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

On the equidistribution of closed geodesics and geodesic nets

Xinze Li and Bruno Staffa
Abstract.

We show that given a closed nn-manifold MM, for a Baire-generic set of Riemannian metrics gg on MM there exists a sequence of closed geodesics that are equidistributed in MM if n=2n=2; and an equidistributed sequence of embedded stationary geodesic nets if n=3n=3. One of the main tools that we use is the Weyl Law for the volume spectrum for 11-cycles, proved in [14] for n=2n=2 and in [11] for n=3n=3. We show that our proof of the equidistribution of stationary geodesic nets can be generalized for any dimension n2n\geq 2 provided the Weyl Law for 11-cycles in nn-manifolds holds.

1. Introduction

Marques, Neves and Song proved in [18] that for a generic set of Riemannian metrics in a closed manifold MnM^{n}, 3n73\leq n\leq 7 there exists a sequence of closed, embedded, connected minimal hypersurfaces which is equidistributed in MM. In this paper, we study the equidistribution of closed geodesics and stationary geodesic nets (which are 11-dimensional analogs of minimal hypersurfaces) on a Riemannian manifold (Mn,g)(M^{n},g), n2n\geq 2. We prove the following two results, for dimensions 22 and 33 of the ambient manifold respectively:

Theorem 1.1.

Let MM be a closed 22-manifold. For a Baire-generic set of CC^{\infty} Riemannian metrics gg on MM, there exists a set of closed geodesics that is equidistributed in MM. Specifically, for every gg in the generic set, there exists a sequence {γi:S1M}\{\gamma_{i}:S^{1}\rightarrow M\} of closed geodesics in (M,g)(M,g), such that for every CC^{\infty} function f:Mf:M\rightarrow\mathbb{R} we have

limki=1kγifdLgi=1kLg(γi)=MfdVolgVol(M,g).\lim_{k\rightarrow\infty}\frac{\sum_{i=1}^{k}\int_{\gamma_{i}}f\operatorname{dL}_{g}}{\sum_{i=1}^{k}\operatorname{L}_{g}(\gamma_{i})}=\frac{\int_{M}f\operatorname{dVol}_{g}}{\operatorname{Vol}(M,g)}.
Theorem 1.2.

Let MM be a closed 33-manifold. For a Baire-generic set of CC^{\infty} Riemannian metrics gg on MM, there exists a set of connected embedded stationary geodesic nets that is equidistributed in MM. Specifically, for every gg in the generic set , there exists a sequence {γi:ΓiM}\{\gamma_{i}:\Gamma_{i}\rightarrow M\} of connected embedded stationary geodesic nets in (M,g)(M,g), such that for every CC^{\infty} function f:Mf:M\rightarrow\mathbb{R} we have

limki=1kγifdLgi=1kLg(γi)=MfdVolgVol(M,g).\lim_{k\rightarrow\infty}\frac{\sum_{i=1}^{k}\int_{\gamma_{i}}f\operatorname{dL}_{g}}{\sum_{i=1}^{k}\operatorname{L}_{g}(\gamma_{i})}=\frac{\int_{M}f\operatorname{dVol}_{g}}{\operatorname{Vol}(M,g)}.
Remark 1.3.

We have an equivalent notion of equidistribution for a sequence of closed geodesics or geodesic nets: we say that {γi}i\{\gamma_{i}\}_{i\in\mathbb{N}} is equidistributed in (M,g)(M,g) if for every open subset UMU\subseteq M it holds

limki=1kLg(γiU)i=1kLg(γi)=VolgUVolgM.\lim_{k\to\infty}\frac{\sum_{i=1}^{k}\operatorname{L}_{g}(\gamma_{i}\cap U)}{\sum_{i=1}^{k}\operatorname{L}_{g}(\gamma_{i})}=\frac{\operatorname{Vol}_{g}U}{\operatorname{Vol}_{g}M}.

Theorem 1.2 is, as far as the authors know, the first result on equidistribution of kk-stationary varifolds in Riemannian nn-manifolds for k<n1k<n-1 (i.e. in codimension greater than 11). 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 (Mn,g)(M^{n},g) with n2n\geq 2 (possibly with boundary), we have

limpωpn1(M,g)p1n=α(n)Vol(M,g)n1n\lim_{p\rightarrow\infty}\omega_{p}^{n-1}(M,g)p^{-\frac{1}{n}}=\alpha(n)\operatorname{Vol}(M,g)^{\frac{n-1}{n}}

for some constant α(n)>0\alpha(n)>0. Here, given 1kn11\leq k\leq n-1 we denote by ωpk(M,g)\omega_{p}^{k}(M,g) the kk-dimensional pp-width of MM with respect to the metric gg (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 11-dimensional cycles. The following is the most general version of the Weyl law we could expect for 11-cycles.

Conjecture 1.4.

Let (Mn,g)(M^{n},g) be a closed nn-dimensional manifold, n2n\geq 2. Then there exists a constant α(n,1)>0\alpha(n,1)>0 such that

limpωp1(Mn,g)pn1n=α(n,1)Vol(Mn,g)1n.\lim_{p\rightarrow\infty}\omega^{1}_{p}(M^{n},g)p^{-\frac{n-1}{n}}=\alpha(n,1)\operatorname{Vol}(M^{n},g)^{\frac{1}{n}}.

So far, Conjecture 1.4 has been proved for n=2n=2 as a particular case of [14] and recently for n=3n=3 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 22. The only obstruction to extend our proof of the equidistribution of stationary geodesic nets to arbitrary dimensions of the ambient manifold MM is that Conjecture 1.4 has not been proved yet if n>3n>3. As all the rest of our argument works for any dimension n>3n>3, 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 MnM^{n}, n2n\geq 2 be a closed manifold. Assume that the Weyl law for 11-cycles in nn-manifolds holds. Then for a Baire-generic set of CC^{\infty} Riemannian metrics gg on MM, there exists a set of connected embedded stationary geodesic nets that is equidistributed in MM. Specifically, for every gg in the generic set , there exists a sequence {γi:ΓiM}\{\gamma_{i}:\Gamma_{i}\rightarrow M\} of connected embedded stationary geodesic nets in (M,g)(M,g), such that for every CC^{\infty} function f:Mf:M\rightarrow\mathbb{R} we have

limki=1kγifdLgi=1kLg(γi)=MfdVolgVol(M,g).\lim_{k\rightarrow\infty}\frac{\sum_{i=1}^{k}\int_{\gamma_{i}}f\operatorname{dL}_{g}}{\sum_{i=1}^{k}\operatorname{L}_{g}(\gamma_{i})}=\frac{\int_{M}f\operatorname{dVol}_{g}}{\operatorname{Vol}(M,g)}.

In order to simplify the exposition, we consider integrals of CC^{\infty} functions instead of the more general traces of 22-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 gg be a Riemannian metric on MM. We want to do a very small perturbation of gg to obtain a new metric g^\hat{g} which admits a sequence of equidistributed stationary geodesic networks. Let f:Mf:M\to\mathbb{R} be a smooth function. Consider a conformal perturbation g^:(δ,δ)\hat{g}:(-\delta,\delta)\to\mathcal{M}^{\infty} (for some δ>0\delta>0 small) defined as

g^(t)=e2tfg.\hat{g}(t)=e^{2tf}g.

By [15, Lemma 3.4] the normalized pp-widths tpn1nωp1(M,g^(t))t\mapsto p^{-\frac{n-1}{n}}\omega_{p}^{1}(M,\hat{g}(t)) are uniformly locally Lipschitz. This combined with the Weyl Law (recall that we assume it holds) implies that the sequence of functions hp:(δ,δ)h_{p}:(-\delta,\delta)\to\mathbb{R}

hp(t)=pn1nωp1(M,g^(t))Vol(M,g^(t))1nh_{p}(t)=\frac{p^{-\frac{n-1}{n}}\omega_{p}^{1}(M,\hat{g}(t))}{\operatorname{Vol}(M,\hat{g}(t))^{\frac{1}{n}}}

converges uniformly to the constant α(n,1)\alpha(n,1). Considering

h~p(t)=log(hp(t))=n1nlog(p)+log(ωp(M,g^(t)))1nlog(Vol(M,g^(t)))\tilde{h}_{p}(t)=\log(h_{p}(t))=-\frac{n-1}{n}\log(p)+\log(\omega_{p}(M,\hat{g}(t)))-\frac{1}{n}\log(\operatorname{Vol}(M,\hat{g}(t)))

we have that h~p\tilde{h}_{p} converges uniformly to the constant log(α(n,1))\log(\alpha(n,1)). On the other hand, Almgren showed that there is a correspondence between 11-widths and the volumes of stationary varifolds (see [1], [2], [5], [19], [20], [21]) such that for each pp\in\mathbb{N} and t(δ,δ)t\in(-\delta,\delta) there exists a (possibly non unique) stationary geodesic network γp(t)\gamma_{p}(t) such that

(1) Lg^(t)(γp(t))=ωp1(g^(t)).\operatorname{L}_{\hat{g}(t)}(\gamma_{p}(t))=\omega_{p}^{1}(\hat{g}(t)).

Assume that the γp(t)\gamma_{p}(t)’s can be chosen so that all of them are parametrized by the same graph Γ\Gamma and the maps (δ,δ)Ω(Γ,M)(-\delta,\delta)\to\Omega(\Gamma,M), tγp(t)t\mapsto\gamma_{p}(t) are differentiable (this is a very strong assumption and doesn’t necessarily hold, as the map tωp1(g^(t))t\mapsto\omega_{p}^{1}(\hat{g}(t)) may not be differentiable; a counterexample is shown below). In that case we can differentiate h~p\tilde{h}_{p} and obtain

ddth~p(t)\displaystyle\frac{d}{dt}\tilde{h}_{p}(t) =1ωp(M,g^(t))ddtωp1(M,g^(t))1nVol(M,g^(t))ddtVol(M,g^(t))\displaystyle=\frac{1}{\omega_{p}(M,\hat{g}(t))}\frac{d}{dt}\omega_{p}^{1}(M,\hat{g}(t))-\frac{1}{n\operatorname{Vol}(M,\hat{g}(t))}\frac{d}{dt}\operatorname{Vol}(M,\hat{g}(t))
=1Lg^(t)(γp(t))γp(t)fdLg^(t)1nVol(M,g^(t))MnfdVolg^(t)\displaystyle=\frac{1}{L_{\hat{g}(t)}(\gamma_{p}(t))}\int_{\gamma_{p}(t)}f\operatorname{dL}_{\hat{g}(t)}-\frac{1}{n\operatorname{Vol}(M,\hat{g}(t))}\int_{M}nf\operatorname{dVol}_{\hat{g}(t)}
=γp(t)fdLg^(t)MfdVolg^(t).\displaystyle=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{p}(t)}f\operatorname{dL}_{\hat{g}(t)}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\hat{g}(t)}.

As {h~p}p\{\tilde{h}_{p}\}_{p} converges uniformly to a constant, we could expect that the sequence {h~p(t)}p\{\tilde{h}_{p}^{\prime}(t)\}_{p} converges to 0 for some values of tt. If that was the case, the sequence {γp(t)}p\{\gamma_{p}(t)\}_{p} would verify the equidistribution formula for the function ff with respect to the metric g^(t)\hat{g}(t). 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 0 at any point. Indeed, we can construct a sequence of zigzag functions which converges uniformly to 0 but hp(t)h_{p}^{\prime}(t) does not converge to 0 for any tt. The second one is that the differentiability of tγp(t)t\mapsto\gamma_{p}(t) could fail, a counterexample is shown in the next paragraph. And even if that reasoning was true and such tt existed, the sequence {γp(t)}p\{\gamma_{p}(t)\}_{p} constructed would only give an equidistribution formula for the function ff (which is used to construct the sequence) instead of for all CC^{\infty} functions at the same time; and with respect to a metric g^(t)\hat{g}(t) which could also vary with ff.

An example when tωp1(g^(t))t\mapsto\omega_{p}^{1}(\hat{g}(t)) is not differentiable is the following. Let us consider a dumbbell metric gg on S2S^{2} obtained by constructing a connected sum of two identical round 22-spheres S12S^{2}_{1} and S22S^{2}_{2} of radius 11 by a thin neck. Define a 11-parameter family of metrics {g^(t)}t(1,1)\{\hat{g}(t)\}_{t\in(-1,1)} such that g^(t)=(1+t)2g\hat{g}(t)=(1+t)^{2}g along S12S^{2}_{1}, g^(t)=(1t)2g\hat{g}(t)=(1-t)^{2}g along S22S^{2}_{2} (interpolating along the neck so that it is still very thin). It is clear than for t0t\geq 0, the 11-width is realized by a great circle in S12S^{2}_{1} with length 1+t1+t, and for t0t\leq 0 it is realized by a great circle in S22S^{2}_{2} of length 1t1-t. Therefore

ω11(g^(t))={1tif t01+tif t0\omega_{1}^{1}(\hat{g}(t))=\begin{cases}1-t&\text{if }t\leq 0\\ 1+t&\text{if }t\geq 0\end{cases}

and hence it is not differentiable at 0.

To fix the previous issue (differentiability of ωp1(g(t))\omega^{1}_{p}(g(t))), we prove Proposition 4.1 which is a version for stationary geodesic networks of [18, Lemma 2]. Regarding the convergence of hp(t)h_{p}^{\prime}(t) to 0 for certain values of tt, 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 CC^{\infty} 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 pp-widths in a similar way as the γp(t)\gamma_{p}(t)’s above. The key idea here is that the integral of any CC^{\infty} function ff over MM can be approximated by Riemann sums along small regions with piecewise smooth boundary where ff 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 fC(M,)f\in C^{\infty}(M,\mathbb{R}). 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 11 cycles in 33 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 pp-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 22-manifold (M,g)(M,g), its unit cotangent bundle UgMU^{*}_{g}M is a closed 33-manifold equipped with a natural contact structure induced by the contact form λg\lambda_{g} which is the restriction of the Liouville form λ\lambda on TMT^{*}M to UgMU^{*}_{g}M. It is a well known fact that the Reeb vector field associated to λg\lambda_{g} generates the geodesic flow of (M,g)(M,g). Additionally, given a function f:Mf:M\to\mathbb{R}, the Riemannian metric g=efgg^{\prime}=e^{f}g corresponds to the conformal perturbation λg=efπ2λg\lambda_{g^{\prime}}=e^{\frac{f\circ\pi}{2}}\lambda_{g} of the contact form in UgMU^{*}_{g}M (here π:UgMM\pi:U^{*}_{g}M\to M is the projection map); and both λg\lambda_{g} and λg\lambda_{g^{\prime}} are compatible with the same contact structure on UgMU^{*}_{g}M. Thus one would like to apply [12, Corollary 1.4] to UgMU^{*}_{g}M with the contact structure induced by λg\lambda_{g}. However, that result is about generic perturbations of the contact form of the type ef~λge^{\tilde{f}}\lambda_{g}, where f~:UgM\tilde{f}:U^{*}_{g}M\to\mathbb{R} and we only want to consider perturbations f~=fπ\tilde{f}=f\circ\pi which are liftings to UgMU^{*}_{g}M of maps f:Mf:M\to\mathbb{R} 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 gg. 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 Γ=(,𝒱,{n(E)}E)\Gamma=(\mathscr{E},\mathscr{V},\{n(E)\}_{E\in\mathscr{E}}) consisting of a set of edges \mathscr{E}, a set of vertices 𝒱\mathscr{V} and a multiplicity n(E)n(E)\in\mathbb{N} assigned to each edge EE\in\mathscr{E}. A weighted multigraph is good if it is connected and either it is a closed loop with mutiplicity or each vertex v𝒱v\in\mathscr{V} has at least three different incoming edges (here loop edges EE at vv count twice as an incoming edge at vv, see [24] for a more detailed discussion). In the later case we say Γ\Gamma is good*.

Definition 2.2.

Given a weighted multigraph (,𝒱,{n(E)}E)(\mathscr{E},\mathscr{V},\{n(E)\}_{E\in\mathscr{E}}), we identify each edge EE\in\mathscr{E} with the interval [0,1][0,1] and we denote π=πE:{0,1}𝒱\pi=\pi_{E}:\{0,1\}\to\mathscr{V} the map sending i{0,1}i\in\{0,1\} to the vertex v𝒱v\in\mathscr{V} under the identification E[0,1]E\cong[0,1].

Definition 2.3 (Γ\Gamma-net).

A Γ\Gamma-net γ\gamma on MM is a continuous map γ:ΓM\gamma:\Gamma\rightarrow M which is a C2C^{2}-immersion restricted to the edges of Γ\Gamma. We will denote Ω(Γ,M)\Omega(\Gamma,M) the space of Γ\Gamma-nets on MM. It has a natural Banach manifold structure as a subspace of EC2(E,M)\prod_{E\in\mathscr{E}}C^{2}(E,M) (see [24]).

Definition 2.4.

We say that two Γ\Gamma-nets γ1\gamma_{1} and γ2\gamma_{2} are equivalent if for every edge EE of Γ\Gamma the map γ1|E\gamma_{1}|_{E} is a C2C^{2} reparametrization of γ2|E\gamma_{2}|_{E} fixing the endpoints. This defines an equivalence relation \sim in Ω(Γ,M)\Omega(\Gamma,M). We denote Ω^(Γ,M)=Ω(Γ,M)/\hat{\Omega}(\Gamma,M)=\Omega(\Gamma,M)/\sim the quotient space. Given γΩ^(Γ,M)\gamma\in\hat{\Omega}(\Gamma,M) we will often denote a representative of the equivalence class γ\gamma also by γ\gamma, and regard different representatives as different parametrizations of the geometric object γΩ^(Γ,M)\gamma\in\hat{\Omega}(\Gamma,M).

Notation 2.5.

Given a Γ\Gamma-net γ\gamma and an edge EE\in\mathscr{E}, we denote γE\gamma_{E} the restriction of γ\gamma to EE. We also define γE(0):=γE(πE(0))\gamma_{E}(0):=\gamma_{E}(\pi_{E}(0)) and γE(1):=γE(πE(1))\gamma_{E}(1):=\gamma_{E}(\pi_{E}(1)).

Notation 2.6.

Given 1q1\leq q\leq\infty, let us denote q\mathcal{M}^{q} the set of CqC^{q} Riemannian metrics on MM.

Definition 2.7.

Let γΩ^(Γ,M)\gamma\in\hat{\Omega}(\Gamma,M) and let hh be a continuous function defined in Im(γ)M\operatorname{Im}(\gamma)\subseteq M. Given a metric gqg\in\mathcal{M}^{q} we define

γhdLg=En(E)Ehγ(u)gγ(u)(γ˙(u),γ˙(u))𝑑u.\int_{\gamma}h\operatorname{dL}_{g}=\sum_{E\in\mathscr{E}}n(E)\int_{E}h\circ\gamma(u)\sqrt{g_{\gamma(u)}(\dot{\gamma}(u),\dot{\gamma}(u))}du.

Observe that the right hand side is independent of the parametrization we choose and therefore γhdLg\int_{\gamma}h\operatorname{dL}_{g} is well defined.

Definition 2.8 (gg-Length).

Given gqg\in\mathcal{M}^{q} and γΩ^(Γ,M)\gamma\in\hat{\Omega}(\Gamma,M), we define the gg-length of γ\gamma by

Lg(γ)=Γ1dLg=En(E)Egγ(u)(γ˙(u),γ˙(u))𝑑u.\operatorname{L}_{g}(\gamma)=\int_{\Gamma}1\operatorname{dL}_{g}=\sum_{E\in\mathscr{E}}n(E)\int_{E}\sqrt{g_{\gamma(u)}(\dot{\gamma}(u),\dot{\gamma}(u))}du.
Definition 2.9 (Stationary Geodesic Network).

