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

The analytic minimal rank Sard Conjecture

A. Belotto da Silva Université Paris Cité and Sorbonne Université, UFR de Mathématiques, Institut de Mathématiques de Jussieu-Paris Rive Gauche, UMR7586, F-75013 Paris, France. Institut universitaire de France (IUF). andre.belotto@imj-prg.fr A. Parusiński Université Côte d’Azur, CNRS, Labo. J.-A. Dieudonné, UMR CNRS 7351, Parc Valrose, 06108 Nice Cedex 02, France adam.parusinski@unice.fr  and  L. Rifford Université Côte d’Azur, CNRS, Labo. J.-A. Dieudonné, UMR CNRS 7351, Parc Valrose, 06108 Nice Cedex 02, France ludovic.rifford@math.cnrs.fr
Abstract.

We obtain, under an additional assumption on the subanalytic abnormal distribution constructed in [4], a proof of the minimal rank Sard conjecture in the analytic category. It establishes that from a given point the set of points accessible through singular horizontal curves of minimal rank, which corresponds to the rank of the distribution, has Lebesgue measure zero. The minimal rank Sard Conjecture is equivalent to the Sard Conjecture for co-rank 11 distributions.

1. Introduction

The topic of this paper is the minimal rank Sard Conjecture in sub-Riemannian geometry. This is a follow-up of our previous work [4], where we provide a geometrical setting to study the Sard Conjecture in arbitrary dimensions. We rely on [4, Sections 1.1, 1.2 and 1.4] for a complete presentation of the Conjecture and its importance.

Let MM be a smooth connected manifold of dimension n3n\geq 3 equipped with a distribution Δ\Delta of rank m<nm<n which is bracket generating. Recall that the Sard Conjecture is only known in very few cases whenever dimM>3\dim M>3. In fact, all previous results focus on Carnot groups [1, 6, 12, 16, 17, 19, 13],. Whenever dim(M)=3\mbox{dim}(M)=3, much more is known by following a geometrical approach inspired by the foundation paper of Zelenko and Zhimtomirskii [23]. In fact, in a joint work with Figalli [2], we have proved the strong version of the Sard Conjecture in the analytic three dimensional case, improving a previous result of Belotto and Rifford [5]. The proof is based on a delicate study of the geometrical properties of the, so-called, characteristic foliation introduced in [23]. It makes use of methods from symplectic geometry, differential topology and singularity theory. Some of the singularity methods – such as resolution of singularities of metrics and foliations, and regularity of transition maps – are constrained to small dimensions.

The goal of this paper is to generalize the heart of the arguments of [2] to arbitrary dimensions, in order to address the minimal rank Sard Conjecture. This is the first step in our program to understand the Sard Conjecture in arbitrary dimensions, as explained in [4, Section 1.4]. To this end, we follow the framework introduced in [4], where we developed the properties of the natural generalization of characteristic foliations, which we called abnormal distribution 𝒦\vec{\mathcal{K}}, see [4, Theorem 1.1]. We then proceed to replace several of the delicate singularity arguments from [2] by subanalytic and symplectic arguments; we highlight the new notion of witness transverse sections which we expect to be of independent interest to foliation theory, see Theorem 3.2. This allows us to prove the minimal rank Sard Conjecture under an additional qualitative hypothesis on the abnormal distribution, which we call splitability, see Theorem 1.1 and Definition 1.3. The remaining difficulty is, non-surprisingly, related to singularity theory: one needs to exclude pathological asymptotic behaviors of a foliation of rank at least 22 over its singular set, see Section 2.2 for an example.

1.1. Main result

Let us briefly recall the main objects introduced in [4] – we equally rely on that after-mentioned paper for extended discussion on these objects.

Given an analytic totally nonholonomic distribution Δ\Delta on a real-analytic connected manifold MM, we consider the nonzero annihilator of Δ\Delta in the cotagent bundle TMT^{*}M:

(1.1) Δ:={𝔞=(x,p)TM|p0 and pv=0,vΔ(x)}.\Delta^{\perp}:=\Bigl{\{}{\mathfrak{a}}=(x,p)\in T^{*}M\,|\,p\neq 0\mbox{ and }p\cdot v=0,\,\forall v\in\Delta(x)\Bigr{\}}.

