brian White [25] and the Weyl Law for the Volume Spectrum proved by Liokumovich, Marques and Neves in [14]: given a compact Riemannian manifold with (possibly with boundary), we have
for some constant . Here, given we denote by the -dimensional -width of with respect to the metric (for background on this, see [10], [15], [17] [13]). It was conjectured by Gromov (see [9, section 8.4]) that the Weyl law can be extended to other dimensions and codimensions. In this work, we are interested in the case of -dimensional cycles. The following is the most general version of the Weyl law we could expect for -cycles.
Conjecture 1.4.
Let be a closed -dimensional manifold, . Then there exists a constant such that
So far, Conjecture 1.4 has been proved for as a particular case of [14] and recently for by Guth and Liokumovich in their work [11]. In this article, we use those two versions of the Weyl law to prove Theorem 1.1 and Theorem 1.2; and we also use the Structure Theorem for Stationary Geodesic Networks proved by Staffa in [24] and the Structure Theorem of White ([25]) for the case of embedded closed geodesics. The work of Chodosh and Mantoulidis in [7] is used to upgrade the equidistribution result for stationary geodesic networks to closed geodesics in dimension . The only obstruction to extend our proof of the equidistribution of stationary geodesic nets to arbitrary dimensions of the ambient manifold is that Conjecture 1.4 has not been proved yet if . As all the rest of our argument works for any dimension , what we do is to prove the following result and then show that it implies Theorem 1.1 and Theorem 1.2.
Theorem 1.5.
Let , be a closed manifold. Assume that the Weyl law for -cycles in -manifolds holds. Then for a Baire-generic set of Riemannian metrics on , there exists a set of connected embedded stationary geodesic nets that is equidistributed in . Specifically, for every in the generic set , there exists a sequence of connected embedded stationary geodesic nets in , such that for every function we have
In order to simplify the exposition, we consider integrals of functions instead of the more general traces of -tensors discussed in [18]. Next we proceed to describe the intuition behind the proof, the technical issues which appear when one tries to carry on that intuition and how to sort them.
Let be a Riemannian metric on . We want to do a very small perturbation of to obtain a new metric which admits a sequence of equidistributed stationary geodesic networks. Let be a smooth function. Consider a conformal perturbation (for some small) defined as
By [15, Lemma 3.4] the normalized -widths are uniformly locally Lipschitz. This combined with the Weyl Law (recall that we assume it holds) implies that the sequence of functions
converges uniformly to the constant . Considering
we have that converges uniformly to the constant . On the other hand, Almgren showed that there is a correspondence between -widths and the volumes of stationary varifolds (see [1], [2], [5], [19], [20], [21]) such that for each and there exists a (possibly non unique) stationary geodesic network such that
(1) |
Assume that the ’s can be chosen so that all of them are parametrized by the same graph and the maps , are differentiable (this is a very strong assumption and doesn’t necessarily hold, as the map may not be differentiable; a counterexample is shown below). In that case we can differentiate and obtain
As converges uniformly to a constant, we could expect that the sequence converges to for some values of . If that was the case, the sequence would verify the equidistribution formula for the function with respect to the metric . Nevertheless, this does not have to be true, because of two reasons. The first one is that the uniform convergence of a sequence of functions to a constant does not imply convergence of the derivatives to at any point. Indeed, we can construct a sequence of zigzag functions which converges uniformly to but does not converge to for any . The second one is that the differentiability of could fail, a counterexample is shown in the next paragraph. And even if that reasoning was true and such existed, the sequence constructed would only give an equidistribution formula for the function (which is used to construct the sequence) instead of for all functions at the same time; and with respect to a metric which could also vary with .
An example when is not differentiable is the following. Let us consider a dumbbell metric on obtained by constructing a connected sum of two identical round -spheres and of radius by a thin neck. Define a -parameter family of metrics such that along , along (interpolating along the neck so that it is still very thin). It is clear than for , the -width is realized by a great circle in with length , and for it is realized by a great circle in of length . Therefore
and hence it is not differentiable at .
To fix the previous issue (differentiability of ), we prove Proposition 4.1 which is a version for stationary geodesic networks of [18, Lemma 2]. Regarding the convergence of to for certain values of , we use Lemma 4.6 which is exactly [18, Lemma 3]. To obtain a sequence of stationary geodesic networks that verifies the equidistribution formula for all functions (and not only for a particular one as above), we carry on a construction described in Section 5 using certain stationary geodesic nets which realize the -widths in a similar way as the ’s above. The key idea here is that the integral of any function over can be approximated by Riemann sums along small regions with piecewise smooth boundary where is almost constant. Therefore, if we have an equidistribution formula for the characteristic functions of those regions (or some suitable smooth approximations), then we will be able to deduce it for an arbitrary . The advantage of doing this is that we reduce the problem to a countable family of functions. This argument is also inspired by [18].