We say that γΩ(Γ,M)\gamma\in\Omega(\Gamma,M) is a stationary geodesic network with respect to a metric gqg\in\mathcal{M}^{q} (q2q\geq 2) if it is a critical point of the length functional Lg:Ω(Γ,M)\operatorname{L}_{g}:\Omega(\Gamma,M)\rightarrow\mathbb{R}. This means that given any smooth one parameter family γ~:(δ,δ)Ω(Γ,M)\tilde{\gamma}:(-\delta,\delta)\to\Omega(\Gamma,M) with γ~(0)=γ\tilde{\gamma}(0)=\gamma we have

dds|s=0Lg(γ~(s))=0.\frac{d}{ds}\bigg{|}_{s=0}\operatorname{L}_{g}(\tilde{\gamma}(s))=0.

Assuming that the edges of γ\gamma are parametrized by constant speed, if X(t)=γ~s(0,t)X(t)=\frac{\partial\tilde{\gamma}}{\partial s}(0,t) (here we regard γ~:(δ,δ)×ΓM\tilde{\gamma}:(-\delta,\delta)\times\Gamma\to M) then

(2) dds|s=0Lg(γ~(s))=En(E)l(E)Eγ¨(t),X(t)g𝑑t+v𝒱Vv(γ),X(v)g\frac{d}{ds}\bigg{|}_{s=0}\operatorname{L}_{g}(\tilde{\gamma}(s))=-\sum_{E\in\mathscr{E}}\frac{n(E)}{l(E)}\int_{E}\langle\ddot{\gamma}(t),X(t)\rangle_{g}dt+\sum_{v\in\mathscr{V}}\langle V_{v}(\gamma),X(v)\rangle_{g}

where l(E)=Lg(γE)l(E)=\operatorname{L}_{g}(\gamma_{E}) and

Vv(γ)=(E,i):πE(i)=v(1)i+1n(E)γ˙E(i)|γ˙E(i)|.V_{v}(\gamma)=\sum_{(E,i):\pi_{E}(i)=v}(-1)^{i+1}n(E)\frac{\dot{\gamma}_{E}(i)}{|\dot{\gamma}_{E}(i)|}.

Equation (2) is called the First Variation Formula and was computed in [24, Section 1]. It implies that γ:ΓM\gamma:\Gamma\to M is stationary with respect to gg if and only if each edge is mapped to a geodesic segment in (M,g)(M,g) and the stability condition at the vertices Vv(γ)=0V_{v}(\gamma)=0 is verified. The latter means that for each v𝒱v\in\mathscr{V}, the sum of the inward pointing unit tangent vectors to each edge at vv is 0.

Definition 2.10.

We say that γΩ^(Γ,M)\gamma\in\hat{\Omega}(\Gamma,M) is a stationary geodesic network if every representative γ~Ω(Γ,M)\tilde{\gamma}\in\Omega(\Gamma,M) of γ\gamma is a stationary geodesic network.

Definition 2.11.

We denote C2(γ)C^{2}(\gamma) the space of continuous vector fields along γ\gamma whose restriction to each edge is of class C2C^{2}.

Remark 2.12.

If gqg\in\mathcal{M}^{q}, q2q\geq 2 and γΩ(Γ,M)\gamma\in\Omega(\Gamma,M) is stationary with respect to gg then by the regularity of the solutions of an ODE, γE\gamma_{E} is of class CqC^{q} for every EE\in\mathscr{E}. This is why we only ask C2C^{2} regularity to Γ\Gamma-nets and vector fields along them.

Assume γΩ(Γ,M)\gamma\in\Omega(\Gamma,M) is a stationary geodesic net with respect to a CqC^{q} metric with q3q\geq 3 (so that the Riemann curvature tensor is of class C1C^{1}). Let γ~:(δ,δ)2Ω(Γ,M)\tilde{\gamma}:(-\delta,\delta)^{2}\to\Omega(\Gamma,M) be a smooth 22-parameter family of Γ\Gamma-nets with γ~(0,0)=γ\tilde{\gamma}(0,0)=\gamma. Let X(t)=γ~x(0,0,t)X(t)=\frac{\partial\tilde{\gamma}}{\partial x}(0,0,t) and Y(t)=γ~s(0,0,t)Y(t)=\frac{\partial\tilde{\gamma}}{\partial s}(0,0,t). We define the Hessian HessγLg:C2(γ)×C2(γ)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}:C^{2}(\gamma)\times C^{2}(\gamma)\to\mathbb{R} of the length functional at γ\gamma as the bilinear form

HessγLg(X,Y)=2xs|(0,0)Lg(γ~(x,s)).\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(X,Y)=\frac{\partial^{2}}{\partial x\partial s}\bigg{|}_{(0,0)}\operatorname{L}_{g}(\tilde{\gamma}(x,s)).

In [24, Section 2] it was shown that HessγLg\operatorname{Hess}_{\gamma}\operatorname{L}_{g} is well defined (i.e. it does not depend on which two parameter variation γ~\tilde{\gamma} with directional derivatives XX and YY we choose) and in fact it holds

(3) HessγLg(X,Y)=EEAE(X)(t),Y(t)g𝑑t+v𝒱Bv(X),Y(v)g\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(X,Y)=\sum_{E\in\mathscr{E}}\int_{E}\langle A_{E}(X)(t),Y(t)\rangle_{g}dt+\sum_{v\in\mathscr{V}}\langle B_{v}(X),Y(v)\rangle_{g}

where

AE(X)\displaystyle A_{E}(X) =n(E)l(E)[X¨E+R(γ˙,XE)γ˙]\displaystyle=-\frac{n(E)}{l(E)}[\ddot{X}_{E}^{\perp}+R(\dot{\gamma},X_{E}^{\perp})\dot{\gamma}]
Bv(X)\displaystyle B_{v}(X) =(E,i):πE(i)=v(1)i+1n(E)l(E)X˙E(i)\displaystyle=\sum_{(E,i):\pi_{E}(i)=v}(-1)^{i+1}\frac{n(E)}{l(E)}\dot{X}_{E}^{\perp}(i)

being XEX_{E} the restriction of the vector field XX to the edge EE and XEX_{E}^{\perp} the component of XEX_{E} orthogonal to γE\gamma_{E}. Observe that AEA_{E} is (up to a positive constant) the Jacobi operator along γE\gamma_{E}. Equation (3) is the Second Variation Formula.

Definition 2.13.

We say that a vector field JC2(γ)J\in C^{2}(\gamma) is Jacobi if it is a null vector of HessγLg\operatorname{Hess}_{\gamma}\operatorname{L}_{g}, i.e. if HessγLg(J,X)=0\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(J,X)=0 for every XC2(γ)X\in C^{2}(\gamma). By the Second Variation Formula, JJ is Jacobi along γ\gamma if and only if AE(J)=0A_{E}(J)=0 for every EE\in\mathscr{E} and Bv(J)=0B_{v}(J)=0 for every v𝒱v\in\mathscr{V}.

Definition 2.14.

A vector field XC2(γ)X\in C^{2}(\gamma) is said to be parallel if XEX_{E} is parallel along γE\gamma_{E} for every EE\in\mathscr{E}.

Remark 2.15.

Observe that every parallel vector field JJ along γ\gamma is Jacobi.

Definition 2.16.

A stationary geodesic net γ:Γ(M,g)\gamma:\Gamma\to(M,g) is nondegenerate if every Jacobi field along γ\gamma is parallel.

Remark 2.17.

In [15, Lemma 2.5], it was shown that every stationary geodesic network with respect to a metric gg can be represented by a map γ:ΓM\gamma:\Gamma\to M, where Γ=i=1PΓi\Gamma=\bigcup_{i=1}^{P}\Gamma_{i} is the finite union of the good weighted multigraphs {Γi}1iP\{\Gamma_{i}\}_{1\leq i\leq P} and γ|Γi\gamma|_{\Gamma_{i}} is an embedded stationary geodesic network for each 1iP1\leq i\leq P (moreover, the map γ:ΓM\gamma:\Gamma\to M is a topological embedding).

Definition 2.18.

Given a stationary geodesic network γ:Γ(M,g)\gamma:\Gamma\to(M,g), we say that its connected components are nondegenerate if

  1. (1)

    We can express Γ=i=1PΓi\Gamma=\bigcup_{i=1}^{P}\Gamma_{i} as a disjoint union of good weighted multigraphs.

  2. (2)

    γ|Γi\gamma|_{\Gamma_{i}} is an embedded nondegenerate stationary geodesic network for every 1iP1\leq i\leq P.

Definition 2.19.

An almost embedded closed geodesic in a Riemannian manifold (M,g)(M,g) is a map γ:S1(M,g)\gamma:S^{1}\to(M,g) such that

  1. (1)

    γ\gamma is geodesic (i.e. γ¨(t)=0\ddot{\gamma}(t)=0 for every tS1t\in S^{1}).

  2. (2)

    γ\gamma is an immersion (i.e. γ˙(t)0\dot{\gamma}(t)\neq 0 for every tS1t\in S^{1}).

  3. (3)

    All self-intersections of γ\gamma are transverse, which means that for every s,tS1s,t\in S^{1} such that γ(s)=γ(t)\gamma(s)=\gamma(t) and sts\neq t, the velocities γ˙(s)\dot{\gamma}(s) and γ˙(t)\dot{\gamma}(t) 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 22-tensor TT, a metric gqg\in\mathcal{M}^{q}, a stationary geodesic network γ:ΓM\gamma:\Gamma\to M on (M,g)(M,g) and tΓt\in\Gamma, we denote

trγ,gT(t)=T(γ˙(t)|γ˙(t)|g,γ˙(t)|γ˙(t)|g)\operatorname{tr}_{\gamma,g}T(t)=T(\frac{\dot{\gamma}(t)}{|\dot{\gamma}(t)|_{g}},\frac{\dot{\gamma}(t)}{|\dot{\gamma}(t)|_{g}})

which is the trace of the tensor TT along γ\gamma with respect to the metric gg.

Definition 2.21 (Average integral along γ\gamma).

Let Γ\Gamma be a weighted multigraph. Given γΩ(Γ,M)\gamma\in\Omega(\Gamma,M), a metric gqg\in\mathcal{M}^{q} and a continuous function hh defined in Im(γ)\operatorname{Im}(\gamma), we define the average integral of hh with respect to metric gg as

γhdLg\displaystyle\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma}h\operatorname{dL}_{g} :=1Lg(γ)γhdLg.\displaystyle:=\frac{1}{\operatorname{L}_{g}(\gamma)}\int_{\gamma}h\operatorname{dL}_{g}.

3. The Jacobi Operator

In this section we will study some properties of the Jacobi operator of an embedded stationary geodesic network γ:Γ(M,g)\gamma:\Gamma\to(M,g), where Γ\Gamma is a good weighted multigraph and gqg\in\mathcal{M}^{q}, q3q\geq 3. We will focus on the case when Γ\Gamma is good* (i.e. every vertex has at least three different incoming edges), because when Γ\Gamma 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

C2(γ)\displaystyle C^{2}(\gamma) ={X continuous vector field along γ:XE is C2 E}\displaystyle=\{X\text{ continuous vector field along }\gamma:X_{E}\text{ is }C^{2}\text{ }\forall E\in\mathscr{E}\}
C2(γ)\displaystyle C^{2}(\gamma)^{\parallel} ={XC2(γ):X(t)γ˙(t) tΓ}\displaystyle=\{X\in C^{2}(\gamma):X(t)\in\langle\dot{\gamma}(t)\rangle\text{ }\forall t\in\Gamma\}
C02(γ)\displaystyle C^{2}_{0}(\gamma)^{\perp} ={XC2(γ):X(t)γ˙(t) tΓ𝒱 and X(v)=0 v𝒱}\displaystyle=\{X\in C^{2}(\gamma):X(t)\perp\dot{\gamma}(t)\text{ }\forall t\in\Gamma\setminus\mathscr{V}\text{ and }X(v)=0\text{ }\forall v\in\mathscr{V}\}
C2(E)\displaystyle C^{2}(E)^{\perp} ={XC2(E):X(t)γ˙(t) E}\displaystyle=\{X\in C^{2}(E):X(t)\perp\dot{\gamma}(t)\text{ }\forall E\in\mathscr{E}\}
C2()\displaystyle C^{2}(\mathscr{E})^{\perp} =EC2(E).\displaystyle=\prod_{E\in\mathscr{E}}C^{2}(E)^{\perp}.

Observe that as Γ\Gamma is good*, if XC2(γ)X\in C^{2}(\gamma)^{\parallel} then X(v)=0X(v)=0 for every v𝒱v\in\mathscr{V}. Denote

T𝒱=v𝒱Tγ(v)M.T\mathscr{V}=\prod_{v\in\mathscr{V}}T_{\gamma(v)}M.

By the second variation formula (3), we can define the Jacobi operator L:C2(γ)C0()×T𝒱L:C^{2}(\gamma)\to C^{0}(\mathscr{E})^{\perp}\times T\mathscr{V} as

(4) L(J)=((n(E)l(E)(J¨E+R(γ˙,JE)γ˙))E,(Bv(J))v𝒱).L(J)=((-\frac{n(E)}{l(E)}(\ddot{J}_{E}^{\perp}+R(\dot{\gamma},J_{E}^{\perp})\dot{\gamma}))_{E\in\mathscr{E}},(B_{v}(J))_{v\in\mathscr{V}}).

We know that each XC2(γ)X\in C^{2}(\gamma)^{\parallel} is Jacobi (i.e. it verifies L(J)=0L(J)=0). We want to construct a complement of C2(γ)C^{2}(\gamma)^{\parallel} in C2(γ)C^{2}(\gamma), and show that when we restrict LL 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 S2(γ)C2(γ)S^{2}(\gamma)\subseteq C^{2}(\gamma) such that the evaluation map ev:S2(γ)T𝒱\operatorname{ev}:S^{2}(\gamma)\to T\mathscr{V}, J(J(v))v𝒱J\mapsto(J(v))_{v\in\mathscr{V}} is a linear isomorphism. This can be done by taking a basis v\mathcal{B}_{v} of Tγ(v)MT_{\gamma(v)}M for each v𝒱v\in\mathscr{V}, and for each pair (v,w)(v,w) with v𝒱v\in\mathscr{V} and wvw\in\mathcal{B}_{v} defining a vector field J(v,w)C2(γ)J_{(v,w)}\in C^{2}(\gamma) such that J(v,w)(v)=wJ_{(v,w)}(v)=w and J(v,w)(v)=0J_{(v,w)}(v^{\prime})=0 for every vvv^{\prime}\neq v. Then we can define S2(γ)=J(v,w):v𝒱,wvS^{2}(\gamma)=\langle J_{(v,w)}:v\in\mathscr{V},w\in\mathcal{B}_{v}\rangle. Of course the choice of S2(γ)S^{2}(\gamma) 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 S2(γ)S^{2}(\gamma)). It is clear that

C2(γ)=C2(γ)C02(γ)S2(γ).C^{2}(\gamma)=C^{2}(\gamma)^{\parallel}\oplus C^{2}_{0}(\gamma)^{\perp}\oplus S^{2}(\gamma).

Denote C2(γ)C=C02(γ)S2(γ)C^{2}(\gamma)^{C}=C^{2}_{0}(\gamma)^{\perp}\oplus S^{2}(\gamma) which is a complement of the space of parallel vector fields along γ\gamma. Same as in the theory of elliptic operators, we can extend the Jacobi operator to Sobolev spaces of vector fields along γ\gamma once we have a suitable definition of them. Denote

H02(E)\displaystyle H^{2}_{0}(E) ={X normal vector field of class H02 along E}\displaystyle=\{X\text{ normal vector field of class }H^{2}_{0}\text{ along }E\}
H02(γ)\displaystyle H^{2}_{0}(\gamma) =EH02(E)\displaystyle=\prod_{E\in\mathscr{E}}H^{2}_{0}(E)
H2(γ)\displaystyle H^{2}(\gamma) =H02(γ)S2(γ)\displaystyle=H^{2}_{0}(\gamma)\oplus S^{2}(\gamma)
L2(E)\displaystyle L^{2}(E) ={X normal vector field of class L2 along E}\displaystyle=\{X\text{ normal vector field of class }L^{2}\text{ along }E\}
L2()\displaystyle L^{2}(\mathscr{E}) =EL2(E)\displaystyle=\prod_{E\in\mathscr{E}}L^{2}(E)
L2(γ)\displaystyle L^{2}(\gamma) =L2()T𝒱.\displaystyle=L^{2}(\mathscr{E})\oplus T\mathscr{V}.

Notice that H2(γ)H^{2}(\gamma) is the H2H^{2}-version of C2(γ)CC^{2}(\gamma)^{C} and will be the domain of the Jacobi operator we will work with (as that operator vanishes on C2(γ)C^{2}(\gamma)^{\parallel}). The previous spaces are defined in analogy with the spaces of C2C^{2}, H2H^{2} and L2L^{2} normal vector fields along a smooth closed submanifold which appear when studying the ellipticity of its Jacobi operator. The space L2(γ)L^{2}(\gamma) is a Hilbert space with the inner product

((XE)E,(uv)v),((YE)E,(wv)v)=EEXE(t),YE(t)g𝑑t+v𝒱uv,wvg\langle((X_{E})_{E},(u_{v})_{v}),((Y_{E})_{E},(w_{v})_{v})\rangle=\sum_{E\in\mathscr{E}}\int_{E}\langle X_{E}(t),Y_{E}(t)\rangle_{g}dt+\sum_{v\in\mathscr{V}}\langle u_{v},w_{v}\rangle_{g}

and we have a monomorphism ι:H2(γ)L2(γ)\iota:H^{2}(\gamma)\to L^{2}(\gamma) with dense image given by

ι(J)=((JE)E,(J(v))v𝒱)\iota(J)=((J_{E})^{\perp}_{E\in\mathscr{E}},(J(v))_{v\in\mathscr{V}})

which allow us to write the Hessian HessγLg:H2(γ)×H2(γ)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}:H^{2}(\gamma)\times H^{2}(\gamma)\to\mathbb{R} as

HessγLg(J,J~)=L(J),ι(J~)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(J,\tilde{J})=\langle L(J),\iota(\tilde{J})\rangle

where ,\langle,\rangle is the inner product in L2(γ)L^{2}(\gamma). Here we considered L:H2(γ)L2(γ)L:H^{2}(\gamma)\to L^{2}(\gamma) given by (4) which is a bounded linear operator. As in the smooth case, we can also regard LL as an unbounded operator L:L2(γ)L2(γ)L:L^{2}(\gamma)\to L^{2}(\gamma) whose domain is the dense linear subspace H2(γ)H^{2}(\gamma). We would therefore expect that for sufficiently big λ\lambda\in\mathbb{R} the operator L+λι:L2(γ)L2(γ)L+\lambda\iota:L^{2}(\gamma)\to L^{2}(\gamma) has a compact inverse, and from that get an orthonormal basis of L2(γ)L^{2}(\gamma) consisting of eigenvectors of LL. This indeed holds, as it is shown in the following proposition.

Proposition 3.1.