By [4, Theorem 1.1], there exists an open and dense subanalytic set 𝒮0Δ\mathcal{S}_{0}\subset\Delta^{\perp} called essential domain. Its complement, the set Σ=Δ𝒮0\Sigma=\Delta^{\perp}\setminus\mathcal{S}_{0}, is in fact a proper analytic subvariety of Δ\Delta^{\perp}. Moreover, there exists a subanalytic distribution 𝒦\vec{\mathcal{K}} over Δ\Delta^{\perp} which is, in fact, a subanalytic isotropic foliation over 𝒮0\mathcal{S}_{0}; we denote the restriction of 𝒦\vec{\mathcal{K}} to S0S_{0} by 𝒦|S0\vec{\mathcal{K}}_{|S_{0}}. We call 𝒦\vec{\mathcal{K}} the abnormal distribution because a horizontal path γ:[0,1]M\gamma:[0,1]\to M is singular if, and only if, it admits a lift ξ:[0,1]Δ\xi:[0,1]\to\Delta^{\perp} which is horizontal with respect to 𝒦\vec{\mathcal{K}}, see [4, Theorem 1.1(iii)].

In our main result, we will need to add an extra hypothesis on the foliation 𝒦|S0\vec{\mathcal{K}}_{|S_{0}}, which we call splittability. We postpone the precise definition to Section 1.3, and we anticipate in here that any line foliation is splittable. We can now state our main result:

Theorem 1.1 (Minimal rank Sard Conjecture for splittable foliatons).

Assume that both MM and Δ\Delta are real-analytic. If the foliation 𝒦|S0\vec{\mathcal{K}}_{|S_{0}} is splittable, then the minimal rank Sard conjecture holds true.

The proof of Theorem 1.1 will consist in showing by contradiction that if the set of minimal rank singular horizontal paths from a given point reaches a set of positive Lebesgue measure in MM, then we can, roughly speaking, lift all those horizontal paths into abnormal curves sitting in the leaves of the foliations given by 𝒦\vec{\mathcal{K}} on its essential domain and from here get a contradiction. This strategy requires to be able to select from a given set of positive measure contained in a transverse local section of the foliation 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}} a subset of positive measure whose all points belong to distinct leaves of 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}. A foliation subject to such a selection result will be called splittable; this is the heuristic behind the definition of splittability given in Section 1.3.

Thanks to [4, Theorem 1.1(iv)], moreover, we know that the rank of the abnormal distribution restricted to its essential domain 𝒦|S0\vec{\mathcal{K}}_{|S_{0}} is less than or equal to m2m-2. Hence, the Minimal rank Sard conjecture holds true whenever Δ\Delta has rank 3\leq 3. Furthermore, the equivalence of the minimal rank Sard Conjecture with the Sard Conjecture in the case of corank-11 distributions yields the following immediate corollary:

Corollary 1.2.

Assume that both MM and Δ\Delta are analytic. If Δ\Delta has codimension one (m=n1m=n-1) and the distribution 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}} is splittable, then the Sard conjecture holds.

In particular, the Sard Conjecture holds true when n=4n=4 and m=3m=3. This four dimensional result remains true in the CC^{\infty} category, as we show in our follow-up work [3], where we extend a few results (as much as we could) from this work to the smooth category.

Finally, our approach to the proof of Theorem 1.1 requires to lift the set of singular horizontal curves in MM to a subset of Δ\Delta^{\perp} of positive transverse volume with respect to 𝒦\vec{\mathcal{K}}. As a consequence, we cannot prove the Sard conjecture for distribution of corank strictly greater than one yet. For an extended discussion, see [4, Section 1.4 and 1.5].

1.2. Witness transverse sections

The proof of Theorem 1.1 follows from a combination of the description of abnormal lifts given in [4, Theorem 1.1]; differential geometry methods; and a new result on the existence of special transverse sections for singular analytic foliations, which we call witness transverse sections, see Theorem 3.2.