The paper is structured as follows. In Section 2, we introduce the set up and necessary preliminaries. In Section 3, we define the Jacobi Operator along a stationary geodesic net and show that it has all the nice properties that an elliptic operator has (mainly, admitting an orthonormal basis of eigenfunctions and therefore having a min-max characterization for its eigenvalues). This is crucial to prove Proposition 4.5. In Section 4, the technical propositions necessary to prove Theorem 1.5 are discussed. In Section 5 we prove Theorem 1.5 and get Theorem 1.2 as a corollary using the Weyl law for cycles in manifolds proved in [11]. In Section 6, we use the Weyl law from [14] and the proof of Theorem 1.5 combined with the work of Chodosh and Mantoulidis in [7] (where it is proved that the -widths on a surface are realized by finite unions of closed geodesics) to prove Theorem 1.1.
Remark 1.6.
Rohil Prasad pointed out that an alternative proof of Theorem 1.1 could be obtained using the methods of Irie in [12]. Given a closed Riemannian -manifold , its unit cotangent bundle is a closed -manifold equipped with a natural contact structure induced by the contact form which is the restriction of the Liouville form on to . It is a well known fact that the Reeb vector field associated to generates the geodesic flow of . Additionally, given a function , the Riemannian metric corresponds to the conformal perturbation of the contact form in (here is the projection map); and both and are compatible with the same contact structure on . Thus one would like to apply [12, Corollary 1.4] to with the contact structure induced by . However, that result is about generic perturbations of the contact form of the type , where and we only want to consider perturbations which are liftings to of maps so some work should be done here in order to apply Irie’s result in our setting. This issue was pointed out in [6, Remark 2.3], where a similar problem is studied for Finsler metrics and a solution is given for that class of metrics. Additionally, Irie’s theorem would give us an equidistributed sequence for a generic conformal perturbation of each metric . This immediately implies that for a dense set of Riemannian metrics such an equidistribution result holds, but some additional arguments are needed to prove it for a Baire-generic metric. It is important to point out that the result in [12] uses the ideas of [18] but in the different setting of contact geometry, applying results of Embedded Contact Homology with the purpose of finding closed orbits of the Reeb vector field; while in [18] Almgren-Pitts theory is used to find closed minimal surfaces.
Aknowledgements. We are very grateful to Yevgeny Liokumovich for suggesting this problem and for his valuable guidance and support while we were working on it. We are deeply thankful to the referee for reading the article very carefully and for their great feedback, particularly for pointing out a gap in the proof of Proposition 5.1 which was filled by proving Proposition 4.5. We also want to thank Rohil Prasad for his suggestion about using the approach mentioned in Remark 1.6 to get an alternative proof of Theorem 1.1 and for the valuable discussion derived from there. We are thankful to Wenkui Du as well, because of his comments on the preliminary version of this work. Bruno Staffa was partially supported by Discovery Grant RGPIN-2019-06912.
2. Preliminaries
Definition 2.1 (Weighted Multigraph).
A weighted multigraph is a graph consisting of a set of edges , a set of vertices and a multiplicity assigned to each edge . A weighted multigraph is good if it is connected and either it is a closed loop with mutiplicity or each vertex has at least three different incoming edges (here loop edges at count twice as an incoming edge at , see [24] for a more detailed discussion). In the later case we say is good*.
Definition 2.2.
Given a weighted multigraph , we identify each edge with the interval and we denote the map sending to the vertex under the identification .
Definition 2.3 (-net).
A -net on is a continuous map which is a -immersion restricted to the edges of . We will denote the space of -nets on . It has a natural Banach manifold structure as a subspace of (see [24]).
Definition 2.4.
We say that two -nets and are equivalent if for every edge of the map is a reparametrization of fixing the endpoints. This defines an equivalence relation in . We denote the quotient space. Given we will often denote a representative of the equivalence class also by , and regard different representatives as different parametrizations of the geometric object .
Notation 2.5.
Given a -net and an edge , we denote the restriction of to . We also define and .
Notation 2.6.
Given , let us denote the set of Riemannian metrics on .
Definition 2.7.
Let and let be a continuous function defined in . Given a metric we define
Observe that the right hand side is independent of the parametrization we choose and therefore is well defined.
Definition 2.8 (-Length).
Given and , we define the -length of by
Definition 2.9 (Stationary Geodesic Network).
We say that is a stationary geodesic network with respect to a metric () if it is a critical point of the length functional . This means that given any smooth one parameter family with we have
Assuming that the edges of are parametrized by constant speed, if (here we regard ) then
(2) |
where and
Equation (2) is called the First Variation Formula and was computed in [24, Section 1]. It implies that is stationary with respect to if and only if each edge is mapped to a geodesic segment in and the stability condition at the vertices is verified. The latter means that for each , the sum of the inward pointing unit tangent vectors to each edge at is .
Definition 2.10.
We say that is a stationary geodesic network if every representative of is a stationary geodesic network.
Definition 2.11.
We denote the space of continuous vector fields along whose restriction to each edge is of class .
Remark 2.12.
If , and is stationary with respect to then by the regularity of the solutions of an ODE, is of class for every . This is why we only ask regularity to -nets and vector fields along them.