For every λ\lambda\in\mathbb{R}, the operator Lλι:H2(γ)L2(γ)L-\lambda\iota:H^{2}(\gamma)\to L^{2}(\gamma) defined as (Lλι)(J)=L(J)λι(J)(L-\lambda\iota)(J)=L(J)-\lambda\iota(J) is Fredholm of index 0. The spectrum of LL consists of an increasing sequence of eigenvalues λ1λ2\lambda_{1}\leq\lambda_{2}\leq... with limnλn=+\lim_{n\to\infty}\lambda_{n}=+\infty (i.e. LλιL-\lambda\iota has nontrivial kernel if and only if λ{λn}n\lambda\in\{\lambda_{n}\}_{n\in\mathbb{N}} and has a continuous inverse (Lλι)1:L2(γ)H2(γ)(L-\lambda\iota)^{-1}:L^{2}(\gamma)\to H^{2}(\gamma) otherwise). In addition, there exists sequence {Jn}n\{J_{n}\}_{n\in\mathbb{N}} in H2(γ)H^{2}(\gamma) such that {ι(Jn)}n\{\iota(J_{n})\}_{n\in\mathbb{N}} is an orthonormal basis of L2(γ)L^{2}(\gamma) and L(Jn)=λnι(Jn)L(J_{n})=\lambda_{n}\iota(J_{n}) for each nn\in\mathbb{N}. Therefore, we have the following min-max characterization of the eigenvalues of LL

λn=minWmaxJW{0}L(J),ι(J)ι(J),ι(J)\lambda_{n}=\min_{W}\max_{J\in W\setminus\{0\}}\frac{\langle L(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle}

where the minimum is taken over all nn-dimensional subspaces WH2(γ)W\subseteq H^{2}(\gamma).

Proof.

Let λ\lambda\in\mathbb{R}. Then if JH02(γ)J\in H^{2}_{0}(\gamma) and J~S2(γ)\tilde{J}\in S^{2}(\gamma),

(Lλι)(J+J~)=((LE(JE)λJE)E+(LE(J~E)λJ~E)E,(Bv(J+J~)λJ~(v))v)(L-\lambda\iota)(J+\tilde{J})=((L_{E}(J_{E})-\lambda J_{E})_{E}+(L_{E}(\tilde{J}_{E}^{\perp})-\lambda\tilde{J}_{E}^{\perp})_{E},(B_{v}(J+\tilde{J})-\lambda\tilde{J}(v))_{v})

where LE:H02(E)L2(E)L_{E}:H^{2}_{0}(E)\to L^{2}(E) is (a constant multiple of) the Jacobi operator along γE\gamma_{E} given by Jn(E)l(E)(J¨+R(γ˙,J)γ˙)J\mapsto-\frac{n(E)}{l(E)}(\ddot{J}+R(\dot{\gamma},J)\dot{\gamma}). We know that each LEL_{E} is elliptic, and therefore LEλL_{E}-\lambda is Fredholm of index 0 for every λ\lambda\in\mathbb{R}. This implies that the product operator L~:H02(γ)L2()\tilde{L}:H^{2}_{0}(\gamma)\to L^{2}(\mathscr{E}), L~=(LE)E\tilde{L}=(L_{E})_{E} verifies that L~λ\tilde{L}-\lambda is Fredholm of index 0 for every λ\lambda\in\mathbb{R}. Thus the fact that LλιL-\lambda\iota is always Fredholm of index 0 can be deduced from the following lemma:

Lemma 3.2.

Let E1E_{1}, E2E_{2}, E¯1\overline{E}_{1}, E¯2\overline{E}_{2} be Banach spaces with dim(E2)=dim(E¯2)<\dim(E_{2})=\dim(\overline{E}_{2})<\infty. Let L:E1E2E¯1E¯2L:E_{1}\oplus E_{2}\to\overline{E}_{1}\oplus\overline{E}_{2} be a continuous linear map, and write L(e1.e2)=(L11(e1)+L21(e2),L12(e1)+L22(e2))L(e_{1}.e_{2})=(L_{11}(e_{1})+L_{21}(e_{2}),L_{12}(e_{1})+L_{22}(e_{2})) with Lij:EiE¯jL_{ij}:E_{i}\to\overline{E}_{j}. Assume L11L_{11} is Fredholm of index 0. Then LL is Fredholm of index 0.

Proof of Lemma 3.2.

Let L~:E1E2E¯1E¯2\tilde{L}:E_{1}\oplus E_{2}\to\overline{E}_{1}\oplus\overline{E}_{2} be the operator L~(e1,e2)=(L11(e1),0)\tilde{L}(e_{1},e_{2})=(L_{11}(e_{1}),0). Because L11L_{11} is Fredholm of index 0 and dim(E2)=dim(E¯2)\dim(E_{2})=\dim(\overline{E}_{2}), we see that L~\tilde{L} is also Fredholm of index 0. As L=L~+FL=\tilde{L}+F with F(e1,e2)=(L21(e2),L12(e1)+L22(e2))F(e_{1},e_{2})=(L_{21}(e_{2}),L_{12}(e_{1})+L_{22}(e_{2})) compact because of the finite dimensionality of E2E_{2} and E¯2\overline{E}_{2}, by [16, Theorem 12-5.13] we deduce that LL is also Fredholm of index 0. ∎

Now we are going to show that the quadratic form HessγLg:H2(γ)×H2(γ)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}:H^{2}(\gamma)\times H^{2}(\gamma)\to\mathbb{R} is bounded from below. We know

HessγLg(J,J~)=EELE(JE)(t),J~E(t)g𝑑t+v𝒱Bv(J),J~(v)g.\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(J,\tilde{J})=\sum_{E\in\mathscr{E}}\int_{E}\langle L_{E}(J_{E}^{\perp})(t),\tilde{J}_{E}^{\perp}(t)\rangle_{g}dt+\sum_{v\in\mathscr{V}}\langle B_{v}(J),\tilde{J}(v)\rangle_{g}.

Denote by C:H2(γ)×H2(γ)C:H^{2}(\gamma)\times H^{2}(\gamma)\to\mathbb{R} the form C(J,J~)=v𝒱Bv(J),J~(v)gC(J,\tilde{J})=\sum_{v\in\mathscr{V}}\langle B_{v}(J),\tilde{J}(v)\rangle_{g}. CC is symmetric because so are HessγLg\operatorname{Hess}_{\gamma}\operatorname{L}_{g} and LEL_{E} for each EE\in\mathscr{E}. If we endow S2(γ)S^{2}(\gamma) with the inner product

J,J~=v𝒱J(v),J~(v)g\langle J,\tilde{J}\rangle=\sum_{v\in\mathscr{V}}\langle J(v),\tilde{J}(v)\rangle_{g}

then as dim(S2(γ))<\dim(S^{2}(\gamma))<\infty, we can see that there exists some constant α>0\alpha>0 such that

(5) |C(J,J)|αv𝒱J(v),J(v)g|C(J,J)|\leq\alpha\sum_{v\in\mathscr{V}}\langle J(v),J(v)\rangle_{g}

for every JS2(γ)J\in S^{2}(\gamma). But then as CC vanishes on H02(γ)H^{2}_{0}(\gamma), by its bilinearity and symmetry we can see that in fact (5) is valid for every JH2(γ)J\in H^{2}(\gamma).

On the other hand, using that each LEL_{E} is elliptic, for each EE\in\mathscr{E} there exists βE\beta_{E}\in\mathbb{R} such that

(6) ELE(JE)(t),JE(t)g𝑑tβEEJE(t),JE(t)g𝑑t.\int_{E}\langle L_{E}(J^{\perp}_{E})(t),J_{E}^{\perp}(t)\rangle_{g}dt\geq\beta_{E}\int_{E}\langle J_{E}^{\perp}(t),J_{E}^{\perp}(t)\rangle_{g}dt.

Thus if β=min{βE:E}\beta=\min\{\beta_{E}:E\in\mathscr{E}\} and γ=min{β,α}\gamma=\min\{\beta,-\alpha\}, from (5) and (6) we deduce

HessγLg(J,J)βEEJE(t),JE(t)g𝑑tαv𝒱J(v),J(v)gγι(J),ι(J)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(J,J)\geq\beta\sum_{E\in\mathscr{E}}\int_{E}\langle J_{E}^{\perp}(t),J_{E}^{\perp}(t)\rangle_{g}dt-\alpha\sum_{v\in\mathscr{V}}\langle J(v),J(v)\rangle_{g}\geq\gamma\langle\iota(J),\iota(J)\rangle

which considering that HessγLg(J,J)=L(J),ι(J)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(J,J)=\langle L(J),\iota(J)\rangle implies that for every λ\lambda\in\mathbb{R} it holds

(L+λι)(J),ι(J)(λ+γ)ι(J),ι(J)\langle(L+\lambda\iota)(J),\iota(J)\rangle\geq(\lambda+\gamma)\langle\iota(J),\iota(J)\rangle

and in particular if λ>γ\lambda>-\gamma implies that L+λιL+\lambda\iota is a monomorphism. Because we also know that these operators are Fredholm of index 0, by the Open Mapping Theorem we conclude that L+λι:H2(γ)L2(γ)L+\lambda\iota:H^{2}(\gamma)\to L^{2}(\gamma) is a continuous linear isomorphism for every λ>γ\lambda>-\gamma.

Fix λ>γ\lambda>-\gamma. We will now show that ι(L+λι)1:L2(γ)L2(γ)\iota\circ(L+\lambda\iota)^{-1}:L^{2}(\gamma)\to L^{2}(\gamma) is compact. Let (Xn)n(X_{n})_{n\in\mathbb{N}} be a bounded sequence in L2(γ)L^{2}(\gamma) and define (Jn,J~n)=(L+λι)1(Xn)(J^{n},\tilde{J}^{n})=(L+\lambda\iota)^{-1}(X_{n}) with JnH02(γ)J^{n}\in H^{2}_{0}(\gamma) and J~nS2(γ)\tilde{J}^{n}\in S^{2}(\gamma). As (L+λι)1(L+\lambda\iota)^{-1} is bounded, (Jn,J~n)(J^{n},\tilde{J}^{n}) is a bounded sequence in H2(γ)H^{2}(\gamma). Therefore, for each EE\in\mathscr{E} the sequence of normal vector fields (JEn)n(J^{n}_{E})_{n\in\mathbb{N}} along γE\gamma_{E} is bounded in H02(E)H^{2}_{0}(E) and therefore in H01(E)H^{1}_{0}(E). Hence, by the Rellich-Kondrachov Compactness Theorem we can find a subsequence (nk)k(n_{k})_{k\in\mathbb{N}} such that (JEnk)k(J^{n_{k}}_{E})_{k\in\mathbb{N}} converges in L2(E)L^{2}(E) for every EE\in\mathscr{E}. On the other hand, using that S2(γ)S^{2}(\gamma) is finite dimensional, we can extract a further subsequence (nkl)l(n_{k_{l}})_{l} to have the additional property that (J~nkl)l(\tilde{J}^{n_{k_{l}}})_{l\in\mathbb{N}} converges in S2(γ)S^{2}(\gamma). This implies that the sequence of general term ι(L+λι)1(Xnkl)=ι(Jnkl,J~nkl)\iota\circ(L+\lambda\iota)^{-1}(X_{n_{k_{l}}})=\iota(J^{n_{k_{l}}},\tilde{J}^{n_{k_{l}}}) converges in L2(γ)L^{2}(\gamma), and this completes the proof that ι(L+λι)1\iota\circ(L+\lambda\iota)^{-1} is compact.

The symmetry of HessγLg(J,J~)=L(J),ι(J~)\operatorname{Hess}_{\gamma}\operatorname{L}_{g}(J,\tilde{J})=\langle L(J),\iota(\tilde{J})\rangle implies that ι(L+λι)1\iota\circ(L+\lambda\iota)^{-1} is self-adjoint, which together with its compactness implies the existence of an orthonormal basis {Xn}n\{X_{n}\}_{n\in\mathbb{N}} of L2(γ)L^{2}(\gamma) such that ι(L+λι)1Xn=δnXn\iota\circ(L+\lambda\iota)^{-1}X_{n}=\delta_{n}X_{n} for some decreasing sequence δn0+\delta_{n}\to 0^{+} (because by our choice of λ\lambda, ι(L+λι)10\iota\circ(L+\lambda\iota)^{-1}\geq 0). But we claim that XL2(γ)X\in L^{2}(\gamma) is an eigenvector of ι(L+λι)1\iota\circ(L+\lambda\iota)^{-1} of eigenvalue δ\delta\in\mathbb{R} if and only if X=ι(J)X=\iota(J) for some JH2(γ)J\in H^{2}(\gamma) such that L(J)=(δ1λ)ι(J)L(J)=(\delta^{-1}-\lambda)\iota(J). This is because ι(L+λι)1(X)=δX\iota\circ(L+\lambda\iota)^{-1}(X)=\delta X if and only if there exists JH2(γ)J\in H^{2}(\gamma) with ι(J)=X\iota(J)=X which verifies any of the the following equivalent conditions:

ι(L+λι)1ι(J)\displaystyle\iota\circ(L+\lambda\iota)^{-1}\circ\iota(J) =δι(J)\displaystyle=\delta\iota(J)
(L+λι)1ι(J)\displaystyle(L+\lambda\iota)^{-1}\circ\iota(J) =δJ\displaystyle=\delta J
ι(J)\displaystyle\iota(J) =δ(L+λι)(J)\displaystyle=\delta(L+\lambda\iota)(J)
L(J)\displaystyle L(J) =(δ1λ)ι(J).\displaystyle=(\delta^{-1}-\lambda)\iota(J).

From the previous, we conclude that if λn:=δn1λ\lambda_{n}:=\delta_{n}^{-1}-\lambda then spec(L)={λn}n\operatorname{spec}(L)=\{\lambda_{n}\}_{n}, limnλn=+\lim_{n\to\infty}\lambda_{n}=+\infty and L(Jn)=λnι(Jn)L(J_{n})=\lambda_{n}\iota(J_{n}). This implies the min-max theorem for the eigenvalues holds for LL, which completes the proof. ∎

4. Some auxiliary results

Proposition 4.1.

Let g:INqg:I^{N}\to\mathcal{M}^{q} be a smooth embedding, NN\in\mathbb{N}, I=(1,1)I=(-1,1). If qN+3q\geq N+3, there exists an arbitrarily small perturbation in the CC^{\infty} topology g:INqg^{\prime}:I^{N}\to\mathcal{M}^{q} such that there is a full measure subset 𝒜IN\mathcal{A}\subseteq I^{N} with the following properties: for any pp\in\mathbb{N} and any t𝒜t\in\mathcal{A}, the function sωp1(g(s))s\mapsto\omega^{1}_{p}(g^{\prime}(s)) is differentiable at tt, and there exists a (possibly disconnected) weighted multigraph Γ\Gamma and a stationary geodesic network γp=γp(t):Γ(M,g(t))\gamma_{p}=\gamma_{p}(t):\Gamma\to(M,g^{\prime}(t)) such that the following two conditions hold

  1. (1)

    ωp1(g(t))=Lg(t)(γp(t))\omega_{p}^{1}(g^{\prime}(t))=L_{g^{\prime}(t)}(\gamma_{p}(t)).

  2. (2)

    v(ωp1g)|s=t=12γp(t)trγp(t),g(t)gv(t)dLg(t)\frac{\partial}{\partial v}(\omega^{1}_{p}\circ g^{\prime})\big{|}_{s=t}=\frac{1}{2}\int_{\gamma_{p}(t)}\operatorname{tr}_{\gamma_{p}(t),g^{\prime}(t)}\frac{\partial g^{\prime}}{\partial v}(t)\operatorname{dL}_{g^{\prime}(t)}.

To prove the proposition, we will need to have a condition for a sequence of embedded stationary geodesic nets γn:Γ(M,gn)\gamma_{n}:\Gamma\to(M,g_{n}) converging to some γ:Γ(M,g)\gamma:\Gamma\to(M,g) that guarantees that γ\gamma 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 (g,γ)(g,\gamma) where γ\gamma is stationary with respect to gg. We proceed to describe those functionals.

The first one is

F1(g,γ)=min{|γ˙E(t)|g:tE,E}.F_{1}(g,\gamma)=\min\{|\dot{\gamma}_{E}(t)|_{g}:t\in E,E\in\mathscr{E}\}.

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 F2(E1,i1),(E2,i2)F_{2}^{(E_{1},i_{1}),(E_{2},i_{2})} defined for each pair ((E1,i1),(E2,i2))(×{0,1})2((E_{1},i_{1}),(E_{2},i_{2}))\in(\mathscr{E}\times\{0,1\})^{2} such that πE1(i1)=πE2(i2)\pi_{E_{1}}(i_{1})=\pi_{E_{2}}(i_{2}) (see Section 2 for the notation) as follows

F2(E1,i1),(E2,i2)(g,γ)=(1)i1+i2γ˙E1(i1),γ˙E2(i2)g|γ˙E1(i1)|g|γ˙E2(i2)|g.F_{2}^{(E_{1},i_{1}),(E_{2},i_{2})}(g,\gamma)=(-1)^{i_{1}+i_{2}}\frac{\langle\dot{\gamma}_{E_{1}}(i_{1}),\dot{\gamma}_{E_{2}}(i_{2})\rangle_{g}}{|\dot{\gamma}_{E_{1}}(i_{1})|_{g}|\dot{\gamma}_{E_{2}}(i_{2})|_{g}}.

Notice that (1)ijγ˙Ej(ij)|γ˙Ej(ij)|g(-1)^{i_{j}}\frac{\dot{\gamma}_{E_{j}}(i_{j})}{|\dot{\gamma}_{E_{j}}(i_{j})|_{g}} is the unit inward tangent vector to γ\gamma at v=πEj(ij)v=\pi_{E_{j}}(i_{j}) along EjE_{j}, j=1,2j=1,2 (and observe that in case EE is a loop at vv, there are two inward tangent vectors to γ\gamma along EE at vv represented by the pairs (E,0)(E,0) and (E,1)(E,1)). The condition F2(E1,i1),(E2,i2)(gn,γn)1δF_{2}^{(E_{1},i_{1}),(E_{2},i_{2})}(g_{n},\gamma_{n})\leq 1-\delta for some δ>0\delta>0 and for every possible choice (E1,i1)(E2,i2)(E_{1},i_{1})\neq(E_{2},i_{2}) with πE1(i1)=πE2(i2)\pi_{E_{1}}(i_{1})=\pi_{E_{2}}(i_{2}) implies that the limit (g,γ)(g,\gamma) has the property that given v𝒱v\in\mathscr{V}, there exists an open neighborhood UvU_{v} of vv in Γ\Gamma such that γ:Uvγ(Uv)\gamma:U_{v}\to\gamma(U_{v}) is a homeomorphism. Explicitly,

Uv=(E,i):πE(i)=v{tE:|ti|<min{inj(g)Lg(γE),12}}U_{v}=\bigcup_{(E,i):\pi_{E}(i)=v}\{t\in E:|t-i|<\min\{\frac{\operatorname{inj}(g)}{\operatorname{L}_{g}(\gamma_{E})},\frac{1}{2}\}\}

where inj:q>0\operatorname{inj}:\mathcal{M}^{q}\to\mathbb{R}_{>0}, ginj(g)g\mapsto\operatorname{inj}(g) is a continuous choice of the injectivity radius for each CqC^{q} Riemannian metric gg. This is because if we consider UvU_{v} as a graph obtained by gluing at vv one edge for each pair (E,i)×{0,1}(E,i)\in\mathscr{E}\times\{0,1\} such that πE(i)=v\pi_{E}(i)=v, this graph is mapped by γ\gamma into a geodesic ball centered at γ(v)\gamma(v) of radius inj(g)\operatorname{inj}(g) and the image of each incoming edge at vv has a different inward tangent vector at γ(v)\gamma(v).