In fact, roughly speaking, we show that if \mathcal{F} is a singular analytic foliation of generic corank rr in a real-analytic manifold NN equipped with a smooth Riemannian metric gg, then we can construct locally, for every point xx in the singular set Σ\Sigma of \mathcal{F}, a special subanalytic set XVΣX\subset V\setminus\Sigma where VV is an open neighborhood of xx, called witness transverse section. This section has the property that its slices Xc:=Xh1(c)X^{c}:=X\cap h^{-1}(c) (c>0c>0) with respect to some nonnegative analytic function hh (verifying ΣV={h=0}\Sigma\cap V=\{h=0\}) have dimension r\leq r with rr-dimensional volume uniformly bounded (w.r.t cc) and such that any point of VV can be connected to XcX^{c} through a horizontal curve (w.r.t. \mathcal{F}) of length no greater than \ell (w.r.t. gg). We refer to Section 3 for further detail.

1.3. Splittable foliation

Let NN be a real-analytic manifold of dimension n2n\geq 2 equipped with a smooth Riemannian metric hh (not necessary assumed to be complete) and \mathcal{F} a (regular) analytic foliation of constant rank d[1,n1]d\in[1,n-1]. Given >0\ell>0, we say that two points xx and yNy\in N are (,)(\mathcal{F},\ell)-related if there exists a smooth path φ:[0,1]N\varphi:[0,1]\rightarrow N with length [0,]\in[0,\ell] with respect to hh which is horizontal with respect to \mathcal{F} and joins xx to yy. Note that the (Δ,)(\Delta,\ell)-relation is not an equivalence relation, since it is not transitive. Moreover, given a point x¯N\bar{x}\in N, we call local transverse section at x¯\bar{x} any set SNS\subset N containing x¯\bar{x} which is a smooth submanifold diffeomorphic to an open disc of dimension ndn-d and transverse to the leaves of \mathcal{F}.

Definition 1.3 (Splittable foliation).

We say that the foliation \mathcal{F} is splittable in (N,h)(N,h) if for every x¯N\bar{x}\in N, every local transverse section SS at x¯\bar{x} and every >0\ell>0, the following property is satisfied:

For every Lebesgue measurable set ESE\subset S with nd(E)>0\mathcal{L}^{n-d}(E)>0, there is a Lebesgue measurable set FEF\subset E such that nd(F)>0\mathcal{L}^{n-d}(F)>0 and for all distinct points x,yFx,\,y\in F, xx and yy are not (,)(\mathcal{F},\ell)-related.

We provide in Section 2 a sufficient conditions for a foliation to be splittable. Indeed, we introduce the notion of foliation having locally horizontal balls with finite volume (with respect to the metric hh in NN), see Definition 2.1, and we prove that this property implies the splittability, see Proposition 2.3. As a consequence, we infer that every line foliation is splittable, as well as every foliation whose leaves have Ricci curvatures uniformly bounded from below. In particular, all regular foliations in a compact manifold are splittable. An example of non-splittable analytic foliation in a non-compact manifold equipped with a smooth metric is presented in Section 2.2; we do not know if such examples do exist with an analytic metric.

1.4. Paper structure

The paper is organized as follows: in Section 2, we discuss the notion of spplitable foliations. In section 3 we construct Witness transverse sections for analytic foliations, see Theorem 3.2. Finally, the proof of Theorem 1.1 is given in Section 4.

Acknowledgment: The first author is supported by the project “Plan d’investissements France 2030”, IDEX UP ANR-18-IDEX-0001, and partially supported by the Agence Nationale de la Recherche (ANR), project ANR-22-CE40-0014.