Assume is a stationary geodesic net with respect to a metric with (so that the Riemann curvature tensor is of class ). Let be a smooth -parameter family of -nets with . Let and . We define the Hessian of the length functional at as the bilinear form
In [24, Section 2] it was shown that is well defined (i.e. it does not depend on which two parameter variation with directional derivatives and we choose) and in fact it holds
(3) |
where
being the restriction of the vector field to the edge and the component of orthogonal to . Observe that is (up to a positive constant) the Jacobi operator along . Equation (3) is the Second Variation Formula.
Definition 2.13.
We say that a vector field is Jacobi if it is a null vector of , i.e. if for every . By the Second Variation Formula, is Jacobi along if and only if for every and for every .
Definition 2.14.
A vector field is said to be parallel if is parallel along for every .
Remark 2.15.
Observe that every parallel vector field along is Jacobi.
Definition 2.16.
A stationary geodesic net is nondegenerate if every Jacobi field along is parallel.
Remark 2.17.
In [15, Lemma 2.5], it was shown that every stationary geodesic network with respect to a metric can be represented by a map , where is the finite union of the good weighted multigraphs and is an embedded stationary geodesic network for each (moreover, the map is a topological embedding).
Definition 2.18.
Given a stationary geodesic network , we say that its connected components are nondegenerate if
-
(1)
We can express as a disjoint union of good weighted multigraphs.
-
(2)
is an embedded nondegenerate stationary geodesic network for every .
Definition 2.19.
An almost embedded closed geodesic in a Riemannian manifold is a map such that
-
(1)
is geodesic (i.e. for every ).
-
(2)
is an immersion (i.e. for every ).
-
(3)
All self-intersections of are transverse, which means that for every such that and , the velocities and are not colinear.
This terminology is inspired in [26, Definition 2.2], where Brian White extends his Bumpy Metrics Theorem to almost embedded minimal surfaces.
Notation 2.20.
Given a symmetric -tensor , a metric , a stationary geodesic network on and , we denote
which is the trace of the tensor along with respect to the metric .
Definition 2.21 (Average integral along ).
Let be a weighted multigraph. Given , a metric and a continuous function defined in , we define the average integral of with respect to metric as
3. The Jacobi Operator
In this section we will study some properties of the Jacobi operator of an embedded stationary geodesic network , where is a good weighted multigraph and , . We will focus on the case when is good* (i.e. every vertex has at least three different incoming edges), because when is a loop with multiplicity what we get is the Jacobi operator along an embedded closed geodesic acting on normal vector fields, which is known to be elliptic; and hence it has all the nice properties that we will describe below. We first introduce some notation. Let
Observe that as is good*, if then for every . Denote
By the second variation formula (3), we can define the Jacobi operator as
(4) |
We know that each is Jacobi (i.e. it verifies ). We want to construct a complement of in , and show that when we restrict to that complement it behaves like an elliptic operator (this complement will play the role of the space of normal Jacobi fields along a minimal submanifold in the smooth case, when it is known that the stability operator is elliptic).
To do this, we will need to define a finite dimensional subspace such that the evaluation map , is a linear isomorphism. This can be done by taking a basis of for each , and for each pair with and defining a vector field such that and for every . Then we can define . Of course the choice of is not canonical, but we fix one choice and work with it for the rest of the section (it will be deduced from the arguments below that the results that we prove hold independently of the choice of ). It is clear that
Denote which is a complement of the space of parallel vector fields along . Same as in the theory of elliptic operators, we can extend the Jacobi operator to Sobolev spaces of vector fields along once we have a suitable definition of them. Denote
Notice that is the -version of and will be the domain of the Jacobi operator we will work with (as that operator vanishes on ). The previous spaces are defined in analogy with the spaces of , and normal vector fields along a smooth closed submanifold which appear when studying the ellipticity of its Jacobi operator. The space is a Hilbert space with the inner product
and we have a monomorphism with dense image given by
which allow us to write the Hessian as
where is the inner product in . Here we considered given by (4) which is a bounded linear operator. As in the smooth case, we can also regard as an unbounded operator whose domain is the dense linear subspace . We would therefore expect that for sufficiently big the operator has a compact inverse, and from that get an orthonormal basis of consisting of eigenvectors of . This indeed holds, as it is shown in the following proposition.
Proposition 3.1.
For every , the operator defined as is Fredholm of index . The spectrum of consists of an increasing sequence of eigenvalues with (i.e. has nontrivial kernel if and only if and has a continuous inverse otherwise). In addition, there exists sequence in such that is an orthonormal basis of and for each . Therefore, we have the following min-max characterization of the eigenvalues of
where the minimum is taken over all -dimensional subspaces .
Proof.
Let . Then if and ,
where is (a constant multiple of) the Jacobi operator along given by . We know that each is elliptic, and therefore is Fredholm of index for every . This implies that the product operator , verifies that is Fredholm of index for every . Thus the fact that is always Fredholm of index can be deduced from the following lemma:
Lemma 3.2.
Let , , , be Banach spaces with . Let be a continuous linear map, and write with . Assume is Fredholm of index . Then is Fredholm of index .
Proof of Lemma 3.2.