To ensure injectivity along the edges, we define for each edge EE\in\mathscr{E} a function

d(g,γ)E(t)=min{dg(γ(t),γ(s)):sE,|ts|inj(g)Lg(γE)}.d^{E}_{(g,\gamma)}(t)=\min\{d_{g}(\gamma(t),\gamma(s)):s\in E,|t-s|\geq\frac{\operatorname{inj}(g)}{\operatorname{L}_{g}(\gamma_{E})}\}.

In case πE(0)=πE(1)\pi_{E}(0)=\pi_{E}(1), the distance |st||s-t| between two points s,tEs,t\in E is measured with respect of the length of S1=E/01S^{1}=E/0\sim 1.

To ensure that the images of different edges under γ\gamma do not overlap, we define for each pair E,EE,E^{\prime}\in\mathscr{E}, EEE\neq E^{\prime} a function d(g,γ)E,E:E0d^{E,E^{\prime}}_{(g,\gamma)}:E\to\mathbb{R}_{\geq 0} as

d(g,γ)E,E(t)=\displaystyle d^{E,E^{\prime}}_{(g,\gamma)}(t)= min{dg(γ(t),γ(s)):sE,|si|inj(g)Lg(γE) for each i{0,1}\displaystyle\min\{d_{g}(\gamma(t),\gamma(s)):s\in E^{\prime},|s-i|\geq\frac{\operatorname{inj}(g)}{\operatorname{L}_{g}(\gamma_{E^{\prime}})}\text{ for each }i\in\{0,1\}
s.t. j{0,1} with πE(i)=πE(j) and |tj|inj(g)Lg(γE)}.\displaystyle\text{ s.t. }\exists j\in\{0,1\}\text{ with }\pi_{E^{\prime}}(i)=\pi_{E}(j)\text{ and }|t-j|\leq\frac{\operatorname{inj}(g)}{\operatorname{L}_{g}(\gamma_{E})}\}.

Let us fix an auxiliary embedding ψ:Ml\psi:M\to\mathbb{R}^{l} and identify from now on our manifold MM with the submanifold ψ(M)l\psi(M)\subseteq\mathbb{R}^{l}. Given a multigraph Γ\Gamma and a continuous map γ:ΓM\gamma:\Gamma\to M which is C3C^{3} when restricted to each edge, we can consider

γ3=γ0+γ˙0+γ¨0+γ˙˙˙0\|\gamma\|_{3}=\|\gamma\|_{0}+\|\dot{\gamma}\|_{0}+\|\ddot{\gamma}\|_{0}+\|\dddot{\gamma}\|_{0}

where given a collection u=(uE)Eu=(u_{E})_{E\in\mathscr{E}} of continuous functions along the edges of Γ\Gamma, we define

u0=max{|uE(t)|:tE,E}\|u\|_{0}=\max\{|u_{E}(t)|:t\in E,E\in\mathscr{E}\}

being |||\cdot| the Euclidean norm in l\mathbb{R}^{l}. We have the following compactness result.

Lemma 4.2.

Let (gn)n(g_{n})_{n\in\mathbb{N}} be a sequence of C3C^{3} Riemannian metrics converging to some metric g3g\in\mathcal{M}^{3}. Let γn:Γ(M,gn)\gamma_{n}:\Gamma\to(M,g_{n}) be a sequence of stationary geodesic networks. Assume γn3M\|\gamma_{n}\|_{3}\leq M for some M>0M\in\mathbb{R}_{>0}. Then there exists a subsequence (γnk)k(\gamma_{n_{k}})_{k} and γΩ(Γ,M)\gamma\in\Omega(\Gamma,M) such that limkγnk=γ\lim_{k\to\infty}\gamma_{n_{k}}=\gamma in Ω(Γ,M)\Omega(\Gamma,M) and γ:Γ(M,g)\gamma:\Gamma\to(M,g) is stationary.

Proof.

The Arzela-Ascoli Theorem gives a subsequence γnkγ\gamma_{n_{k}}\to\gamma in Ω(Γ,M)\Omega(\Gamma,M). The fact that γ\gamma is stationary with respect to gg comes from the continuity of the operator HH defined in [24] (which plays the role of the mean curvature operator on minimal surfaces) which vanishes in a pair (g,[γ])(g,[\gamma]) if and only if γ\gamma is stationary with respect to gg. ∎

We will also need the following two lemmas.

Lemma 4.3.

Let F:nF:\mathbb{R}^{n}\to\mathbb{N} be a function. Then there exists mm\in\mathbb{N} and a basis {v1,,vn}\{v_{1},...,v_{n}\} of n\mathbb{R}^{n} such that F(vi)=mF(v_{i})=m for all 1in1\leq i\leq n.

Proof.

Observe that n=mF1(m)\mathbb{R}^{n}=\bigcup_{m\in\mathbb{N}}F^{-1}(m) and therefore n=mF1(m)\mathbb{R}^{n}=\bigcup_{m\in\mathbb{N}}\langle F^{-1}(m)\rangle where given AnA\subseteq\mathbb{R}^{n} we denote A\langle A\rangle the subspace spanned by AA. If F1(m)F^{-1}(m) did not contain a basis of n\mathbb{R}^{n} for every mm\in\mathbb{N}, F1(m)\langle F^{-1}(m)\rangle would a proper subspace for every mm. Therefore, n\mathbb{R}^{n} 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 γ:(1,1)NΩ(Γ,M)\gamma:(-1,1)^{N}\to\Omega(\Gamma,M) and g:(1,1)Nqg:(-1,1)^{N}\to\mathcal{M}^{q} be smooth maps. Assume that γ(s)\gamma(s) is stationary with respect to g(s)g(s) for every s(1,1)Ns\in(-1,1)^{N}. Then for every t(1,1)Nt\in(-1,1)^{N} and every vNv\in\mathbb{R}^{N}

v|s=tL(g(s),γ(s))=12γ(t)trγ(t),g(t)gv(t)dLg(t).\frac{\partial}{\partial v}\big{|}_{s=t}\operatorname{L}(g(s),\gamma(s))=\frac{1}{2}\int_{\gamma(t)}\operatorname{tr}_{\gamma(t),g(t)}\frac{\partial g}{\partial v}(t)\operatorname{dL}_{g(t)}.
Proof.

Using that the length functional is a smooth function L:q×Ω(Γ,M)\operatorname{L}:\mathcal{M}^{q}\times\Omega(\Gamma,M)\to\mathbb{R} and the chain rule, we get

v|s=tL(g(s),γ(s))\displaystyle\frac{\partial}{\partial v}\big{|}_{s=t}\operatorname{L}(g(s),\gamma(s)) =DL(g(t),γ(t))(D(g×γ)t(v))\displaystyle=D\operatorname{L}_{(g(t),\gamma(t))}(D(g\times\gamma)_{t}(v))
=DL(g(t),γ(t))(gv(t),γv(t))\displaystyle=D\operatorname{L}_{(g(t),\gamma(t))}(\frac{\partial g}{\partial v}(t),\frac{\partial\gamma}{\partial v}(t))
=D1L(g(t),γ(t))(gv(t))+D2Lg(t),γ(t))(γv(t))\displaystyle=D_{1}\operatorname{L}_{(g(t),\gamma(t))}(\frac{\partial g}{\partial v}(t))+D_{2}\operatorname{L}_{g(t),\gamma(t))}(\frac{\partial\gamma}{\partial v}(t))
=D1L(g(t),γ(t))(gv(t)).\displaystyle=D_{1}\operatorname{L}_{(g(t),\gamma(t))}(\frac{\partial g}{\partial v}(t)).

The second term in the penultimate equation vanishes because γ(t)\gamma(t) is stationary with respect to g(t)g(t). Hence

v|s=tL(g(s),γ(s))\displaystyle\frac{\partial}{\partial v}\big{|}_{s=t}\operatorname{L}(g(s),\gamma(s)) =dds|s=0L(g(t+sv),γ(t))\displaystyle=\frac{d}{ds}\big{|}_{s=0}\operatorname{L}(g(t+sv),\gamma(t))
=dds|s=0En(E)Egt+sv(γ˙t(u),γ˙t(u))𝑑u\displaystyle=\frac{d}{ds}\big{|}_{s=0}\sum_{E\in\mathscr{E}}n(E)\int_{E}\sqrt{g_{t+sv}(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}du
=En(E)Edds|s=0gt+sv(γ˙t(u),γ˙t(u))du\displaystyle=\sum_{E\in\mathscr{E}}n(E)\int_{E}\frac{d}{ds}\big{|}_{s=0}\sqrt{g_{t+sv}(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}du
=En(E)Egv(t)(γ˙t(u),γ˙t(u))2gt(γ˙t(u),γ˙t(u))𝑑u\displaystyle=\sum_{E\in\mathscr{E}}n(E)\int_{E}\frac{\frac{\partial g}{\partial v}(t)(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}{2\sqrt{g_{t}(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}}du
=12En(E)Egv(t)(γ˙t(u),γ˙t(u))gt(γ˙t(u),γ˙t(u))gt(γ˙t(u),γ˙t(u))𝑑u\displaystyle=\frac{1}{2}\sum_{E\in\mathscr{E}}n(E)\int_{E}\frac{\frac{\partial g}{\partial v}(t)(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}{g_{t}(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}\sqrt{g_{t}(\dot{\gamma}_{t}(u),\dot{\gamma}_{t}(u))}du
=12En(E)γ(t)Etrγ(t),g(t)gv(t)dLg(t)\displaystyle=\frac{1}{2}\sum_{E\in\mathscr{E}}n(E)\int_{\gamma(t)_{E}}\operatorname{tr}_{\gamma(t),g(t)}\frac{\partial g}{\partial v}(t)\operatorname{dL}_{g(t)}
=12γ(t)trγ(t),g(t)gv(t)dLg(t).\displaystyle=\frac{1}{2}\int_{\gamma(t)}\operatorname{tr}_{\gamma(t),g(t)}\frac{\partial g}{\partial v}(t)\operatorname{dL}_{g(t)}.

Proof of Proposition 4.1.

Notice that it suffices to show that for each pp\in\mathbb{N}, there exists a full measure subset 𝒜(p)IN\mathcal{A}(p)\subseteq I^{N} where (1) and (2) hold, because in that case 𝒜=p𝒜(p)\mathcal{A}=\bigcap_{p\in\mathbb{N}}\mathcal{A}(p) will have the desired property. Therefore we will assume pp\in\mathbb{N} is fixed.

Let g:INqg:I^{N}\to\mathcal{M}^{q} be a smooth embedding. Let {Γi}i1\{\Gamma_{i}\}_{i\geq 1} be a sequence enumerating the countable collection of all good weighted multigraphs. For each i1i\geq 1, let 𝒮q(Γi)\mathcal{S}^{q}(\Gamma_{i}) be the space of pairs (g,[γ])(g,[\gamma]) where gqg\in\mathcal{M}^{q}, γ:Γi(M,g)\gamma:\Gamma_{i}\to(M,g) is an embedded stationary geodesic net and [γ][\gamma] 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 𝒮q(Γi)\mathcal{S}^{q}(\Gamma_{i}) is a second countable Banach manifold and the projection map Πi:𝒮q(Γi)q\Pi_{i}:\mathcal{S}^{q}(\Gamma_{i})\to\mathcal{M}^{q} mapping (g,[γ])g(g,[\gamma])\mapsto g is Fredholm of index 0. A pair (g,[γ])𝒮q(Γi)(g,[\gamma])\in\mathcal{S}_{q}(\Gamma_{i}) is a critical point of Πi\Pi_{i} if and only if γ\gamma admits a nontrivial Jacobi field with respect to the metric gg.

By Smale’s transversality theorem, we can perturb g:INqg:I^{N}\to\mathcal{M}^{q} slightly in the CC^{\infty} topology to a CC^{\infty} embedding g:INqg^{\prime}:I^{N}\to\mathcal{M}^{q} which is transversal to Πi:𝒮q(Γi)q\Pi_{i}:\mathcal{S}^{q}(\Gamma_{i})\to\mathcal{M}^{q} for every ii\in\mathbb{N}. Transversality implies that Mi=Πi1(g(IN))M_{i}=\Pi_{i}^{-1}(g^{\prime}(I^{N})) is an NN-dimensional embedded submanifold of 𝒮q(Γi)\mathcal{S}^{q}(\Gamma_{i}) for every ii\in\mathbb{N}. Let πi=(g)1Πi|Mi:MiIN\pi_{i}=(g^{\prime})^{-1}\circ\Pi_{i}\big{|}_{M_{i}}:M_{i}\to I^{N}. Let 𝒜~iIn\tilde{\mathcal{A}}_{i}\subseteq I^{n} be the set of regular values of πi\pi_{i}, which is a set of full measure by Sard’s theorem. Let 𝒜~0IN\tilde{\mathcal{A}}_{0}\subseteq I^{N} be the set of points for which the Lipschitz function sωp1(g(s))s\mapsto\omega^{1}_{p}(g^{\prime}(s)) is differentiable. Observe that 𝒜~0\tilde{\mathcal{A}}_{0} has full measure by Rademacher’s theorem. Therefore, 𝒜~=i0𝒜~i\tilde{\mathcal{A}}=\bigcap_{i\geq 0}\tilde{\mathcal{A}}_{i} is a full measure subset of INI^{N}. Notice that by transversality, if t𝒜~t\in\tilde{\mathcal{A}} then g(t)g^{\prime}(t) is a bumpy metric, i.e. all embedded stationary geodesic nets with respect to g(t)g^{\prime}(t) and with domain a good weighted multigraph are nondegenerate; and also the map sωp1(g(s))s\mapsto\omega_{p}^{1}(g^{\prime}(s)) is differentiable at s=ts=t.

Given a weighted multigraph Γ=i=1PΓi\Gamma=\bigcup_{i=1}^{P}\Gamma_{i} whose connected components Γi\Gamma_{i} are good and a natural number MM\in\mathbb{N}, we define Γ,M\mathcal{B}_{\Gamma,M} as the set of all tINt\in I^{N} such that there exists a stationary geodesic network γ:Γ(M,g(t))\gamma:\Gamma\to(M,g^{\prime}(t)) verifying

  1. (1)

    γi=γ|Γi\gamma_{i}=\gamma|_{\Gamma_{i}} is an embedding for each 1iP1\leq i\leq P.

  2. (2)

    γi3M\|\gamma_{i}\|_{3}\leq M for every 1iP1\leq i\leq P.

  3. (3)

    F1(g(t),γi)1MF_{1}(g^{\prime}(t),\gamma_{i})\geq\frac{1}{M} for every 1iP1\leq i\leq P.

  4. (4)

    F2(E1,i1),(E2,i2)(g(t),γi)11MF_{2}^{(E_{1},i_{1}),(E_{2},i_{2})}(g^{\prime}(t),\gamma_{i})\leq 1-\frac{1}{M} for every 1iP1\leq i\leq P and every pair (E1,i1)(E2,i2)(E_{1},i_{1})\neq(E_{2},i_{2}) in i×{0,1}\mathscr{E}_{i}\times\{0,1\} such that πE1(i1)=πE2(i2)\pi_{E_{1}}(i_{1})=\pi_{E_{2}}(i_{2}).

  5. (5)

    d(g(t),γi)E(s)1Md^{E}_{(g^{\prime}(t),\gamma_{i})}(s)\geq\frac{1}{M} for every 1iP1\leq i\leq P, EiE\in\mathscr{E}_{i} and sEs\in E.

  6. (6)

    d(g(t),γi)E,E(s)1Md^{E,E^{\prime}}_{(g^{\prime}(t),\gamma_{i})}(s)\geq\frac{1}{M} for every 1iP1\leq i\leq P, EEiE\neq E^{\prime}\in\mathscr{E}_{i} and sEs\in E.

  7. (7)

    ωp1(g(t))=Lg(t)(γ)\omega_{p}^{1}(g^{\prime}(t))=\operatorname{L}_{g^{\prime}(t)}(\gamma).

where i\mathscr{E}_{i} denotes the set of edges of Γi\Gamma_{i}. Observe that IN=Γ,MΓ,MI^{N}=\bigcup_{\Gamma,M}\mathcal{B}_{\Gamma,M} because of (1) and Remark 2.17. We claim that each Γ,MIN\mathcal{B}_{\Gamma,M}\subseteq I^{N} is closed.

Indeed, suppose we have a sequence {tj}jΓ,M\{t_{j}\}_{j\in\mathbb{N}}\subseteq\mathcal{B}_{\Gamma,M} converging to some tINt\in I^{N}. Let γj\gamma^{j} be the stationary geodesic network corresponding to g(tj)g^{\prime}(t_{j}) and verifying properties (1) to (7) above. By property (2) and Lemma (4.2), passing to a subsequence we have that if γij=γj|Γi\gamma^{j}_{i}=\gamma^{j}\big{|}_{\Gamma_{i}} then there exists γi:ΓiM\gamma_{i}:\Gamma_{i}\to M such that limjγij=γi\lim_{j\to\infty}\gamma_{i}^{j}=\gamma_{i} in Ω(Γi,M)\Omega(\Gamma_{i},M) and γi\gamma_{i} is stationary with respect to g(t)g^{\prime}(t) for each 1iP1\leq i\leq P. Observe also that if γ=jγi\gamma=\bigcup_{j}\gamma_{i}

Lg(t)(γ)=limjLg(tj)(γj)=limjωp(tj)=ωp(t).L_{g^{\prime}(t)}(\gamma)=\lim_{j\to\infty}L_{g^{\prime}(t_{j})}(\gamma^{j})=\lim_{j\to\infty}\omega_{p}(t_{j})=\omega_{p}(t).

Properties (2) to (6) are preserved when we take the limit of the sequence γj\gamma^{j}, so it suffices to show that γ|Γi\gamma|_{\Gamma_{i}} is embedded for each 1iP1\leq i\leq P. Fix such ii. Properties (3), (4) and (5) imply that γi\gamma_{i} 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 Γ,M\mathcal{B}_{\Gamma,M} is closed, they are measurable and therefore so are the sets 𝒜~Γ,M=𝒜~Γ,M\tilde{\mathcal{A}}_{\Gamma,M}=\tilde{\mathcal{A}}\cap\mathcal{B}_{\Gamma,M} (whose union is 𝒜~\tilde{\mathcal{A}}). Let 𝒜Γ,M\mathcal{\mathcal{A}}_{\Gamma,M} be the set of points t𝒜~Γ,Mt\in\tilde{\mathcal{A}}_{\Gamma,M} where the Lebesgue density of 𝒜~Γ,M\tilde{\mathcal{A}}_{\Gamma,M} at tt is 11. By the Lebesgue Differentiation Theorem, 𝒜~Γ,M𝒜Γ,M\tilde{\mathcal{A}}_{\Gamma,M}\setminus\mathcal{A}_{\Gamma,M} has Lebesgue measure 0 for each pair (Γ,M)(\Gamma,M). Let us define 𝒜=Γ,M𝒜Γ,M\mathcal{A}=\bigcup_{\Gamma,M}\mathcal{A}_{\Gamma,M}, observe that as 𝒜~𝒜\mathcal{\tilde{A}}\setminus\mathcal{A} has measure 0, 𝒜IN\mathcal{A}\subseteq I^{N} has full measure.