2. Splittable foliations

2.1. Sufficient conditions for splittability

The notion of splittable foliation and of (,)(\mathcal{F},\ell)-related points have been provided in the Introduction, see Definition 1.3, and we follow its notation. Let us start our discussion by providing a more structured way to describe two (,)(\mathcal{F},\ell)-related points. In fact, let us consider horizontal balls with respect to \mathcal{F} and hh. Given xNx\in N, we denote by x\mathcal{L}_{x} the leaf of \mathcal{F} through xx in NN. Then, for every >0\ell>0, we call horizontal ball with respect to \mathcal{F} and hh the subset of x\mathcal{L}_{x} given by

x:={yLx|φ:[0,1]x abs. cont. s.t. φ(0)=x,φ(1)=y,lengthh(φ)}.\mathcal{L}_{x}^{\ell}:=\left\{y\in L_{x}\,|\,\exists\,\varphi:[0,1]\to\mathcal{L}_{x}\mbox{ abs. cont. s.t. }\varphi(0)=x,\,\varphi(1)=y,\,\mbox{length}^{h}(\varphi)\leq\ell\right\}.

We check easily that two points x,yNx,y\in N are (,)(\mathcal{F},\ell)-related if and only if yxy\in\mathcal{L}_{x}^{\ell} (or xyx\in\mathcal{L}_{y}^{\ell}). Let us now introduce the following definition where volh,(A)\mbox{vol}^{h,\mathcal{F}}(A) stands for the volume of a Borel set AA contained in a leaf of \mathcal{F} with respect to the Riemannian metric induced by hh on that leaf:

Definition 2.1.

We say that the foliation \mathcal{F} has locally horizontal balls with finite volume (w.r.t. hh) if for every xNx\in N and every >0\ell>0, there are V>0V>0 and a neighborhood 𝒰\mathcal{U} of xx such that volh,(y)V\mbox{\em vol}^{h,\mathcal{F}}(\mathcal{L}_{y}^{\ell})\leq V for all y𝒰y\in\mathcal{U}.

The first example of foliations having locally horizontal balls with finite volume is given by foliations associated with complete Riemannian metrics. As a matter of fact, if hh is complete, then by the Hopf-Rinow Theorem, all balls y\mathcal{L}_{y}^{\ell} with yy close to xx are contained in the ball centered at xx with radius +1\ell+1 which happens to be compact, so all of those horizontal balls are compact sets with a volume which is finite and depends continuously upon yy. Another example is given by foliations whose curvature satisfy a lower bound:

Proposition 2.2.

If \mathcal{F} has rank 11 then it has locally horizontal balls with finite volume, indeed we have for any xNx\in N and >0\ell>0, volh,(x)2\mbox{\em vol}^{h,\mathcal{F}}(\mathcal{L}_{x}^{\ell})\leq 2\ell. Moreover, if \mathcal{F} has rank 2\geq 2 and the Ricci curvature (w.r.t. hh) of all its leaves is uniformly bounded from below, then it has locally horizontal balls with finite volume (w.r.t. hh).

The proof of this result is left to the reader. We draw their attention to the fact that the comparison theorem required for the proof (of the second part) remains true for a non-complete metric (see e.g. [8, §4]). We end this section with the result that justifies the introduction of Definition 2.1 and provide many examples of splittable foliations.

Proposition 2.3.

If \mathcal{F} has locally horizontal balls with finite volume (w.r.t. hh), then it is splittable in (N,h).

Proof.

Let x¯N\bar{x}\in N and 0>0\ell_{0}>0 be fixed. Let >0\ell>\ell_{0} and denote by x<x\mathcal{L}_{x}^{<\ell}\subset\mathcal{L}_{x}^{\ell} the union of all x\mathcal{L}_{x}^{\ell^{\prime}} with <\ell^{\prime}<\ell. Let V>0V>0 be such that volh,(x4)V\mbox{vol}^{h,\mathcal{F}}(\mathcal{L}_{x}^{4\ell})\leq V for all xx in an open neighborhood UU of x¯\bar{x}. By considering a foliation chart (see [4, Section 3]) and shrinking UU if necessary, there exists a diffeomorphism Φ:WU\Phi:W\to U such that W=(1,1)nnW=(-1,1)^{n}\subset\mathbb{R}^{n}, Φ(0)=x¯\Phi(0)=\bar{x}, the pull-back foliation is given by (xnd+1==xn=cte)(x_{n-d+1}=\ldots=x_{n}=cte) (where dd is the rank of \mathcal{F}) and the following properties are verified:

  • there exists a smooth transverse section 𝒟\mathcal{D} diffeomorphic to a disc of dimension ndn-d;

  • there exists ϵ>0\epsilon>0 such that, for every point x𝒟x\in\mathcal{D}, the connected component of xU\mathcal{L}_{x}\cap U containing xx, which we denote by x,U\mathcal{L}_{x,U}, is such that volh,(x,U)>ϵ\mbox{vol}^{h,\mathcal{F}}(\mathcal{L}_{x,U})>\epsilon and diam(x,U)<\mbox{diam}(\mathcal{L}_{x,U})<\ell (where diam stands for the diameter w.r.t. hh).