Let be the operator . Because is Fredholm of index and , we see that is also Fredholm of index . As with compact because of the finite dimensionality of and , by [16, Theorem 12-5.13] we deduce that is also Fredholm of index . ∎
Now we are going to show that the quadratic form is bounded from below. We know
Denote by the form . is symmetric because so are and for each . If we endow with the inner product
then as , we can see that there exists some constant such that
(5) |
for every . But then as vanishes on , by its bilinearity and symmetry we can see that in fact (5) is valid for every .
On the other hand, using that each is elliptic, for each there exists such that
(6) |
Thus if and , from (5) and (6) we deduce
which considering that implies that for every it holds
and in particular if implies that is a monomorphism. Because we also know that these operators are Fredholm of index , by the Open Mapping Theorem we conclude that is a continuous linear isomorphism for every .
Fix . We will now show that is compact. Let be a bounded sequence in and define with and . As is bounded, is a bounded sequence in . Therefore, for each the sequence of normal vector fields along is bounded in and therefore in . Hence, by the Rellich-Kondrachov Compactness Theorem we can find a subsequence such that converges in for every . On the other hand, using that is finite dimensional, we can extract a further subsequence to have the additional property that converges in . This implies that the sequence of general term converges in , and this completes the proof that is compact.
The symmetry of implies that is self-adjoint, which together with its compactness implies the existence of an orthonormal basis of such that for some decreasing sequence (because by our choice of , ). But we claim that is an eigenvector of of eigenvalue if and only if for some such that . This is because if and only if there exists with which verifies any of the the following equivalent conditions:
From the previous, we conclude that if then , and . This implies the min-max theorem for the eigenvalues holds for , which completes the proof. ∎
4. Some auxiliary results
Proposition 4.1.
Let be a smooth embedding, , . If , there exists an arbitrarily small perturbation in the topology such that there is a full measure subset with the following properties: for any and any , the function is differentiable at , and there exists a (possibly disconnected) weighted multigraph and a stationary geodesic network such that the following two conditions hold
-
(1)
.
-
(2)
.
To prove the proposition, we will need to have a condition for a sequence of embedded stationary geodesic nets converging to some that guarantees that is also embedded. The condition we will work with can be expressed as a collection of lower and upper bounds of certain functionals defined for pairs where is stationary with respect to . We proceed to describe those functionals.
The first one is
A lower bound for this functional will imply that the limit net is an immersion along each edge.
Then we have a family of functionals defined for each pair such that (see Section 2 for the notation) as follows
Notice that is the unit inward tangent vector to at along , (and observe that in case is a loop at , there are two inward tangent vectors to along at represented by the pairs and ). The condition for some and for every possible choice with implies that the limit has the property that given , there exists an open neighborhood of in such that is a homeomorphism. Explicitly,
where , is a continuous choice of the injectivity radius for each Riemannian metric . This is because if we consider as a graph obtained by gluing at one edge for each pair such that , this graph is mapped by into a geodesic ball centered at of radius and the image of each incoming edge at has a different inward tangent vector at .
To ensure injectivity along the edges, we define for each edge a function
In case , the distance between two points is measured with respect of the length of .
To ensure that the images of different edges under do not overlap, we define for each pair , a function as
Let us fix an auxiliary embedding and identify from now on our manifold with the submanifold . Given a multigraph and a continuous map which is when restricted to each edge, we can consider
where given a collection of continuous functions along the edges of , we define
being the Euclidean norm in . We have the following compactness result.
Lemma 4.2.
Let be a sequence of Riemannian metrics converging to some metric . Let be a sequence of stationary geodesic networks. Assume for some . Then there exists a subsequence and such that in and is stationary.
Proof.
The Arzela-Ascoli Theorem gives a subsequence in . The fact that is stationary with respect to comes from the continuity of the operator defined in [24] (which plays the role of the mean curvature operator on minimal surfaces) which vanishes in a pair if and only if is stationary with respect to . ∎
We will also need the following two lemmas.
Lemma 4.3.
Let be a function. Then there exists and a basis of such that for all .
Proof.
Observe that and therefore where given we denote the subspace spanned by . If did not contain a basis of for every , would a proper subspace for every . Therefore, would be a countable union of closed subspaces with empty interior, which leads to a contradiction due to the Baire Category Theorem. ∎
Lemma 4.4.
Let and be smooth maps. Assume that is stationary with respect to for every . Then for every and every
Proof.
Using that the length functional is a smooth function and the chain rule, we get
The second term in the penultimate equation vanishes because is stationary with respect to . Hence
∎
Proof of Proposition 4.1.
Notice that it suffices to show that for each , there exists a full measure subset where (1) and (2) hold, because in that case will have the desired property. Therefore we will assume is fixed.
Let be a smooth embedding. Let be a sequence enumerating the countable collection of all good weighted multigraphs. For each , let be the space of pairs where , is an embedded stationary geodesic net and denotes its class modulo reparametrization as defined in [24] for connected multigraphs with at least three incoming edges at each vertex and in [25] for embedded closed geodesics. By the structure theorems proved in [24] and [25], each is a second countable Banach manifold and the projection map mapping is Fredholm of index . A pair is a critical point of if and only if admits a nontrivial Jacobi field with respect to the metric .