Fix t𝒜t\in\mathcal{A}. Let (Γ,M)(\Gamma,M) be such that t𝒜Γ,Mt\in\mathcal{A}_{\Gamma,M}. As the density of 𝒜~Γ,M\tilde{\mathcal{A}}_{\Gamma,M} at tt is 11, given vNv\in\mathbb{R}^{N} with |v|=1|v|=1 we can find a sequence {tm(v)}m𝒜~Γ,M\{t_{m}(v)\}_{m\in\mathbb{N}}\subseteq\tilde{\mathcal{A}}_{\Gamma,M} such that limmtm(v)=t\lim_{m\to\infty}t_{m}(v)=t and limmttm(v)|ttm(v)|=v\lim_{m\to\infty}\frac{t-t_{m}(v)}{|t-t_{m}(v)|}=v. Denoting ωp(s)=ωp1(g(s))\omega_{p}(s)=\omega_{p}^{1}(g^{\prime}(s)), using that ωp\omega_{p} is a Lipschitz function we can see that

(7) limmωp(tm(v))ωp(t)|ttm|=vωp(t).\lim_{m\to\infty}\frac{\omega_{p}(t_{m}(v))-\omega_{p}(t)}{|t-t_{m}|}=\frac{\partial}{\partial v}\omega_{p}(t).

As tm(v)𝒜~Γ,Mt_{m}(v)\in\tilde{\mathcal{A}}_{\Gamma,M}, for each mm\in\mathbb{N} there exists a stationary geodesic network γm:ΓM\gamma_{m}:\Gamma\to M with respect to g(tm(v))g^{\prime}(t_{m}(v)) such that

ωp(tm(v))=ωp1(g(tm(v)))=Lg(tm(v))(γm)\omega_{p}(t_{m}(v))=\omega^{1}_{p}(g^{\prime}(t_{m}(v)))=\operatorname{L}_{g^{\prime}(t_{m}(v))}(\gamma_{m})

and properties (1) to (6) above hold. By the reasoning used to prove that the Γ,M\mathcal{B}_{\Gamma,M} are closed, we can construct a stationary geodesic net γ:Γ(M,g(t))\gamma:\Gamma\to(M,g^{\prime}(t)) which is embedded when restricted to each connected component Γi\Gamma_{i} of Γ\Gamma, is the limit of (a subsequence of) the γm\gamma_{m}’s in the C2C^{2} topology and realizes the width ωp1(g(t))\omega_{p}^{1}(g^{\prime}(t)). Hence from (7) we get

vωp(t)=limmLg(tm)(γm)Lg(t)(γ)|ttm|.\frac{\partial}{\partial v}\omega_{p}(t)=\lim_{m\to\infty}\frac{\operatorname{L}_{g^{\prime}(t_{m})}(\gamma_{m})-\operatorname{L}_{g^{\prime}(t)}(\gamma)}{|t-t_{m}|}.

As γ|Γi\gamma|_{\Gamma_{i}} is an embedded stationary geodesic net with respect to g(t)g^{\prime}(t) for each 1iP1\leq i\leq P and g(t)g^{\prime}(t) is bumpy, Πi:𝒮q(Γi)q\Pi_{i}:\mathcal{S}^{q}(\Gamma_{i})\to\mathcal{M}^{q} is a diffeomorphism from a neighborhood UiU_{i} of (g(t),[γi])(g^{\prime}(t),[\gamma_{i}]) to a neighborhood Wi=Πi(U)W_{i}=\Pi_{i}(U) of g(t)g^{\prime}(t). Denote Ξi\Xi_{i} its inverse. As there exists m0m_{0}\in\mathbb{N} such that g(tm)W=i=1PWig^{\prime}(t_{m})\in W=\bigcap_{i=1}^{P}W_{i} and [γm|Γi]Ui[\gamma_{m}\big{|}_{\Gamma_{i}}]\in U_{i} for every mm0m\geq m_{0}, we deduce that [γm|Γi]=Ξi(g(tm(v)))[\gamma_{m}\big{|}_{\Gamma_{i}}]=\Xi_{i}(g^{\prime}(t_{m}(v))) for each mm0m\geq m_{0} and each 1iP1\leq i\leq P. Let us define Ξ:WΩ^(Γ,M)\Xi:W\to\hat{\Omega}(\Gamma,M) as Ξ(g)=h\Xi(g)=h where h|Γi=Ξi(g)h|_{\Gamma_{i}}=\Xi_{i}(g). Thus by Lemma 4.4

vωp(t)\displaystyle\frac{\partial}{\partial v}\omega_{p}(t) =limmLg(tm)(Ξ(g(tm)))Lg(t)(Ξ(g(t)))|tmt|\displaystyle=\lim_{m\to\infty}\frac{\operatorname{L}_{g^{\prime}(t_{m})}(\Xi(g^{\prime}(t_{m})))-\operatorname{L}_{g^{\prime}(t)}(\Xi(g^{\prime}(t)))}{|t_{m}-t|}
=v|s=tL(g(s),Ξ(g(s)))\displaystyle=\frac{\partial}{\partial v}\big{|}_{s=t}\operatorname{L}(g^{\prime}(s),\Xi(g^{\prime}(s)))
=12γvtrγv,g(t)gv(t)dLg(t).\displaystyle=\frac{1}{2}\int_{\gamma_{v}}\operatorname{tr}_{\gamma_{v},g^{\prime}(t)}\frac{\partial g^{\prime}}{\partial v}(t)\operatorname{dL}_{g^{\prime}(t)}.

Where γv=Ξ(g(t))\gamma_{v}=\Xi(g^{\prime}(t)) is the one constructed before. Observe that γv\gamma_{v} depends on vv and that the previous formula holds for each vNv\in\mathbb{R}^{N}, |v|=1|v|=1. Notice that each γv\gamma_{v} is a stationary geodesic network with respect to g(t)g^{\prime}(t), and as g(t)g^{\prime}(t) is bumpy there are countably many possible γvs\gamma_{v}^{\prime}s, say {hj}j\{h_{j}\}_{j\in\mathbb{N}}. This induces a map F:NF:\mathbb{R}^{N}\to\mathbb{N} defined as F(0)=1F(0)=1 and if w0w\neq 0 then F(w)=jF(w)=j where γw|w|=hj\gamma_{\frac{w}{|w|}}=h_{j}. By Lemma 4.3 we can obtain mm\in\mathbb{N} and a basis w1,,wNw_{1},...,w_{N} of N\mathbb{R}^{N} with the property γ(wi)=m\gamma(w_{i})=m for every 1iN1\leq i\leq N. Therefore if we set vi:=wi|wi|v_{i}:=\frac{w_{i}}{|w_{i}|}, v1,,vNv_{1},...,v_{N} is still a basis and by definition γvi=hm\gamma_{v_{i}}=h_{m} for every ii. By linearity of directional derivatives, denoting γ=hm\gamma=h_{m} we deduce that

vωp(t)=12γtrγ,g(t)gv(t)dLg(t)\frac{\partial}{\partial v}\omega_{p}(t)=\frac{1}{2}\int_{\gamma}\operatorname{tr}_{\gamma,g^{\prime}(t)}\frac{\partial g^{\prime}}{\partial v}(t)\operatorname{dL}_{g^{\prime}(t)}

for every unit vNv\in\mathbb{R}^{N}, which completes the proof. ∎

Proposition 4.5.

Let MM be a closed manifold and let gg be a CqC^{q} Riemannian metric on MM, q3q\geq 3. Let γ1,,γk\gamma_{1},...,\gamma_{k} be a finite collection collection of connected, embedded stationary geodesic networks on (M,g)(M,g) whose domains are good weighted multigraphs and let 𝒰q\mathcal{U}\subseteq\mathcal{M}^{q} be an open neighborhood of gg. Then there exists g𝒰g^{\prime}\in\mathcal{U} such that γ1,γk\gamma_{1},...\gamma_{k} are non-degenerate stationary geodesic nets with respect to gg^{\prime}.

Proof.

Following [18, Lemma 4], we will consider conformal perturbations of the metric of the form gε(x)=e2εϕ(x)g(x)g_{\varepsilon}(x)=e^{-2\varepsilon\phi(x)}g(x). Let us denote γ~=i=1kγi\tilde{\gamma}=\bigcup_{i=1}^{k}\gamma_{i}, Γ~=i=1kΓi\tilde{\Gamma}=\bigcup_{i=1}^{k}\Gamma_{i} (where γi:ΓiM\gamma_{i}:\Gamma_{i}\to M) and ~\tilde{\mathscr{E}} the set of edges of γ~\tilde{\gamma}. Notice that γ~:Γ~M\tilde{\gamma}:\tilde{\Gamma}\to M is a stationary geodesic network whose edges may overlap, even non-transversally. Given EE\in\mathscr{E}, let Reg(γ~E)\operatorname{Reg}(\tilde{\gamma}_{E}) be the set of interior points of γ~E\tilde{\gamma}_{E} which are not points of transverse intersection with any other edge γ~E\tilde{\gamma}_{E}. We define a finite poset

𝒫={i=1lReg(γ~Ei):E1,,El~, EiEj ij}\mathcal{P}=\{\bigcap_{i=1}^{l}\operatorname{Reg}(\tilde{\gamma}_{E_{i}})\neq\emptyset:E_{1},...,E_{l}\in\tilde{\mathscr{E}},\text{ }E_{i}\neq E_{j}\text{ }\forall i\neq j\}

which is the collection of finite non-empty intersections of sets in {Reg(γ~E):E}\{\operatorname{Reg}(\tilde{\gamma}_{E}):E\in\mathscr{E}\}, with the order given by the inclusion. Denote by 𝒫\mathcal{P}^{\prime} the set of minimal elements in 𝒫\mathcal{P}. Observe that if α,α\alpha,\alpha^{\prime} are two different elements of 𝒫\mathcal{P}^{\prime} then they are disjoint. Given α𝒫\alpha\in\mathcal{P}^{\prime}, write α=i=1lReg(γ~Ei)\alpha=\bigcap_{i=1}^{l}\operatorname{Reg}(\tilde{\gamma}_{E_{i}}) in the unique way such that αγ~E=\alpha\cap\tilde{\gamma}_{E}=\emptyset for every E~{E1,,El}E\in\tilde{\mathscr{E}}\setminus\{E_{1},...,E_{l}\}. Pick tααt_{\alpha}\in\alpha for every α𝒫\alpha\in\mathcal{P}^{\prime}, and let η>0\eta>0 be such that the geodesic balls Bα=B(pα,η)B_{\alpha}=B(p_{\alpha},\eta) verify

  • BαBα=B_{\alpha}\cap B_{\alpha^{\prime}}=\emptyset if αα\alpha\neq\alpha^{\prime}.

  • BαγE=B_{\alpha}\cap\gamma_{E}=\emptyset if E{E1,,El}E\notin\{E_{1},...,E_{l}\}.

  • BαγEiαB_{\alpha}\cap\gamma_{E_{i}}\subseteq\alpha for every i=1,,li=1,...,l.

  • There exists a diffeomorphism ρα:Bαn\rho_{\alpha}:B_{\alpha}\to\mathbb{R}^{n} such that ρα(γEiBα)=ρα(αBα)={(t,0,0,,0):t}\rho_{\alpha}(\gamma_{E_{i}}\cap B_{\alpha})=\rho_{\alpha}(\alpha\cap B_{\alpha})=\{(t,0,0,...,0):t\in\mathbb{R}\} for each i=1,,li=1,...,l.

Denote Bα=B(pα,η2)B^{\prime}_{\alpha}=B(p_{\alpha},\frac{\eta}{2}). Observe that for each E~E\in\tilde{\mathscr{E}} there exists at least one α𝒫\alpha\in\mathcal{P}^{\prime} such that αγ~E\alpha\subseteq\tilde{\gamma}_{E}. Choose such an α\alpha for each EE\in\mathscr{E} and denote BE=BαB_{E}=B_{\alpha} and BE=BαB^{\prime}_{E}=B^{\prime}_{\alpha}. We can now proceed to define the function ϕ\phi which will induce the one-parameter family of metrics gε(x)=e2εϕ(x)g_{\varepsilon}(x)=e^{-2\varepsilon\phi(x)} mentioned before.

For each α𝒫\alpha\in\mathcal{P}^{\prime}, let ψα:M\psi_{\alpha}:M\to\mathbb{R} be a smooth function with 0ψα10\leq\psi_{\alpha}\leq 1, spt(ψα)Bα\operatorname{spt}(\psi_{\alpha})\subseteq B_{\alpha} and ψα1\psi_{\alpha}\equiv 1 in BαB^{\prime}_{\alpha}. Let fα:Bαf_{\alpha}:B_{\alpha}\to\mathbb{R} be given in local coordinates under the chart (Bα,ρα)(B_{\alpha},\rho_{\alpha}) by fα(x)=i=2nxi2f_{\alpha}(x)=\sum_{i=2}^{n}x_{i}^{2}. We define ϕ=α𝒫ψαfα\phi=\sum_{\alpha\in\mathcal{P}^{\prime}}\psi_{\alpha}f_{\alpha}. An easy computation shows that DϕD\phi vanishes along γ~\tilde{\gamma} and in local coordinates Hessγ~(t)ϕ(X,Y)=ψα(x)i=2nxiyi\operatorname{Hess}_{\tilde{\gamma}(t)}\phi(X,Y)=\psi_{\alpha}(x)\sum_{i=2}^{n}x_{i}y_{i} if γ~(t)Bα\tilde{\gamma}(t)\in B_{\alpha}, X=(x1,,xn)X=(x_{1},...,x_{n}) and Y=(y1,,yn)Y=(y_{1},...,y_{n}); and Hessγ~(t)ϕ0\operatorname{Hess}_{\tilde{\gamma}(t)}\phi\equiv 0 if tα𝒫Bαt\notin\bigcup_{\alpha\in\mathcal{P}^{\prime}}B_{\alpha}. In particular, if γ~(t)Bα\tilde{\gamma}(t)\in B^{\prime}_{\alpha} for some α𝒫\alpha\in\mathcal{P}^{\prime} then Hessγ~(t)ϕ(X,X)=0\operatorname{Hess}_{\tilde{\gamma}(t)}\phi(X,X)=0 if and only if Xγ˙Ej(t)X\in\langle\dot{\gamma}_{E_{j}}(t)\rangle for some (or equivalently, for every) j{1,,l}j\in\{1,...,l\} where α=i=1lReg(γ~Ei)\alpha=\bigcap_{i=1}^{l}\operatorname{Reg}(\tilde{\gamma}_{E_{i}}).

Therefore we know that ϕ\phi and DϕD\phi vanish along each γi\gamma_{i}. Hence by [3, Theorem 1.159], the γ1,,γk\gamma_{1},...,\gamma_{k} are still stationary with respect to gε(x)=e2εϕ(x)g(x)g_{\varepsilon}(x)=e^{-2\varepsilon\phi(x)}g(x). Fix γ=γi:ΓM\gamma=\gamma_{i}:\Gamma\to M with set of vertices 𝒱\mathscr{V} and set of edges \mathscr{E}. We assume that Γ\Gamma is good* (i.e. every vertex has at least 33 different incoming edges), the case when γ\gamma 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 γ\gamma with respect to gg is the map L:H2(γ)L2(γ)L:H^{2}(\gamma)\to L^{2}(\gamma) given by

L(J)=((LE(J))E,(Bv(J))v𝒱)L(J)=((L_{E}(J))_{E\in\mathscr{E}},(B_{v}(J))_{v\in\mathscr{V}})

where

LE(J)\displaystyle L_{E}(J) =n(E)l(E)[J¨E+R(γ˙,JE),γ˙]\displaystyle=-\frac{n(E)}{l(E)}\bigg{[}\ddot{J}_{E}^{\perp}+R(\dot{\gamma},J_{E}^{\perp}),\dot{\gamma}\bigg{]}
Bv(J)\displaystyle B_{v}(J) =(E,i):πE(i)=v(1)i+1n(E)l(E)J˙E(i).\displaystyle=\sum_{(E,i):\pi_{E}(i)=v}(-1)^{i+1}\frac{n(E)}{l(E)}\dot{J}^{\perp}_{E}(i).

Let us compute which is the change in the Jacobi operator along γ\gamma when we switch from the metric gg to gεg_{\varepsilon}. We will denote LεL^{\varepsilon} the operator corresponding to gεg_{\varepsilon}. Using [3, Theorem 1.159] and the fact that Dϕ=0D\phi=0 along γi\gamma_{i}, we can see that Bvε=BvB_{v}^{\varepsilon}=B_{v} for all v𝒱v\in\mathscr{V}, and that

LEε(J)=n(E)l(E)(J¨E+R(γ˙,JE)γ˙+εHessϕ(JE))L_{E}^{\varepsilon}(J)=-\frac{n(E)}{l(E)}(\ddot{J}_{E}^{\perp}+R(\dot{\gamma},J_{E}^{\perp})\dot{\gamma}+\varepsilon\operatorname{Hess}\phi(J_{E}^{\perp}))

where the covariant derivatives and the curvature tensor RR are taken with respect to the metric gg, and at each point pMp\in M, Hessp(ϕ):TpMTpM\operatorname{Hess}_{p}(\phi):T_{p}M\to T_{p}M is the linear transformation such that the Hessian of ϕ\phi at pp is given by (X,Y)Hesspϕ(X),Yg(X,Y)\mapsto\langle\operatorname{Hess}_{p}\phi(X),Y\rangle_{g}.

We know from Section 3 that each Lε:H2(γ)L2(γ)L^{\varepsilon}:H^{2}(\gamma)\to L^{2}(\gamma) admits a non-decreasing sequence of eigenvalues λ1ελ2ελQε\lambda_{1}^{\varepsilon}\leq\lambda_{2}^{\varepsilon}\leq...\leq\lambda_{Q}^{\varepsilon}\leq... which are characterized by

λiε=infWmaxJW{0}Lε(J),ι(J)ι(J),ι(J)\lambda_{i}^{\varepsilon}=\inf_{W}\max_{J\in W\setminus\{0\}}\frac{\langle L^{\varepsilon}(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle}

where the infimum is taken over all ii-dimensional subspaces WW of H2(γ)H^{2}(\gamma). Also, the map εLε\varepsilon\mapsto L^{\varepsilon} is continuous; therefore λiε\lambda_{i}^{\varepsilon} varies continuously with ε\varepsilon for every ii\in\mathbb{N}. We will use these facts to show that for sufficiently small values of ε>0\varepsilon>0, 0 is not an eigenvalue of LεL^{\varepsilon}.