Let KK be a natural number greater than V/ϵV/\epsilon. By construction, if x,y𝒟x,y\in\mathcal{D} are (,)(\mathcal{F},\ell)-related then we have x,U,y,Ux4\mathcal{L}_{x,U},\mathcal{L}_{y,U}\subset\mathcal{L}_{x}^{4\ell}. Thus, since volh,(x,U)>ϵ\mbox{vol}^{h,\mathcal{F}}(\mathcal{L}_{x,U})>\epsilon for every x𝒟x\in\mathcal{D} and volh,(x4)V\mbox{vol}^{h,\mathcal{F}}(\mathcal{L}_{x}^{4\ell})\leq V for all xUx\in U, we infer that for every x𝒟x\in\mathcal{D}, there are at most KK points in 𝒟\mathcal{D} which are (,<)(\mathcal{F},<\ell)-related to xx (that is, they are (,)(\mathcal{F},\ell^{\prime})-related to xx for some <\ell^{\prime}<\ell). Denote by {x1,,xkx}\{x_{1},\ldots,x_{k_{x}}\} these points, where kxKk_{x}\leq K depends on x𝒟x\in\mathcal{D}. Let E𝒟E\subset\mathcal{D} be a measurable set such that nd(E)>0\mathcal{L}^{n-d}(E)>0. Let kk be the maximum value of kxk_{x} for every xEx\in E which is a density point of EE. Fix a density point xEx\in E such that kx=kk_{x}=k and consider the set {x1,,xk}\{x_{1},\ldots,x_{k}\} of (,<)(\mathcal{F},<\ell)-related points to xx in 𝒟\mathcal{D}. Denote by φi:[0,1]x\varphi_{i}:[0,1]\to\mathcal{L}_{x}, for i=1,,ki=1,\ldots,k, the absolutely continuous curves of length <<\ell between xx and xix_{i} respectively. Since \mathcal{F} is everywhere regular and φi\varphi_{i} has compact domain, we conclude from the foliation charts that there exists a transverse section 𝒟x𝒟\mathcal{D}_{x}\subset\mathcal{D} containing xx and diffeomorphic to a disc of dimension ndn-d, such that: for every y𝒟xy\in\mathcal{D}_{x} the curves φi\varphi_{i} can be diffeomorphically deformed into an absolutely continuous curve φi~:[0,1]y\widetilde{\varphi_{i}}:[0,1]\to\mathcal{L}_{y} starting from yy and finishing at a point yi𝒟y_{i}\in\mathcal{D} with length <<\ell, for every i=1,,ki=1,\ldots,k. Now, since all the points {x,x1,,xk}\{x,x_{1},\ldots,x_{k}\} are distinct, apart from shrinking 𝒟x\mathcal{D}_{x}, we may suppose that for every y𝒟xy\in\mathcal{D}_{x}, all other points {y1,,yk}\{y_{1},\ldots,y_{k}\} do not belong to 𝒟x\mathcal{D}_{x}. We now consider F=E𝒟xF=E\cap\mathcal{D}_{x}. First, note that nd(F)>0\mathcal{L}^{n-d}(F)>0 since xx is a density point. Moreover, for yFy\in F, we know that kykk_{y}\geq k since y𝒟xy\in\mathcal{D}_{x}, and that kykk_{y}\leq k since yEy\in E. We conclude easily that every two points of FF are not (,0)(\mathcal{F},\ell_{0})-related, finishing the proof. ∎

As a consequence of Propositions 2.2 and 2.3, we get the following result:

Proposition 2.4.