By Smale’s transversality theorem, we can perturb slightly in the topology to a embedding which is transversal to for every . Transversality implies that is an -dimensional embedded submanifold of for every . Let . Let be the set of regular values of , which is a set of full measure by Sard’s theorem. Let be the set of points for which the Lipschitz function is differentiable. Observe that has full measure by Rademacher’s theorem. Therefore, is a full measure subset of . Notice that by transversality, if then is a bumpy metric, i.e. all embedded stationary geodesic nets with respect to and with domain a good weighted multigraph are nondegenerate; and also the map is differentiable at .
Given a weighted multigraph whose connected components are good and a natural number , we define as the set of all such that there exists a stationary geodesic network verifying
-
(1)
is an embedding for each .
-
(2)
for every .
-
(3)
for every .
-
(4)
for every and every pair in such that .
-
(5)
for every , and .
-
(6)
for every , and .
-
(7)
.
where denotes the set of edges of . Observe that because of (1) and Remark 2.17. We claim that each is closed.
Indeed, suppose we have a sequence converging to some . Let be the stationary geodesic network corresponding to and verifying properties (1) to (7) above. By property (2) and Lemma (4.2), passing to a subsequence we have that if then there exists such that in and is stationary with respect to for each . Observe also that if
Properties (2) to (6) are preserved when we take the limit of the sequence , so it suffices to show that is embedded for each . Fix such . Properties (3), (4) and (5) imply that is injective along the edges and property (6) combined with property (4) imply that the images of different edges do not intersect (except at the common vertices).
As each is closed, they are measurable and therefore so are the sets (whose union is ). Let be the set of points where the Lebesgue density of at is . By the Lebesgue Differentiation Theorem, has Lebesgue measure for each pair . Let us define , observe that as has measure , has full measure.
Fix . Let be such that . As the density of at is , given with we can find a sequence such that and . Denoting , using that is a Lipschitz function we can see that
(7) |
As , for each there exists a stationary geodesic network with respect to such that
and properties (1) to (6) above hold. By the reasoning used to prove that the are closed, we can construct a stationary geodesic net which is embedded when restricted to each connected component of , is the limit of (a subsequence of) the ’s in the topology and realizes the width . Hence from (7) we get
As is an embedded stationary geodesic net with respect to for each and is bumpy, is a diffeomorphism from a neighborhood of to a neighborhood of . Denote its inverse. As there exists such that and for every , we deduce that for each and each . Let us define as where . Thus by Lemma 4.4
Where is the one constructed before. Observe that depends on and that the previous formula holds for each , . Notice that each is a stationary geodesic network with respect to , and as is bumpy there are countably many possible , say . This induces a map defined as and if then where . By Lemma 4.3 we can obtain and a basis of with the property for every . Therefore if we set , is still a basis and by definition for every . By linearity of directional derivatives, denoting we deduce that
for every unit , which completes the proof. ∎
Proposition 4.5.
Let be a closed manifold and let be a Riemannian metric on , . Let be a finite collection collection of connected, embedded stationary geodesic networks on whose domains are good weighted multigraphs and let be an open neighborhood of . Then there exists such that are non-degenerate stationary geodesic nets with respect to .
Proof.
Following [18, Lemma 4], we will consider conformal perturbations of the metric of the form . Let us denote , (where ) and the set of edges of . Notice that is a stationary geodesic network whose edges may overlap, even non-transversally. Given , let be the set of interior points of which are not points of transverse intersection with any other edge . We define a finite poset
which is the collection of finite non-empty intersections of sets in , with the order given by the inclusion. Denote by the set of minimal elements in . Observe that if are two different elements of then they are disjoint. Given , write in the unique way such that for every . Pick for every , and let be such that the geodesic balls verify
-
•
if .
-
•
if .
-
•
for every .
-
•
There exists a diffeomorphism such that for each .
Denote . Observe that for each there exists at least one such that . Choose such an for each and denote and . We can now proceed to define the function which will induce the one-parameter family of metrics mentioned before.
For each , let be a smooth function with , and in . Let be given in local coordinates under the chart by . We define . An easy computation shows that vanishes along and in local coordinates if , and ; and if . In particular, if for some then if and only if for some (or equivalently, for every) where .
Therefore we know that and vanish along each . Hence by [3, Theorem 1.159], the are still stationary with respect to . Fix with set of vertices and set of edges . We assume that is good* (i.e. every vertex has at least different incoming edges), the case when is an embedded closed geodesic can be handled with the same method using the ellipticity of its Jacobi operator. As discussed in Section 3, the stability operator of with respect to is the map given by
where
Let us compute which is the change in the Jacobi operator along when we switch from the metric to . We will denote the operator corresponding to . Using [3, Theorem 1.159] and the fact that along , we can see that for all , and that
where the covariant derivatives and the curvature tensor are taken with respect to the metric , and at each point , is the linear transformation such that the Hessian of at is given by .
We know from Section 3 that each admits a non-decreasing sequence of eigenvalues which are characterized by
where the infimum is taken over all -dimensional subspaces of . Also, the map is continuous; therefore varies continuously with for every . We will use these facts to show that for sufficiently small values of , is not an eigenvalue of .