Let QQ be the unique natural number such that 0=λQ<λQ+10=\lambda_{Q}<\lambda_{Q+1} (here λi:=λi0\lambda_{i}:=\lambda_{i}^{0}). Denote SS the sum of the eigenspaces corresponding to λ1,,λQ\lambda_{1},...,\lambda_{Q}. Let JSJ\in S, J0J\neq 0. Then we have

Lε(J),ι(J)ι(J),ι(J)\displaystyle\frac{\langle L^{\varepsilon}(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle} =L(J),ι(J)ι(J),ι(J)EEn(E)l(E)εHessγ(t)(ϕ)(JE(t)),JE(t)g𝑑tι(J),ι(J)\displaystyle=\frac{\langle L(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle}-\frac{\sum_{E\in\mathscr{E}}\int_{E}\frac{n(E)}{l(E)}\varepsilon\langle\operatorname{Hess}_{\gamma(t)}(\phi)(J_{E}^{\perp}(t)),J_{E}^{\perp}(t)\rangle_{g}dt}{\langle\iota(J),\iota(J)\rangle}
εEn(E)l(E)EHessγ(t)(ϕ)(JE(t)),JE(t)g𝑑tι(J),ι(J)\displaystyle\leq-\varepsilon\sum_{E\in\mathscr{E}}\frac{n(E)}{l(E)}\frac{\int_{E}\langle\operatorname{Hess}_{\gamma(t)}(\phi)(J_{E}^{\perp}(t)),J_{E}^{\perp}(t)\rangle_{g}dt}{\langle\iota(J),\iota(J)\rangle}
0\displaystyle\leq 0

because Hessγ(t)ϕ0\operatorname{Hess}_{\gamma(t)}\phi\geq 0 for every tΓt\in\Gamma. Suppose there is equality for some JS{0}J\in S\setminus\{0\}. Then the two inequalities should be equalities. From the first one we deduce that JJ is Jacobi along γ\gamma for the metric gg, and thus it verifies J¨E+R(γ˙,JE)γ˙=0\ddot{J}^{\perp}_{E}+R(\dot{\gamma},J^{\perp}_{E})\dot{\gamma}=0 for every EE\in\mathscr{E}. From the second one, by considering the values of tt for which γ(t)BE\gamma(t)\in B^{\prime}_{E}, we see that JEJ^{\perp}_{E} is a null vector of Hessγ(t)ϕ\operatorname{Hess}_{\gamma(t)}\phi along γEBE\gamma_{E}\cap B^{\prime}_{E} and therefore JE=0J^{\perp}_{E}=0 on γEBE\gamma_{E}\cap B^{\prime}_{E}; and as it satisfies the Jacobi equation this implies JE=0J^{\perp}_{E}=0 for every EE\in\mathscr{E}. Thus JJ must be parallel and hence J=0J=0 as H2(γ)H^{2}(\gamma) does not contain non-trivial parallel vector fields. But this is a contradiction because we chose JS{0}J\in S\setminus\{0\}. Hence we just proved that

Lε(J),ι(J)ι(J),ι(J)<0\frac{\langle L^{\varepsilon}(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle}<0

for every JS{0}J\in S\setminus\{0\}. As SS is finite dimensional and Lε(J),ι(J)ι(J),ι(J)\frac{\langle L^{\varepsilon}(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle} is invariant under rescaling of JJ, the compactness of the unit ball in SS implies that there exists c(ε)>0c(\varepsilon)>0 such that

Lε(J),ι(J)ι(J),ι(J)c(ε)\frac{\langle L^{\varepsilon}(J),\iota(J)\rangle}{\langle\iota(J),\iota(J)\rangle}\leq-c(\varepsilon)

for every JS{0}J\in S\setminus\{0\}. By the min-max characterization of the eigenvalues for LεL^{\varepsilon}, we see that λ1ελ2ελQεc(ε)<0\lambda_{1}^{\varepsilon}\leq\lambda_{2}^{\varepsilon}\leq...\leq\lambda^{\varepsilon}_{Q}\leq c(\varepsilon)<0. If we also choose ε\varepsilon sufficiently small so that λQ+1ε>0\lambda_{Q+1}^{\varepsilon}>0, we get that for 0<ε<ε(γ)0<\varepsilon<\varepsilon(\gamma), γ\gamma is nondegenerate with respect to gεg_{\varepsilon}. Taking 0<ε<min{ε(γi):1ik}0<\varepsilon<\min\{\varepsilon(\gamma_{i}):1\leq i\leq k\} such that gε𝒰g_{\varepsilon}\in\mathcal{U} and defining g:=gεg^{\prime}:=g_{\varepsilon} we get the desired result.

Lemma 4.6.

Given η>0\eta>0 and NN\in\mathbb{N}, there exists ε>0\varepsilon>0 depending on η\eta and NN such that the following is true: for any Lipschitz function f:INf:I^{N}\rightarrow\mathbb{R} satisfying

|f(x)f(y)|2ε|f(x)-f(y)|\leq 2\varepsilon

for every x,yINx,y\in I^{N}, and for any subset 𝒜\mathcal{A} of INI^{N} of full measure, there exist N+1N+1 sequences of points {y1,m}m,,{yN+1,m}m\{y_{1,m}\}_{m},\cdots,\{y_{N+1,m}\}_{m} contained in 𝒜\mathcal{A} and converging to a common limit y(1,1)Ny\in(-1,1)^{N} such that:

  • f is differentiable at each yi,my_{i,m},

  • the gradients f(yi,m)\nabla f(y_{i,m}) converge to N+1N+1 vectors v1,,vN+1v_{1},\cdots,v_{N+1} with

    dN(0,Conv(v1,,vN+1))<η.d_{\mathbb{R}^{N}}(0,\text{Conv}(v_{1},\cdots,v_{N+1}))<\eta.
Proof.

See [18, Lemma 3]. ∎

5. Proof of the Main Theorem

Fix an nn-dimensional closed manifold MM. We are going to consider several choices and constructions over MM. Let gg be a CC^{\infty} Riemannian metric on MM. Let ε1>0\varepsilon_{1}>0 be a positive constant such that ε1<inj(M,g)\varepsilon_{1}<\operatorname{inj}(M,g), where inj(M,g)\operatorname{inj}(M,g) is the injectivity radius of (M,g)(M,g). Let KK be an integer and B^1,,B^K\hat{B}_{1},...,\hat{B}_{K} be disjoint domains in MM, with piecewise smooth boundary, such that the union of their closures covers MM. Let B1,,BKB_{1},...,B_{K} be some open neighbourhoods of B^1,,B^K\hat{B}_{1},...,\hat{B}_{K} respectively with the property that each of them is contained in a geodesic ball of radius of ε1\varepsilon_{1}. Denote q\mathcal{M}^{q} the space of all CqC^{q} Riemannian metrics on MM. For each 1kK1\leq k\leq K, we define a smooth function 0ϕk10\leq\phi_{k}\leq 1, spt(ϕk)Bk\text{spt}(\phi_{k})\subseteq B_{k} such that

ϕk={1 on B^k0 on Bkc.\phi_{k}=\begin{cases}\text{$1$ on $\hat{B}_{k}$}\\ \text{$0$ on $B_{k}^{c}$}\end{cases}.

Consider also the partition of unity ψk=ϕkl=1Kϕl\psi_{k}=\frac{\phi_{k}}{\sum_{l=1}^{K}\phi_{l}}. We denote

𝒞g,K~,ε1:={(K,{B^k},{Bk},{ϕk})}\mathcal{C}_{g,\tilde{K},\varepsilon_{1}}:=\{(K,\{\hat{B}_{k}\},\{B_{k}\},\{\phi_{k}\})\}

the set of all possible choices as above with KK~K\geq\tilde{K}. Notice that 𝒞g,K~,ε1\mathcal{C}_{g,\tilde{K},\varepsilon_{1}} is non-empty, as we can always find a sufficiently fine triangulation of (M,g)(M,g). We claim that the following property holds:

Proposition 5.1.

Assume that the Weyl law for 11-cycles in nn-manifolds holds as stated in Conjecture 1.4. Then for any metric gg\in\mathcal{M}^{\infty}, for every ε1>0\varepsilon_{1}>0, K~>0\tilde{K}>0 and any choice of

S=(K,{B^k},{Bk},{ϕk})𝒞g,K~,ε1S=(K,\{\hat{B}_{k}\},\{B_{k}\},\{\phi_{k}\})\in\mathcal{C}_{g,\tilde{K},\varepsilon_{1}}

there is a metric g~\tilde{g}\in\mathcal{M}^{\infty} arbitrarily close to gg in the CC^{\infty} topology such that the following holds: there are stationary geodesic networks γ1,,γJ\gamma_{1},...,\gamma_{J} with respect to g~\tilde{g} whose connected components are nondegenerate (according to Definition 2.18) and coefficients α1,,αJ[0,1]\alpha_{1},...,\alpha_{J}\in[0,1] with j=1Jαj=1\sum_{j=1}^{J}\alpha_{j}=1 satisfying

(8) |j=1JαjγjψkdLg~MψkdVolg~|<ε1K\Big{|}\sum_{j=1}^{J}\alpha_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{j}}\psi_{k}dL_{\tilde{g}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}d\operatorname{Vol}_{\tilde{g}}\Big{|}<\frac{\varepsilon_{1}}{K}

for every k=1,,Kk=1,...,K.

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 f,g:(δ,δ)Kf,g:(-\delta,\delta)^{K}\to\mathbb{R} are ε\varepsilon-close if

1δfδ1δgδ<ε\|\frac{1}{\delta}f_{\delta}-\frac{1}{\delta}g_{\delta}\|_{\infty}<\varepsilon

where fδ,gδ:(1,1)Kf_{\delta},g_{\delta}:(-1,1)^{K}\to\mathbb{R} are given by fδ(s)=f(δs)f_{\delta}(s)=f(\delta s) and gδ(s)=g(δs)g_{\delta}(s)=g(\delta s).

Remark 5.3.

Observe that 1δfδ\frac{1}{\delta}f_{\delta} is differentiable at s(1,1)Ks\in(-1,1)^{K} if and only if ff is differentiable at δs(δ,δ)K\delta s\in(-\delta,\delta)^{K} and in that case (1δfδ)(s)=f(δs)\nabla(\frac{1}{\delta}f_{\delta})(s)=\nabla f(\delta s).

Proof of Proposition 5.1.

Let gg\in\mathcal{M}^{\infty}, K~\tilde{K}\in\mathbb{N} and ε1>0\varepsilon_{1}>0. Fix (K,{B^k},{Bk},{ϕk})𝒞g,K~,ε1(K,\{\hat{B}_{k}\},\{B_{k}\},\{\phi_{k}\})\in\mathcal{C}_{g,\tilde{K},\varepsilon_{1}}. Let 𝒰\mathcal{U} be a CC^{\infty} neighborhood of gg. Choose ε0>0\varepsilon^{\prime}_{0}>0 sufficiently small and qK+3q\geq K+3 sufficiently large so that if gg^{\prime}\in\mathcal{M}^{\infty} satisfies ggCq<ε0\|g-g^{\prime}\|_{C^{q}}<\varepsilon^{\prime}_{0}, then g𝒰g^{\prime}\in\mathcal{U}. Let εε0\varepsilon^{\prime}\leq\varepsilon^{\prime}_{0} be a positive real number (which we will have to shrink later in the argument). Our goal is to show that there exists g~\tilde{g}\in\mathcal{M}^{\infty} such that g~gCq<ε0\|\tilde{g}-g\|_{C^{q}}<\varepsilon^{\prime}_{0} and (8) holds for some stationary geodesic nets γ1,,γJ\gamma_{1},...,\gamma_{J} (whose connected components are nondegenerate) with respect to g~\tilde{g} and some coefficients α1,,αJ\alpha_{1},...,\alpha_{J}.

Consider the following KK-parameter family of metrics. For a t=(t1,,tK)(1,1)Kt=(t_{1},...,t_{K})\in(-1,1)^{K}, we define

g^(t)=e2ktkψkg.\hat{g}(t)=e^{2\sum_{k}t_{k}\psi_{k}}g.

At t=0t=0, for each kk, we have

tk|t=0Vol(M,g^(t))\displaystyle\frac{\partial}{\partial t_{k}}\big{|}_{t=0}\operatorname{Vol}(M,\hat{g}(t)) =tk|t=0M(e2ktkψk(x))n2dVolg\displaystyle=\frac{\partial}{\partial t_{k}}\big{|}_{t=0}\int_{M}(e^{2\sum_{k}t_{k}\psi_{k}(x)})^{\frac{n}{2}}\operatorname{dVol}_{g}
=Mnψk(x)dVolg.\displaystyle=\int_{M}n\psi_{k}(x)\operatorname{dVol}_{g}.

As tt goes to zero, we have the following expansion

(9) Vol(M,g^(t))1n=Vol(M,g)1n+k=1KtkVol(M,g)n1nMψk(x)dVolg+R(t)\operatorname{Vol}(M,\hat{g}(t))^{\frac{1}{n}}=\operatorname{Vol}(M,g)^{\frac{1}{n}}+\sum_{k=1}^{K}t_{k}\operatorname{Vol}(M,g)^{-\frac{n-1}{n}}\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}+R(t)

where |R(t)|C1t2|R(t)|\leq C_{1}\|t\|^{2} if t(1,1)Kt\in(-1,1)^{K}, where C1C_{1} is a constant which depends only on gg (this can be checked by computing the second order partial derivatives of tVol(M,g^(t))1nt\mapsto\operatorname{Vol}(M,\hat{g}(t))^{\frac{1}{n}} and using the fact that enVol(M,g)Vol(M,g^(t))enVol(M,g)e^{-n}\operatorname{Vol}(M,g)\leq\operatorname{Vol}(M,\hat{g}(t))\leq e^{n}\operatorname{Vol}(M,g) as Vol(M,g^(t))=Menktkψk(x)dVolg\operatorname{Vol}(M,\hat{g}(t))=\int_{M}e^{n\sum_{k}t_{k}\psi_{k}(x)}\operatorname{dVol}_{g}). Following [18] we can define the following function

f0(t)=Vol(M,g^(t))1nVol(M,g)1nk=1KtkMψk(x)dVolg.f_{0}(t)=\frac{\operatorname{Vol}(M,\hat{g}(t))^{\frac{1}{n}}}{\operatorname{Vol}(M,g)^{\frac{1}{n}}}-\sum_{k=1}^{K}t_{k}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}.

Because of (9), |f0(t)1|=|R(t)Vol(M,g)1n|C2t2|f_{0}(t)-1|=|\frac{R(t)}{\operatorname{Vol}(M,g)^{\frac{1}{n}}}|\leq C_{2}\|t\|^{2} for every t(1,1)Kt\in(-1,1)^{K}; where C2=C1Vol(M,g)1nC_{2}=\frac{C_{1}}{\operatorname{Vol}(M,g)^{\frac{1}{n}}} depends only on gg (as C1C_{1} and other constants CiC_{i} to be defined later).

By the previous, f0f_{0} is C2εC_{2}\varepsilon^{\prime}-close to 11 in (δ,δ)K(-\delta,\delta)^{K} if δ<ε\delta<\varepsilon^{\prime} (see Definition 5.2). Let δ<ε\delta<\varepsilon^{\prime} be such that g^:(δ,δ)Kq\hat{g}:(-\delta,\delta)^{K}\to\mathcal{M}^{q} is an embedding and g^(t)gCq<ε2\|\hat{g}(t)-g\|_{C^{q}}<\frac{\varepsilon^{\prime}}{2} for every t(δ,δ)Kt\in(-\delta,\delta)^{K}. We can slightly perturb g^\hat{g} in the CC^{\infty} topology to another embedding g:(δ,δ)Kqg^{\prime}:(-\delta,\delta)^{K}\to\mathcal{M}^{q} applying Proposition 4.1. We can assume g(t)g^(t)Cq<ε2\|g^{\prime}(t)-\hat{g}(t)\|_{C^{q}}<\frac{\varepsilon^{\prime}}{2} and gvg^vCq<ε\|\frac{\partial g^{\prime}}{\partial v}-\frac{\partial\hat{g}}{\partial v}\|_{C^{q}}<\varepsilon^{\prime} for every t(δ,δ)Kt\in(-\delta,\delta)^{K} and vK:|v|=1v\in\mathbb{R}^{K}:|v|=1. Consider the function

f1(t)=Vol(M,g(t))1nVol(M,g)1nktkMψk(x)dVolg.f_{1}(t)=\frac{\operatorname{Vol}(M,g^{\prime}(t))^{\frac{1}{n}}}{\operatorname{Vol}(M,g)^{\frac{1}{n}}}-\sum_{k}t_{k}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}.

By the properties of gg^{\prime}, there exists C3>0C_{3}>0 such that f1f_{1} is C3εC_{3}\varepsilon^{\prime}-close to the constant function equal to 11 on (δ,δ)K(-\delta,\delta)^{K}.

Now we will use the assumption that the Weyl law for 11-cycles in nn-manifolds holds, which means that

(10) limpωp1(Mn,g)pn1n=α(n,1)Vol(Mn,g)1n.\lim_{p\rightarrow\infty}\omega^{1}_{p}(M^{n},g)p^{-\frac{n-1}{n}}=\alpha(n,1)\operatorname{Vol}(M^{n},g)^{\frac{1}{n}}.

The normalized pp-widths pn1nωp1(g(t))p^{-\frac{n-1}{n}}\omega_{p}^{1}(g^{\prime}(t)) are uniformly Lipschitz continuous on (δ,δ)K(-\delta,\delta)^{K} by [15, Lemma 3.4]. Hence, by (10) the sequence of functions tpn1nωp1(M,g(t))t\mapsto p^{-\frac{n-1}{n}}\omega^{1}_{p}(M,g^{\prime}(t)) converges uniformly to the function ta(n)Vol(M,g(t))1nt\mapsto a(n)\operatorname{Vol}(M,g^{\prime}(t))^{\frac{1}{n}}. This implies that for the previously defined δ>0\delta>0, there exists p0p_{0}\in\mathbb{N} such that pp0p\geq p_{0} implies

|pn1nωp1(M,g(t))α(n,1)Vol(M,g(t))1n|<δε|p^{-\frac{n-1}{n}}\omega_{p}^{1}(M,g^{\prime}(t))-\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(t))^{\frac{1}{n}}|<\delta\varepsilon^{\prime}

and hence

|ωp1(M,g(t))α(n,1)pn1nVol(M,g)1nVol(M,g(t))1nVol(M,g)1n|<C4δε|\frac{\omega^{1}_{p}(M,g^{\prime}(t))}{\alpha(n,1)p^{\frac{n-1}{n}}\operatorname{Vol}(M,g)^{\frac{1}{n}}}-\frac{\operatorname{Vol}(M,g^{\prime}(t))^{\frac{1}{n}}}{\operatorname{Vol}(M,g)^{\frac{1}{n}}}|<C_{4}\delta\varepsilon^{\prime}

for every t(δ,δ)Kt\in(-\delta,\delta)^{K}. The previous means that h(t)=ωp1(M,g(t))α(n,1)pn1nVol(M,g)1nVol(M,g(t))1nVol(M,g)1nh(t)=\frac{\omega^{1}_{p}(M,g^{\prime}(t))}{\alpha(n,1)p^{\frac{n-1}{n}}\operatorname{Vol}(M,g)^{\frac{1}{n}}}-\frac{\operatorname{Vol}(M,g^{\prime}(t))^{\frac{1}{n}}}{\operatorname{Vol}(M,g)^{\frac{1}{n}}} is C4εC_{4}\varepsilon^{\prime}-close to 0 in (δ,δ)K(-\delta,\delta)^{K} and therefore as f1f_{1} is C3εC_{3}\varepsilon^{\prime}-close to 11, by triangle inequality we have that

f2(t)=ωp1(M,g(t))α(n,1)pn1nVol(M,g)1nk=1KtkMψk(x)dVolgf_{2}(t)=\frac{\omega^{1}_{p}(M,g^{\prime}(t))}{\alpha(n,1)p^{\frac{n-1}{n}}\operatorname{Vol}(M,g)^{\frac{1}{n}}}-\sum_{k=1}^{K}t_{k}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}

is C5εC_{5}\varepsilon^{\prime}-close to 11 if pp0p\geq p_{0}, for some C5>0C_{5}>0.