Let be the unique natural number such that (here ). Denote the sum of the eigenspaces corresponding to . Let , . Then we have
because for every . Suppose there is equality for some . Then the two inequalities should be equalities. From the first one we deduce that is Jacobi along for the metric , and thus it verifies for every . From the second one, by considering the values of for which , we see that is a null vector of along and therefore on ; and as it satisfies the Jacobi equation this implies for every . Thus must be parallel and hence as does not contain non-trivial parallel vector fields. But this is a contradiction because we chose . Hence we just proved that
for every . As is finite dimensional and is invariant under rescaling of , the compactness of the unit ball in implies that there exists such that
for every . By the min-max characterization of the eigenvalues for , we see that . If we also choose sufficiently small so that , we get that for , is nondegenerate with respect to . Taking such that and defining we get the desired result.
∎
Lemma 4.6.
Given and , there exists depending on and such that the following is true: for any Lipschitz function satisfying
for every , and for any subset of of full measure, there exist sequences of points contained in and converging to a common limit such that:
-
•
f is differentiable at each ,
-
•
the gradients converge to vectors with
Proof.
See [18, Lemma 3]. ∎
5. Proof of the Main Theorem
Fix an -dimensional closed manifold . We are going to consider several choices and constructions over . Let be a Riemannian metric on . Let be a positive constant such that , where is the injectivity radius of . Let be an integer and be disjoint domains in , with piecewise smooth boundary, such that the union of their closures covers . Let be some open neighbourhoods of respectively with the property that each of them is contained in a geodesic ball of radius of . Denote the space of all Riemannian metrics on . For each , we define a smooth function , such that
Consider also the partition of unity . We denote
the set of all possible choices as above with . Notice that is non-empty, as we can always find a sufficiently fine triangulation of . We claim that the following property holds:
Proposition 5.1.
Assume that the Weyl law for -cycles in -manifolds holds as stated in Conjecture 1.4. Then for any metric , for every , and any choice of
there is a metric arbitrarily close to in the topology such that the following holds: there are stationary geodesic networks with respect to whose connected components are nondegenerate (according to Definition 2.18) and coefficients with satisfying
(8) |
for every .
In the proof, we will need to measure the distance between two rescaled functions. In order to do that, we introduce the following definition.
Definition 5.2.
We say that two functions are -close if
where are given by and .
Remark 5.3.
Observe that is differentiable at if and only if is differentiable at and in that case .
Proof of Proposition 5.1.
Let , and . Fix . Let be a neighborhood of . Choose sufficiently small and sufficiently large so that if satisfies , then . Let be a positive real number (which we will have to shrink later in the argument). Our goal is to show that there exists such that and (8) holds for some stationary geodesic nets (whose connected components are nondegenerate) with respect to and some coefficients .
Consider the following -parameter family of metrics. For a , we define
At , for each , we have
As goes to zero, we have the following expansion
(9) |
where if , where is a constant which depends only on (this can be checked by computing the second order partial derivatives of and using the fact that as ). Following [18] we can define the following function
Because of (9), for every ; where depends only on (as and other constants to be defined later).
By the previous, is -close to in if (see Definition 5.2). Let be such that is an embedding and for every . We can slightly perturb in the topology to another embedding applying Proposition 4.1. We can assume and for every and . Consider the function
By the properties of , there exists such that is -close to the constant function equal to on .
Now we will use the assumption that the Weyl law for -cycles in -manifolds holds, which means that
(10) |
The normalized -widths are uniformly Lipschitz continuous on by [15, Lemma 3.4]. Hence, by (10) the sequence of functions converges uniformly to the function . This implies that for the previously defined , there exists such that implies
and hence
for every . The previous means that is -close to in and therefore as is -close to , by triangle inequality we have that
is -close to if , for some .
On the other hand, by our choice of using Proposition 4.1, there exists a full measure subset such that for each and the map is differentiable at and there exists a stationary geodesic net with respect to so that
-
(1)
-
(2)
.
Define as . We know that . Now we want to use Lemma 4.6. In order to do that we will need to impose more restrictions on . Let . Let be the one depending on and according to Lemma 4.6. Choose small enough so that , and . Observe that this allows us to define and with all the properties in the construction above. Then we have
for every . As is Lipschitz, we can apply Lemma 4.6 to and the full measure subset . After passing to by rescaling and using Remark 5.3, we get sequences of points contained in and converging to a common limit such that:
-
(1)
is differentiable at each .
-
(2)
The gradients converge to vectors with
Let be such that and . Then if is sufficiently large,
and hence
for every . But using the definition of and denoting ,
As the lengths of the ’s are uniformly bounded, passing to a subsequence we can obtain stationary geodesic networks with respect to verifying
(11) |
in the varifold topology for every . Hence from the previous,
for every . Using that , and for every and ; and the fact that , we can see that there exists a constant such that
By definition of , thus
(12) |
Combining (12) with the fact that ,
(13) |
But we know that for every , so
if for some , because of the Weyl law and the fact . Hence from (13),
for some constant depending only on . Let us take and .
Notice that . Let us represent each as a map where each connected component of the weighted multigraph is good and the restrictions of to those connected components are embedded (here we are using Remark 2.17). The metric has all the properties required by Proposition 5.1 except that the components of the ’s may not be non-degenerate and may not be (in principle they are only ). Using Proposition 4.5, we can change for another metric which still verifies , and has the property that are non-degenerate stationary geodesic nets with respect to . If on top of that we apply the Implicit Function Theorem (see [15, Lemma 2.6]), we can find a metric close enough to in the topology so that which admits stationary geodesic networks whose connected components are nondegenerate and verify
for every . This completes the proof.
∎
Now we will show that Proposition 5.1 implies Theorem 1.5. Given , , and we will denote the set of all metrics such that (computed with respect to ) and there exist stationary geodesic networks with respect to whose connected components are nondegenerate (according to Definition 2.18) and coefficients with such that (8) holds for every . By the Implicit Function Theorem, is open (see [15, Lemma 2.6]). Therefore given and the set
is open and by Proposition 5.1 it is also dense in . Define
which is a generic subset of in the Baire sense. We are going to prove that if then there exists a sequence of equidistributed stationary geodesic networks with respect to .
Fix . By definition, given there exists such that for some . Therefore, belongs to a neighborhood of in the topology; and there exist , stationary geodesic networks with respect to and coefficients with satisfying
(14) |
for every . Let . We want to obtain a formula analogous to the previous one but replacing by , which will imply the following proposition.
Proposition 5.4.
Let . For each , there exists depending on , integers and stationary geodesic networks such that
for every , where is a constant depending only on and the metric .
Proof.
Given , consider as above and such that . Define , stationary geodesic networks with respect to and coefficients such that (14) holds. Taking into account, let us choose points with for each . The idea will be to approximate the integral of by the integral of the function . First of all, by using (14) we can see that
(15) |
where depends only on (and not on , or ). On the other hand, given
We used the Mean Value Theorem and the fact that and for every . Combining this and (15) we get
(16) |
where depends only on and . Let us choose integers such that
Then it holds
and hence by (16) and triangle inequality we get
where depends only on and . On the other hand,
because . Hence
for a constant depending only on and , as desired. ∎
Given , using Proposition 5.4 we can find a sequence of finite lists of connected embedded stationary geodesic nets with respect to satisfying the following: given , if we denote and , then
(17) |
where and is a constant depending only on . The lists are obtained from the lists and the coefficients from Proposition 5.4 by decomposing each as a union of embedded stationary geodesic networks whose domain is a good weighted multigraph (see Remark 2.17) and listing each of them times. From the ’s and the ’s, we want to construct two sequences such that
-
•
For all , there exist integers (chosen independently of ) with and ,
-
•
It holds
This can be done as in [18, p. 437-439] and gives us a sequence of connected embedded stationary geodesic networks with respect to (defined as ), which is constructed independently of the constant . It holds
for every . This gives us the desired equidistribution result and completes the proof of Theorem 1.5.
6. Equidistribution of almost embedded closed geodesics in 2-manifolds
In this section we show that the proof of Theorem 1.5 combined with the work of Chodosh and Mantoulidis in [7] (where they show that the -widths on a surface are realized by collections of almost embedded closed geodesics) imply Theorem 1.1. The strategy to show this result will be to follow the proof of Theorem 1.5 replacing “embedded stationary geodesic network” by “almost embedded closed geodesic”. The main change needed in the proof is the following version of Proposition 4.1:
Proposition 6.1.
Let be a closed -manifold. Let be a smooth embedding, . If , there exists an arbitrarily small perturbation in the topology such that there is a full measure subset with the following property: for any and any , the function is differentiable at and there exist almost embedded closed geodesics such that the following two conditions hold
-
(1)
.
-
(2)
.
Proof.
We are going to adapt the proof of Proposition 4.1 by introducing some necessary changes. A priori, the easiest way to do this seems to be substituting “stationary geodesic network” by “finite union of almost embedded closed geodesics” everywhere and use the Bumpy metrics theorem for almost embedded minimal submanifolds proved by Brian White in [26]. Nevertheless, there is not an easy condition (analog to conditions (1) to (7) in the proof of Proposition 4.1) that we can impose on a sequence of almost embedded closed geodesics to converge to another almost embedded closed geodesic without classifying them by their self-intersections and the angles formed there. Therefore, what we will do is to treat the almost embedded closed geodesics as a certain class of stationary geodesic networks, and then proceed as with Proposition 4.1.
To each almost embedded closed geodesic we can associate a connected graph where is the equivalence relation if and only if . This induces a map defined as . Observe that the as the self-intersections of are transverse, the vertices of are mapped precisely to those self-intersections and the map is injective. Moreover, is a good multigraph and is an embedded stationary geodesic network. We replace the set which in the proof of Proposition 4.1 is the set of all good connected multigraphs by the countable set of pairs where is a good multigraph which can be obtained as from an almost embedded closed geodesic with respect to some metric as before and is the set of pairs such that and (in other words, contains the necessary information to reparametrize the geodesic net to an immersed closed geodesic ). Observe that if and is an embedded stationary geodesic network verifying for every then can be reparametrized as an immersed closed geodesic whose self intersections occur precisely at the points .