On the other hand, by our choice of gg^{\prime} using Proposition 4.1, there exists a full measure subset 𝒜(δ,δ)K\mathcal{A}\subseteq(-\delta,\delta)^{K} such that for each t𝒜t\in\mathcal{A} and pp\in\mathbb{N} the map tωp1(g(t))t\mapsto\omega^{1}_{p}(g^{\prime}(t)) is differentiable at tt and there exists a stationary geodesic net γp(t)\gamma_{p}(t) with respect to g(t)g^{\prime}(t) so that

  1. (1)

    ωp1(g(t))=Lg(t)(γp(t))\omega_{p}^{1}(g^{\prime}(t))=\operatorname{L}_{g^{\prime}(t)}(\gamma_{p}(t))

  2. (2)

    v(ωp1g(s))|s=t=12γp(t)trγp(t),g(t)gv(t)dLg(t)\frac{\partial}{\partial v}(\omega_{p}^{1}\circ g^{\prime}(s))|_{s=t}=\frac{1}{2}\int_{\gamma_{p}(t)}\operatorname{tr}_{\gamma_{p}(t),g^{\prime}(t)}\frac{\partial g^{\prime}}{\partial v}(t)\operatorname{dL}_{g^{\prime}(t)}.

Define f3:(1,1)Kf_{3}:(-1,1)^{K}\to\mathbb{R} as f3(t)=1δf2(δt)f_{3}(t)=\frac{1}{\delta}f_{2}(\delta t). We know that f31,(1,1)K<C5ε\|f_{3}-1\|_{\infty,(-1,1)^{K}}<C_{5}\varepsilon^{\prime}. Now we want to use Lemma 4.6. In order to do that we will need to impose more restrictions on ε\varepsilon^{\prime}. Let η>0\eta>0. Let ε>0\varepsilon>0 be the one depending on η\eta and N=KN=K according to Lemma 4.6. Choose ε\varepsilon^{\prime} small enough so that C5ε<εC_{5}\varepsilon^{\prime}<\varepsilon, ε<η\varepsilon^{\prime}<\eta and εε0\varepsilon^{\prime}\leq\varepsilon^{\prime}_{0}. Observe that this allows us to define δ>0\delta>0 and p0p_{0}\in\mathbb{N} with all the properties in the construction above. Then we have

|f3(x)f3(y)|2ε|f_{3}(x)-f_{3}(y)|\leq 2\varepsilon

for every x,y(1,1)Kx,y\in(-1,1)^{K}. As f3f_{3} is Lipschitz, we can apply Lemma 4.6 to f3f_{3} and the full measure subset 𝒜={tδ:t𝒜}\mathcal{A^{\prime}}=\{\frac{t}{\delta}:t\in\mathcal{A}\}. After passing to (δ,δ)K(-\delta,\delta)^{K} by rescaling and using Remark 5.3, we get K+1K+1 sequences of points {s1,m}m,,{sK+1,m}m\{s_{1,m}\}_{m},...,\{s_{K+1,m}\}_{m\in\mathbb{N}} contained in 𝒜\mathcal{A} and converging to a common limit s(δ,δ)Ks\in(-\delta,\delta)^{K} such that:

  1. (1)

    f2f_{2} is differentiable at each sj,ms_{j,m}.

  2. (2)

    The gradients f2(sj,m)\nabla f_{2}(s_{j,m}) converge to K+1K+1 vectors v1,,vK+1v_{1},...,v_{K+1} with

    dN(0,Conv(v1,,vK+1))<η.d_{\mathbb{R}^{N}}(0,\text{Conv}(v_{1},...,v_{K+1}))<\eta.

Let α1,,αK+1[0,1]\alpha_{1},...,\alpha_{K+1}\in[0,1] be such that j=1K+1αj=1\sum_{j=1}^{K+1}\alpha_{j}=1 and |j=1K+1αjvj|<η|\sum_{j=1}^{K+1}\alpha_{j}v_{j}|<\eta. Then if mm is sufficiently large,

|j=1K+1αjf2(sj,m)|<η|\sum_{j=1}^{K+1}\alpha_{j}\nabla f_{2}(s_{j,m})|<\eta

and hence

|j=1K+1αjf2tk(sj,m)|<η|\sum_{j=1}^{K+1}\alpha_{j}\frac{\partial f_{2}}{\partial t_{k}}(s_{j,m})|<\eta

for every k=1,,Kk=1,...,K. But using the definition of f2f_{2} and denoting γj,m=γp(sj,m)\gamma_{j,m}=\gamma_{p}(s_{j,m}),

f2tk(sj,m)\displaystyle\frac{\partial f_{2}}{\partial t_{k}}(s_{j,m}) =tkωp1(M,g(s))|s=sj,mα(n,1)Vol(M,g)1npn1nMψk(x)dVolg\displaystyle=\frac{\frac{\partial}{\partial t_{k}}\omega_{p}^{1}(M,g^{\prime}(s))|_{s=s_{j,m}}}{\alpha(n,1)\operatorname{Vol}(M,g)^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}
=γj,mtrγj,m,g(sj,m)gtk(sj,m)dLg(sj,m)2α(n,1)Vol(M,g)1npn1nMψk(x)dVolg.\displaystyle=\frac{\int_{\gamma_{j,m}}\operatorname{tr}_{\gamma_{j,m},g^{\prime}(s_{j,m})}\frac{\partial g^{\prime}}{\partial t_{k}}(s_{j,m})\operatorname{dL}_{g^{\prime}(s_{j,m})}}{2\alpha(n,1)\operatorname{Vol}(M,g)^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}.

As the lengths Lg(sj,m)(γj,m)=ωp(g(sj,m))\operatorname{L}_{g^{\prime}(s_{j,m})}(\gamma_{j,m})=\omega_{p}(g^{\prime}(s_{j,m})) of the γj,m\gamma_{j,m}’s are uniformly bounded, passing to a subsequence we can obtain stationary geodesic networks γ1,,γK+1\gamma_{1},...,\gamma_{K+1} with respect to g(s)g^{\prime}(s) verifying

(11) limmγj,m=γj\lim_{m\to\infty}\gamma_{j,m}=\gamma_{j}

in the varifold topology for every j=1,,K+1j=1,...,K+1. Hence from the previous,

|j=1K+1αjγjtrγj,g(s)gtk(s)dLg(s)2α(n,1)Vol(M,g)1npn1nMψk(x)dVolg|η|\sum_{j=1}^{K+1}\alpha_{j}\frac{\int_{\gamma_{j}}\operatorname{tr}_{\gamma_{j},g^{\prime}(s)}\frac{\partial g^{\prime}}{\partial t_{k}}(s)\operatorname{dL}_{g^{\prime}(s)}}{2\alpha(n,1)\operatorname{Vol}(M,g)^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g}|\leq\eta

for every k=1,,Kk=1,...,K. Using that g^(t)gCq<ε2\|\hat{g}(t)-g\|_{C^{q}}<\frac{\varepsilon^{\prime}}{2}, g(t)g^(t)Cq<ε2\|g^{\prime}(t)-\hat{g}(t)\|_{C^{q}}<\frac{\varepsilon^{\prime}}{2} and gvg^vCq<ε\|\frac{\partial g^{\prime}}{\partial v}-\frac{\partial\hat{g}}{\partial v}\|_{C^{q}}<\varepsilon^{\prime} for every t(δ,δ)Kt\in(-\delta,\delta)^{K} and vK:|v|=1v\in\mathbb{R}^{K}:|v|=1; and the fact that ε<η\varepsilon^{\prime}<\eta, we can see that there exists a constant C6>0C_{6}>0 such that

|j=1K+1αjγjtrγj,g^(s)g^tk(s)dLg^(s)2α(n,1)Vol(M,g(s))1npn1nMψk(x)dVolg(s)|C6η.|\sum_{j=1}^{K+1}\alpha_{j}\frac{\int_{\gamma_{j}}{\operatorname{tr}_{\gamma_{j},\hat{g}(s)}\frac{\partial\hat{g}}{\partial t_{k}}(s)}\operatorname{dL}_{\hat{g}(s)}}{2\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(s))^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g^{\prime}(s)}|\leq C_{6}\eta.

By definition of g^\hat{g}, g^tk(s)=2ψkg^(s)\frac{\partial\hat{g}}{\partial t_{k}}(s)=2\psi_{k}\hat{g}(s) thus

(12) |j=1K+1αjγjψkdLg^(s)α(n,1)Vol(M,g(s))1npn1nMψk(x)dVolg(s)|C6η.\displaystyle|\sum_{j=1}^{K+1}\alpha_{j}\frac{\int_{\gamma_{j}}\psi_{k}\operatorname{dL}_{\hat{g}(s)}}{\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(s))^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g^{\prime}(s)}|\leq C_{6}\eta.

Combining (12) with the fact that g(s)g^(s)Cq<ε2\|g^{\prime}(s)-\hat{g}(s)\|_{C^{q}}<\frac{\varepsilon^{\prime}}{2},

(13) |j=1K+1αjγjψkdLg(s)α(n,1)Vol(M,g(s))1npn1nMψk(x)dVolg(s)|C7η.|\sum_{j=1}^{K+1}\alpha_{j}\frac{\int_{\gamma_{j}}\psi_{k}\operatorname{dL}_{g^{\prime}(s)}}{\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(s))^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{g^{\prime}(s)}|\leq C_{7}\eta.

But we know that Łg(s)(γj)=ωp1(g(s))\L_{g^{\prime}(s)}(\gamma_{j})=\omega_{p}^{1}(g^{\prime}(s)) for every j=1,,K+1j=1,...,K+1, so

|γjψkdLg(s)α(n,1)Vol(M,g(s))1npn1nγjψkdLg(s)|\displaystyle|\frac{\int_{\gamma_{j}}\psi_{k}\operatorname{dL}_{g^{\prime}(s)}}{\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(s))^{\frac{1}{n}}p^{\frac{n-1}{n}}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{j}}\psi_{k}\operatorname{dL}_{g^{\prime}(s)}| =\displaystyle=
|γjψkdLg(s)||ωp1(g(s))α(n,1)Vol(M,g(s))1npn1n1|\displaystyle|\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{j}}\psi_{k}\operatorname{dL}_{g^{\prime}(s)}||\frac{\omega_{p}^{1}(g^{\prime}(s))}{\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(s))^{\frac{1}{n}}p^{\frac{n-1}{n}}}-1| \displaystyle\leq
|ωp1(g(s))α(n,1)Vol(M,g(s))1npn1n1|\displaystyle|\frac{\omega_{p}^{1}(g^{\prime}(s))}{\alpha(n,1)\operatorname{Vol}(M,g^{\prime}(s))^{\frac{1}{n}}p^{\frac{n-1}{n}}}-1| η\displaystyle\leq\eta

if pp1p\geq p_{1} for some p1p_{1}\in\mathbb{N}, because of the Weyl law and the fact 0ψk10\leq\psi_{k}\leq 1. Hence from (13),

|j=1K+1αjγjψkdLg(s)MψkdVolg(s)|C8η|\sum_{j=1}^{K+1}\alpha_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{j}}\psi_{k}\operatorname{dL}_{g^{\prime}(s)}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}\operatorname{dVol}_{g^{\prime}(s)}|\leq C_{8}\eta

for some constant C8C_{8} depending only on gg. Let us take η=ε12C8K\eta=\frac{\varepsilon_{1}}{2C_{8}K} and pmax{p0,p1}p\geq\max\{p_{0},p_{1}\}.

Notice that g(s)gCqg(s)g^(s)Cq+g^(s)gCq<ε2+ε2<ε0\|g^{\prime}(s)-g\|_{C^{q}}\leq\|g^{\prime}(s)-\hat{g}(s)\|_{C^{q}}+\|\hat{g}(s)-g\|_{C^{q}}<\frac{\varepsilon^{\prime}}{2}+\frac{\varepsilon^{\prime}}{2}<\varepsilon_{0}^{\prime}. Let us represent each γi\gamma_{i} as a map γi:ΓiM\gamma_{i}:\Gamma_{i}\to M where each connected component of the weighted multigraph Γi\Gamma_{i} is good and the restrictions of γi\gamma_{i} to those connected components are embedded (here we are using Remark 2.17). The metric g(s)g^{\prime}(s) has all the properties required by Proposition 5.1 except that the components of the γi\gamma_{i}’s may not be non-degenerate and may not be CC^{\infty} (in principle they are only CqC^{q}). Using Proposition 4.5, we can change g(s)g^{\prime}(s) for another CqC^{q} metric g¯\overline{g} which still verifies g¯gCq<ε0\|\overline{g}-g\|_{C^{q}}<\varepsilon_{0}^{\prime}, and has the property that γ1,,γK+1\gamma_{1},...,\gamma_{K+1} are non-degenerate stationary geodesic nets with respect to g¯\overline{g}. If on top of that we apply the Implicit Function Theorem (see [15, Lemma 2.6]), we can find a CC^{\infty} metric g~\tilde{g} close enough to g¯\overline{g} in the CqC^{q} topology so that g~gCq<ε0\|\tilde{g}-g\|_{C^{q}}<\varepsilon_{0} which admits stationary geodesic networks γ~1,,γ~k+1\tilde{\gamma}_{1},...,\tilde{\gamma}_{k+1} whose connected components are nondegenerate and verify

|j=1K+1αjγ~jψkdLg~MψkdVolg~|<ε1K|\sum_{j=1}^{K+1}\alpha_{j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\tilde{\gamma}_{j}}\psi_{k}\operatorname{dL}_{\tilde{g}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}\operatorname{dVol}_{\tilde{g}}|<\frac{\varepsilon_{1}}{K}

for every k=1,,K+1k=1,...,K+1. This completes the proof.

Now we will show that Proposition 5.1 implies Theorem 1.5. Given gg\in\mathcal{M}^{\infty}, ε1>0\varepsilon_{1}>0, K~\tilde{K}\in\mathbb{N} and S𝒞g,K~,ε1S\in\mathcal{C}_{g,\tilde{K},\varepsilon_{1}} we will denote (g,ε1,K~,S)\mathcal{M}(g,\varepsilon_{1},\tilde{K},S) the set of all metrics g~\tilde{g}\in\mathcal{M}^{\infty} such that g~gCq<ε1\|\tilde{g}-g\|_{C^{q}}<\varepsilon_{1} (computed with respect to gg) and there exist stationary geodesic networks γ1,,γJ\gamma_{1},...,\gamma_{J} with respect to g~\tilde{g} whose connected components are nondegenerate (according to Definition 2.18) and coefficients α1,,αJ[0,1]\alpha_{1},...,\alpha_{J}\in[0,1] with j=1Jαj=1\sum_{j=1}^{J}\alpha_{j}=1 such that (8) holds for every k=1,,Kk=1,...,K. By the Implicit Function Theorem, (g,ε1,K~,S)\mathcal{M}(g,\varepsilon_{1},\tilde{K},S) is open (see [15, Lemma 2.6]). Therefore given ε1>0\varepsilon_{1}>0 and K~\tilde{K}\in\mathbb{N} the set

(ε1,K~)=gS𝒞g,ε1,K~(g,ε1,K,S)\mathcal{M}(\varepsilon_{1},\tilde{K})=\bigcup_{g\in\mathcal{M}^{\infty}}\bigcup_{S\in\mathcal{C}_{g,\varepsilon_{1},\tilde{K}}}\mathcal{M}(g,\varepsilon_{1},K,S)

is open and by Proposition 5.1 it is also dense in \mathcal{M}^{\infty}. Define

~=m(1m,m)\tilde{\mathcal{M}}=\bigcap_{m\in\mathbb{N}}\mathcal{M}(\frac{1}{m},m)

which is a generic subset of \mathcal{M}^{\infty} in the Baire sense. We are going to prove that if g~~\tilde{g}\in\tilde{\mathcal{M}} then there exists a sequence of equidistributed stationary geodesic networks with respect to g~\tilde{g}.

Fix g~~\tilde{g}\in\tilde{\mathcal{M}}. By definition, given mm\in\mathbb{N} there exists gg\in\mathcal{M}^{\infty} such that g~(g,1m,m,S)\tilde{g}\in\mathcal{M}(g,\frac{1}{m},m,S) for some S𝒞g,1m,mS\in\mathcal{C}_{g,\frac{1}{m},m}. Therefore, g~\tilde{g} belongs to a 1m\frac{1}{m} neighborhood of gg in the CKC^{K} topology; and there exist J=JmJ=J_{m}\in\mathbb{N}, stationary geodesic networks γm,1,,γm,Jm\gamma_{m,1},...,\gamma_{m,J_{m}} with respect to g~\tilde{g} and coefficients αm,1,,αm,Jm[0,1]\alpha_{m,1},...,\alpha_{m,J_{m}}\in[0,1] with j=1Jmαm,j=1\sum_{j=1}^{J_{m}}\alpha_{m,j}=1 satisfying

(14) |j=1Jmαm,jγm,jψk(x)dLg~Mψk(x)dVolg~|<1mK|\sum_{j=1}^{J_{m}}\alpha_{m,j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{m,j}}\psi_{k}(x)\operatorname{dL}_{\tilde{g}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}\psi_{k}(x)\operatorname{dVol}_{\tilde{g}}|<\frac{1}{mK}

for every k=1,,Kk=1,...,K. Let fC(M,)f\in C^{\infty}(M,\mathbb{R}). We want to obtain a formula analogous to the previous one but replacing ψk\psi_{k} by ff, which will imply the following proposition.

Proposition 5.4.

Let g~~\tilde{g}\in\tilde{\mathcal{M}}. For each mm\in\mathbb{N}, there exists J=JmJ=J_{m} depending on mm, integers {cm,j}1jJm\{c_{m,j}\}_{1\leq j\leq J_{m}} and stationary geodesic networks {γm,j}1jJm\{\gamma_{m,j}\}_{1\leq j\leq J_{m}} such that

|j=1Jmcm,jγm,jfdLg~j=1Jmcm,jLg~(γm,j)MfdVolg~|D(f)m|\frac{\sum_{j=1}^{J_{m}}c_{m,j}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}}{\sum_{j=1}^{J_{m}}c_{m,j}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\tilde{g}}|\leq\frac{D(f)}{m}

for every fC(M,)f\in C^{\infty}(M,\mathbb{R}), where D(f)>0D(f)>0 is a constant depending only on ff and the metric g~\tilde{g}.

Proof.