Taking the previous into account, instead of the in the proof of Proposition 4.1 we will work with the following. Consider the set of pairs where is a graph, as a union of connected components, and for every . Given such a pair and a natural number we define to be the set of all such that there exists a stationary geodesic network verifying
-
(1)
For each , is an embedding and verifies the relations for every .
-
(2)
for every .
-
(3)
for every .
-
(4)
for every , and every pair such that .
-
(5)
for every , and .
-
(6)
for every , and .
-
(7)
.
One more remark is necessary to adapt the proof of Proposition 5.1. The sequences in (11) have length uniformly bounded by some and consist of finite unions of almost embedded closed geodesics. This implies that the number of closed geodesics whose union is is also bounded (independently on ). Thus by applying Arzela-Ascoli to each of those components we can get a subsequence whose limit is not only a stationary geodesic net but also a union of closed curves with uniform convergence in . The rest of the proof follows that of Theorem 1.5 word for word.
References
- [1] F. Almgren, The homotopy groups of the integral cycle groups, Topology (1962), 257–299.
- [2] by same author, The theory of varifolds, Mimeographed notes, Princeton (1965).
- [3] A. L. Besse, Einstein manifolds, Springer-Verlag, 1987.
- [4] R. Bowen, The equidistribution of closed geodesics, American Journal of Mathematics 94 (1972), no. 2, 413–423.
- [5] E. Calabi and J. Cao, Simple closed geodesics on convex surfaces, Journal of Differential Geometry 36 (1992), no. 3, 517–549.
- [6] D. Chen, On the closing lemma for Hamiltonian flows on symplectic -manifolds, preprint (2019), https://arxiv.org/abs/1904.09900.
- [7] O. Chodosh and C. Mantoulidis, The p-widths of a surface, preprint (2021), https://arxiv.org/abs/2107.11684.
- [8] V. Climenhaga, G. Knieper, and K. War, Closed geodesics on surfaces without conjugate points, preprint (2020), https://arxiv.org/abs/2008.02249.
- [9] M. Gromov, Isoperimetry of waists and concentration of maps, Geometric and Functional Analysis 13 (2003), 178–215.
- [10] L. Guth, Minimax problems related to cup powers and Steenrod squares, Geometric and Functional Analysis 18 (2009), 1917–1987.
- [11] L. Guth and Y. Liokumovich, Parametric inequalites and Weyl law for the volume spectrum, preprint (2022), https://arxiv.org/abs/2202.11805.
- [12] K. Irie, Equidistributed periodic orbits of -generic three-dimensional Reeb flows, Journal of Symplectic Geometry 19 (2019), no. 3, 531–566.
- [13] K. Irie, F. C. Marques, and A. Neves, Density of minimal hypersurfaces for generic metrics, Annals of Mathematics 187 (2018), 963–972.
- [14] Y. Liokumovich, F. C. Marques, and A. Neves, Weyl law for the volume spectrum, Annals of mathematics 187 (2018), 933–961.
- [15] Y. Liokumovich and B. Staffa, Generic density of geodesic nets, preprint (2021), https://arxiv.org/abs/2107.12340.
- [16] Tsoy-Wo Ma, Banach-Hilbert spaces, vector measures and group representations, World Scientific Pub Co Inc, 2002.
- [17] F. Marques and A. Neves, Existence of infinitely many minimal hypersurfaces in positive Ricci curvature, Inventiones mathematicae 209 (2017).
- [18] F. C. Marques, A. Neves, and A. Song, Equidistribution of minimal hypersurfaces for generic metrics, Inventiones mathematicae 216 (2019), 421–443.
- [19] A. Nabutovsky and R. Rotman, Volume, diameter and the minimal mass of a stationary 1-cycle, Geometric and Functional Analysis 14 (2004), 748–790.
- [20] J. T. Pitts, Regularity and singularity of one dimensional stationary integral varifolds on manifolds arising from variational methods in the large, in Symposia Mathematica XIV (1973), 465–472.
- [21] by same author, Existence and regularity of minimal surfaces on Riemannian manifolds, Princeton University Press, 1981.
- [22] M. Pollicott, Asymptotic distribution of closed geodesics, Israel Journal of Mathematics 52 (1985), 209–224.
- [23] B. Schapira and S. Tapie, Narrow equidistribution and counting of closed geodesics on noncompact manifolds, Groups, Geometry, and Dynamics, European Mathematical Society 15 (2021), no. 3, 1085–1101.
- [24] B. Staffa, Bumpy metrics theorem for geodesic nets, preprint (2021), https://arxiv.org/abs/2107.12446.
- [25] B. White, The space of minimal submanifolds for varying Riemannian metrics, Indiana University Mathematics Department 40 (1991), no. 1, 161–200.
- [26] by same author, On the bumpy metrics theorem for minimal submanifolds, American Journal of Mathematics 139 (2017), no. 4, 1149–1155.