Given mm\in\mathbb{N}, consider as above gg\in\mathcal{M}^{\infty} and S𝒞g,1m,mS\in\mathcal{C}_{g,\frac{1}{m},m} such that g~(g,1m,m,S)\tilde{g}\in\mathcal{M}(g,\frac{1}{m},m,S). Define JmJ_{m}\in\mathbb{N}, stationary geodesic networks γm,1,,γm,Jm\gamma_{m,1},...,\gamma_{m,J_{m}} with respect to g~\tilde{g} and coefficients αm,1,,αm,Jm\alpha_{m,1},...,\alpha_{m,J_{m}} such that (14) holds. Taking S=(K,{B^k}k,{Bk}k,{ϕk}k)𝒞g,1m,mS=(K,\{\hat{B}_{k}\}_{k},\{B_{k}\}_{k},\{\phi_{k}\}_{k})\in\mathcal{C}_{g,\frac{1}{m},m} into account, let us choose points q1,,qKq_{1},...,q_{K} with qkB^kq_{k}\in\hat{B}_{k} for each k=1,,Kk=1,...,K. The idea will be to approximate the integral of f(x)f(x) by the integral of the function k=1Kf(qk)ψk(x)\sum_{k=1}^{K}f(q_{k})\psi_{k}(x). First of all, by using (14) we can see that

(15) |j=1Jmαm,jγm,j[k=1Kf(qk)ψk(x)]dLg~M[k=1Kf(qk)ψk(x)]dVolg~|<D1m|\sum_{j=1}^{J_{m}}\alpha_{m,j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{m,j}}[\sum_{k=1}^{K}f(q_{k})\psi_{k}(x)]d\operatorname{L}_{\tilde{g}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}[\sum_{k=1}^{K}f(q_{k})\psi_{k}(x)]d\operatorname{Vol}_{\tilde{g}}|<\frac{D_{1}}{m}

where D1=f=max{f(x):xM}D_{1}=\|f\|_{\infty}=\max\{f(x):x\in M\} depends only on ff (and not on mm, gg or SS). On the other hand, given xMx\in M

|f(x)k=1Kf(qk)ψk(x)|\displaystyle|f(x)-\sum_{k=1}^{K}f(q_{k})\psi_{k}(x)| =|f(x)k=1Kψk(x)k=1Kf(qk)ψk(x)|\displaystyle=|f(x)\sum_{k=1}^{K}\psi_{k}(x)-\sum_{k=1}^{K}f(q_{k})\psi_{k}(x)|
=|k=1Kf(x)ψk(x)f(qk)ψk(x)|\displaystyle=|\sum_{k=1}^{K}f(x)\psi_{k}(x)-f(q_{k})\psi_{k}(x)|
k:xBk|f(x)f(qk)||ψk(x)|\displaystyle\leq\sum_{k:x\in B_{k}}|f(x)-f(q_{k})||\psi_{k}(x)|
=k:xBk|g~f(ck)|dg~(x,qk)ψk(x)\displaystyle=\sum_{k:x\in B_{k}}|\nabla_{\tilde{g}}f(c_{k})|d_{\tilde{g}}(x,q_{k})\psi_{k}(x)
2g~fmk=1Kψk(x)\displaystyle\leq\frac{2\|\nabla_{\tilde{g}}f\|_{\infty}}{m}\sum_{k=1}^{K}\psi_{k}(x)
=2g~fm.\displaystyle=\frac{2\|\nabla_{\tilde{g}}f\|_{\infty}}{m}.

We used the Mean Value Theorem and the fact that supp(ψk)Bk\text{supp}(\psi_{k})\subseteq B_{k} and diamg~(Bk)2diamg(Bk)2m\text{diam}_{\tilde{g}}(B_{k})\leq 2\text{diam}_{g}(B_{k})\leq\frac{2}{m} for every ii. Combining this and (15) we get

(16) |j=1Jmαm,jγm,jfdLg~MfdVolg~|<D2m|\sum_{j=1}^{J_{m}}\alpha_{m,j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\tilde{g}}|<\frac{D_{2}}{m}

where D2D_{2} depends only on ff and g~\tilde{g}. Let us choose integers cm,j,dmc_{m,j},d_{m}\in\mathbb{N} such that

|αm,jLg~(γm,j)cm,jdm|<1mJmLg~(γm.j).|\frac{\alpha_{m,j}}{\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}-\frac{c_{m,j}}{d_{m}}|<\frac{1}{mJ_{m}\operatorname{L}_{\tilde{g}}(\gamma_{m.j})}.

Then it holds

|j=1Jmαm,jγm,jfdLg~j=1Jmcm,jdmγm,jfdLg~|\displaystyle|\sum_{j=1}^{J_{m}}\alpha_{m,j}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}-\sum_{j=1}^{J_{m}}\frac{c_{m,j}}{d_{m}}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}| j=1Jm|αm,jLg~(γm.j)cmjdm||γm,jfdLg~|\displaystyle\leq\sum_{j=1}^{J_{m}}|\frac{\alpha_{m,j}}{\operatorname{L}_{\tilde{g}(\gamma_{m.j})}}-\frac{c_{m_{j}}}{d_{m}}||\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}|
j=1Jm1mJmLg~(γm,j)fLg~(γm,j)\displaystyle\leq\sum_{j=1}^{J_{m}}\frac{1}{mJ_{m}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}\|f\|_{\infty}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})
=D1m\displaystyle=\frac{D_{1}}{m}

and hence by (16) and triangle inequality we get

|j=1Kmcm,jdmγm,jfdLg~MfdVolg~|<D3m|\sum_{j=1}^{K_{m}}\frac{c_{m,j}}{d_{m}}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\tilde{g}}|<\frac{D_{3}}{m}

where D3=D2+D1D_{3}=D_{2}+D_{1} depends only on ff and g~\tilde{g}. On the other hand,

|j=1Jmcm,jdmγm,jfdLg~j=1Jmcm,jγm,jfdLg~j=1Jmcm,jLg~(γm,j)|\displaystyle|\sum_{j=1}^{J_{m}}\frac{c_{m,j}}{d_{m}}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}-\frac{\sum_{j=1}^{J_{m}}c_{m,j}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}}{\sum_{j=1}^{J_{m}}c_{m,j}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}| \displaystyle\leq
|1dm1j=1Jmcm,jLg~(γm,j)||j=1Jmcm,jγm,jfdLg~|\displaystyle|\frac{1}{d_{m}}-\frac{1}{\sum_{j=1}^{J_{m}}c_{m,j}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}||\sum_{j=1}^{J_{m}}c_{m,j}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}| \displaystyle\leq
|1dm1j=1Jmcm,jLg~(γm,j)|j=1Jmcm,jfLg~(γm,j)\displaystyle|\frac{1}{d_{m}}-\frac{1}{\sum_{j=1}^{J_{m}}c_{m,j}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}|\sum_{j=1}^{J_{m}}c_{m,j}\|f\|_{\infty}\operatorname{L}_{\tilde{g}}(\gamma_{m,j}) =\displaystyle=
D1|j=1Jmcm,jdmLg~(γm,j)1|\displaystyle D_{1}|\sum_{j=1}^{J_{m}}\frac{c_{m,j}}{d_{m}}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})-1| D1m\displaystyle\leq\frac{D_{1}}{m}

because |j=1Jmcm,jdmLg~(γm,j)1|<1m|\sum_{j=1}^{J_{m}}\frac{c_{m,j}}{d_{m}}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})-1|<\frac{1}{m}. Hence

|j=1Jmcm,jγm,jfdLg~j=1Jmcm,jLg~(γm,j)MfdVolg~|D4m|\frac{\sum_{j=1}^{J_{m}}c_{m,j}\int_{\gamma_{m,j}}f\operatorname{dL}_{\tilde{g}}}{\sum_{j=1}^{J_{m}}c_{m,j}\operatorname{L}_{\tilde{g}}(\gamma_{m,j})}-\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\tilde{g}}|\leq\frac{D_{4}}{m}

for a constant D4D_{4} depending only on ff and g~\tilde{g}, as desired. ∎

Given g~~\tilde{g}\in\tilde{\mathcal{M}}, using Proposition 5.4 we can find a sequence of finite lists of connected embedded stationary geodesic nets {βm,1,,βm,Km}m\{\beta_{m,1},...,\beta_{m,K_{m}}\}_{m\in\mathbb{N}} with respect to g~\tilde{g} satisfying the following: given fC(M,)f\in C^{\infty}(M,\mathbb{R}), if we denote Xm,j=βm,jfdLg~X_{m,j}=\int_{\beta_{m,j}}f\operatorname{dL}_{\tilde{g}} and X¯m,j=Lg~(βm,j)\bar{X}_{m,j}=\operatorname{L}_{\tilde{g}}(\beta_{m,j}), then

(17) |j=1KmXm,jj=1KmX¯m,jα|D(f)m|\frac{\sum_{j=1}^{K_{m}}X_{m,j}}{\sum_{j=1}^{K_{m}}\bar{X}_{m,j}}-\alpha|\leq\frac{D(f)}{m}

where α=MfdVolg~\alpha=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\tilde{g}} and D(f)D(f) is a constant depending only on ff. The lists {βm,j}1jKm\{\beta_{m,j}\}_{1\leq j\leq K_{m}} are obtained from the lists {γm,j}1jJm\{\gamma_{m,j}\}_{1\leq j\leq J_{m}} and the coefficients {cm,j}1jJm\{c_{m,j}\}_{1\leq j\leq J_{m}} from Proposition 5.4 by decomposing each γm,j\gamma_{m,j} as a union of embedded stationary geodesic networks whose domain is a good weighted multigraph (see Remark 2.17) and listing each of them cm,jc_{m,j} times. From the Xm,jX_{m,j}’s and the X¯m,j\bar{X}_{m,j}’s, we want to construct two sequences {Yi}i,{Y¯i}i\{Y_{i}\}_{i\in\mathbb{N}},\{\bar{Y}_{i}\}_{i\in\mathbb{N}} such that

  • For all ii, there exist integers m(i),j(i)m(i),j(i) (chosen independently of ff) with Yi=Xm(i),j(i)Y_{i}=X_{m(i),j(i)} and Y¯i=X¯m(i),j(i)\bar{Y}_{i}=\bar{X}_{m(i),j(i)},

  • It holds

    limki=1kYii=1kY¯i=α.\lim_{k\rightarrow\infty}\frac{\sum_{i=1}^{k}Y_{i}}{\sum_{i=1}^{k}\bar{Y}_{i}}=\alpha.

This can be done as in [18, p. 437-439] and gives us a sequence {γi}i\{\gamma_{i}\}_{i\in\mathbb{N}} of connected embedded stationary geodesic networks with respect to g~\tilde{g} (defined as γi=βm(i),j(i)\gamma_{i}=\beta_{m(i),j(i)}), which is constructed independently of the constant D(f)D(f). It holds

limki=1kγifdLg~i=1kLg~(γi)=MfdVolg~\lim_{k\to\infty}\frac{\sum_{i=1}^{k}\int_{\gamma_{i}}f\operatorname{dL}_{\tilde{g}}}{\sum_{i=1}^{k}\operatorname{L}_{\tilde{g}}(\gamma_{i})}=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{M}f\operatorname{dVol}_{\tilde{g}}

for every fC(M,)f\in C^{\infty}(M,\mathbb{R}). This gives us the desired equidistribution result and completes the proof of Theorem 1.5.

Remark 5.5.

Observe that combining the Weyl law for 11-cycles in 33-manifolds from [11] with Theorem 1.5 which we just proved, we obtain Theorem 1.2.

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 pp-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 MM be a closed 22-manifold. Let g:INqg:I^{N}\to\mathcal{M}^{q} be a smooth embedding, NN\in\mathbb{N}. If qN+3q\geq N+3, there exists an arbitrarily small perturbation in the CC^{\infty} topology g:INqg^{\prime}:I^{N}\to\mathcal{M}^{q} such that there is a full measure subset 𝒜IN\mathcal{A}\subseteq I^{N} with the following property: for any pp\in\mathbb{N} and any t𝒜t\in\mathcal{A}, the function sωp1(g(s))s\mapsto\omega^{1}_{p}(g^{\prime}(s)) is differentiable at tt and there exist almost embedded closed geodesics γp1,,γpP:S1M\gamma_{p}^{1},...,\gamma_{p}^{P}:S^{1}\to M such that the following two conditions hold

  1. (1)

    ωp1(g(t))=i=1PLg(t)(γpi(t))\omega_{p}^{1}(g^{\prime}(t))=\sum_{i=1}^{P}L_{g^{\prime}(t)}(\gamma_{p}^{i}(t)).

  2. (2)

    v(ωp1g)|s=t=12i=1Pγpitrγpi,g(t)gv(t)dLg(t)\frac{\partial}{\partial v}(\omega^{1}_{p}\circ g^{\prime})\big{|}_{s=t}=\frac{1}{2}\sum_{i=1}^{P}\int_{\gamma_{p}^{i}}\operatorname{tr}_{\gamma_{p}^{i},g^{\prime}(t)}\frac{\partial g^{\prime}}{\partial v}(t)\operatorname{dL}_{g^{\prime}(t)}.

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 γ:S1M\gamma:S^{1}\to M we can associate a connected graph Γ=S1/\Gamma=S^{1}/\sim where \sim is the equivalence relation sts\sim t if and only if γ(s)=γ(t)\gamma(s)=\gamma(t). This induces a map f:ΓMf:\Gamma\to M defined as f([t])=γ(t)f([t])=\gamma(t). Observe that the as the self-intersections of γ\gamma are transverse, the vertices of Γ\Gamma are mapped precisely to those self-intersections and the map f:ΓMf:\Gamma\to M is injective. Moreover, Γ\Gamma is a good multigraph and f:Γ(M,g)f:\Gamma\to(M,g) is an embedded stationary geodesic network. We replace the set {Γi}i\{\Gamma_{i}\}_{i\in\mathbb{N}} which in the proof of Proposition 4.1 is the set of all good connected multigraphs by the countable set of pairs 𝒫={(Γ,r)}\mathcal{P}=\{(\Gamma,r)\} where Γ\Gamma is a good multigraph which can be obtained as Γ=S1/\Gamma=S^{1}/\sim from an almost embedded closed geodesic γ:S1(M,g)\gamma:S^{1}\to(M,g) with respect to some metric gg as before and rr is the set of pairs ((E1,i1),(E2,i2))((E_{1},i_{1}),(E_{2},i_{2})) such that πE1(i1)=πE2(i2)\pi_{E_{1}}(i_{1})=\pi_{E_{2}}(i_{2}) and (1)i1+1f˙E1(i1)|f˙E1(i1)|g=(1)i2f˙E2(i2)|f˙E2(i2)|g(-1)^{i_{1}+1}\frac{\dot{f}_{E_{1}}(i_{1})}{|\dot{f}_{E_{1}}(i_{1})|_{g}}=(-1)^{i_{2}}\frac{\dot{f}_{E_{2}}(i_{2})}{|\dot{f}_{E_{2}}(i_{2})|_{g}} (in other words, rr contains the necessary information to reparametrize the geodesic net f:ΓMf:\Gamma\to M to an immersed closed geodesic γ:S1(M,g)\gamma:S^{1}\to(M,g)). Observe that if (Γ,r)𝒫(\Gamma,r)\in\mathcal{P} and f:Γ(M,g)f:\Gamma\to(M,g) is an embedded stationary geodesic network verifying (1)i1+1f˙E1(i1)|f˙E1(i1)|g=(1)i2f˙E2(i2)|f˙E2(i2)|g(-1)^{i_{1}+1}\frac{\dot{f}_{E_{1}}(i_{1})}{|\dot{f}_{E_{1}}(i_{1})|_{g}}=(-1)^{i_{2}}\frac{\dot{f}_{E_{2}}(i_{2})}{|\dot{f}_{E_{2}}(i_{2})|_{g}} for every ((E1,i1),(E2,i2))r((E_{1},i_{1}),(E_{2},i_{2}))\in r then f:Γ(M,g)f:\Gamma\to(M,g) can be reparametrized as an immersed closed geodesic γ:S1(M,g)\gamma:S^{1}\to(M,g) whose self intersections occur precisely at the points {f(v):v vertex of Γ}\{f(v):v\text{ vertex of }\Gamma\}.

Taking the previous into account, instead of the B~Γ,M\tilde{B}_{\Gamma,M} in the proof of Proposition 4.1 we will work with the following. Consider the set of pairs (Γ,r)(\Gamma,r) where Γ\Gamma is a graph, Γ=i=1PΓi\Gamma=\bigcup_{i=1}^{P}\Gamma_{i} as a union of connected components, r=(ri)1iPr=(r_{i})_{1\leq i\leq P} and (Γi,ri)𝒫(\Gamma_{i},r_{i})\in\mathcal{P} for every 1iP1\leq i\leq P. Given such a pair (Γ,r)(\Gamma,r) and a natural number MM\in\mathbb{N} we define Γ,r,M\mathcal{B}_{\Gamma,r,M} to be the set of all t(1,1)Nt\in(-1,1)^{N} such that there exists a stationary geodesic network f:Γ(M,g(t))f:\Gamma\to(M,g^{\prime}(t)) verifying

  1. (1)

    For each 1iP1\leq i\leq P, fi=f|Γif_{i}=f|_{\Gamma_{i}} is an embedding and verifies the relations (1)i1+1f˙i,E1(i1)|f˙i,E1(i1)|g(t)=(1)i2f˙i,E2(i2)|f˙i,E2(i2)|g(t)(-1)^{i_{1}+1}\frac{\dot{f}_{i,E_{1}}(i_{1})}{|\dot{f}_{i,E_{1}}(i_{1})|_{g^{\prime}(t)}}=(-1)^{i_{2}}\frac{\dot{f}_{i,E_{2}}(i_{2})}{|\dot{f}_{i,E_{2}}(i_{2})|_{g^{\prime}(t)}} for every ((E1,i1),(E2,i2))ri((E_{1},i_{1}),(E_{2},i_{2}))\in r_{i}.

  2. (2)

    fi3M\|f_{i}\|_{3}\leq M for every 1iP1\leq i\leq P.

  3. (3)

    F1(g(t),fi)1MF_{1}(g^{\prime}(t),f_{i})\geq\frac{1}{M} for every 1iP1\leq i\leq P.

  4. (4)

    F2(E1,i1),(E2,i2)(g(t),fi)11MF_{2}^{(E_{1},i_{1}),(E_{2},i_{2})}(g^{\prime}(t),f_{i})\leq 1-\frac{1}{M} for every 1iP1\leq i\leq P, and every pair (E1,i1)(E2,i2)i×{0,1}(E_{1},i_{1})\neq(E_{2},i_{2})\in\mathscr{E}_{i}\times\{0,1\} such that πE1(i1)=πE2(i2)\pi_{E_{1}}(i_{1})=\pi_{E_{2}}(i_{2}).

  5. (5)

    d(g(t),fi)E(s)1Md^{E}_{(g^{\prime}(t),f_{i})}(s)\geq\frac{1}{M} for every 1iP1\leq i\leq P, EiE\in\mathscr{E}_{i} and sEs\in E.

  6. (6)

    d(g(t),fi)E,E(s)1Md^{E,E^{\prime}}_{(g^{\prime}(t),f_{i})}(s)\geq\frac{1}{M} for every 1iP1\leq i\leq P, EEiE\neq E^{\prime}\in\mathscr{E}_{i} and sEs\in E.

  7. (7)

    ωp1(g(t))=Lg(t)(f)\omega_{p}^{1}(g^{\prime}(t))=\operatorname{L}_{g^{\prime}(t)}(f).

Therefore, same as in Proposition 4.1 we have IN=Γ,r,MΓ,r,MI^{N}=\bigcup_{\Gamma,r,M}\mathcal{B}_{\Gamma,r,M} because of the fact showed in [7] that the pp-widths on surfaces are realized by unions of almost embedded closed geodesics; and each Γ,r,M\mathcal{B}_{\Gamma,r,M} is closed. The rest of the proof follows exactly as in Proposition 4.1 if we replace the pairs (Γ,M)(\Gamma,M) by the triples (Γ,r,M)(\Gamma,r,M). ∎

One more remark is necessary to adapt the proof of Proposition 5.1. The sequences (γj,m)m(\gamma_{j,m})_{m} in (11) have length uniformly bounded by some L>0L>0 and consist of finite unions of almost embedded closed geodesics. This implies that the number of closed geodesics whose union is γj,m\gamma_{j,m} is also bounded (independently on mm). 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 C0C^{0}. 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 C{C}^{\infty} closing lemma for Hamiltonian flows on symplectic 44-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 C{C}^{\infty}-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.