Every foliation of rank 11 is splittable.

We provide in Section 2.2 an example of analytic foliation {\mathcal{F}} contained in an analytic manifold with boundary MM, endowed with a (non-complete) CC^{\infty} metric gg, which is non-splittable. This example illustrates the kind of qualitative behavior that we must exclude when studying the minimal rank Sard Conjecture. Nevertheless, note that the example is constructed on an abstract manifold. We do not know the answer to the following question:

Open question. Is there an integrable family of analytic 11-forms Ω=(ω1,,ωt)\Omega=(\omega_{1},\ldots,\omega_{t}) defined over an open set UnU\subset\mathbb{R}^{n} whose associated analytic foliation {\mathcal{F}} defined in UΣU\setminus\Sigma, where Σ\Sigma is the singular set of Ω\Omega, is non-splittable in (U,g0)(U,g^{0}) where g0g^{0} is the Euclidean metric?

If the answer to the above question is negative, then the hypothesis of Theorem 1.1 would always be satisfied provided that MM and Δ\Delta are analytic.

2.2. Example of a non-splittable foliation

We modify a construction of Hirsch [11] in order to define a foliation which is non-splittable in a (non-compact) manifold with border MM. As a matter of fact, Hirsch foliations are two-dimensional analytic foliations which satisfy the topological properties of a non-splittable foliation, but they lack the metric properties. In order to obtain the metric properties, we modify the original construction, and we make use of CC^{\infty}-partitions of the unit to yield a CC^{\infty}-metric.

We start by defining the building-blocks. Consider the double cover immersion f:S1S1f:S^{1}\to S^{1} given by f(t)=2tf(t)=2t, and choose an analytic embedding ι\iota of the solid torus S1×D2S^{1}\times D^{2} onto its interior so that πι=fπ\pi\circ\iota=f\circ\pi, where π:S1×D2S1\pi:S^{1}\times D^{2}\to S^{1} is the projection. Let V=S1×D2Int(ι(S1×D2))V=S^{1}\times D^{2}\setminus\mbox{Int}(\iota(S^{1}\times D^{2})). Then the boundary of VV is two copies of S1×S1S^{1}\times S^{1}, which we denote by VV^{-} and V+V^{+} where ι(V)=V+\iota(V^{-})=V^{+}. Denote by 𝒢\mathcal{G} foliation over VV induced by the the fibration π\pi. Note that the leaves of this foliation are topological pants, whose intersection with VV^{-} is a S1S^{1}, and whose intersection with V+V^{+} is the disjoint union of two S1S^{1}, cf. figure 1(a).

Refer to caption
(a) The building block VV.
Refer to caption
(b) A leaf of \mathcal{F}.
Figure 1. Geometrical illustration of \mathcal{F}.

Now, we consider a countable family of building-blocks (Vn,Vn,Vn+,ιn,𝒢n,gn)(V_{n},V_{n}^{-},V_{n}^{+},\iota_{n},\mathcal{G}_{n},g_{n}), where gng_{n} are analytic metrics over VnV_{n} satisfying the following property: given two points xx and yy in a leaf LL of 𝒢n\mathcal{G}_{n}, the distance of xx and yy in LL is bounded by 4n4^{-n}. We denote by MM the manifold with boundary given by the union of all VnV_{n}, by identifying VnV_{n}^{-} with Vn+1+V_{n+1}^{+} via ι\iota, that is, we take the identification xVnx\in V_{n}^{-} equivalent to ι(x)Vn+1+\iota(x)\in V_{n+1}^{+}. This yields an analytic manifold with border, where the border is a torus M0=V0+=S1×S1M_{0}=V_{0}^{+}=S^{1}\times S^{1}. This construction induces, furthermore, an analytic foliation \mathcal{F} over the manifold with border MM which locally agrees with 𝒢n\mathcal{G}_{n} over each VnV_{n}, because πf=ιπ\pi\circ f=\iota\circ\pi, cf. figure 1(b). Furthermore, we can define a globally defined CC^{\infty} metric gg over MM by patching the metrics gng_{n} via partition of the unit. We may chose such a partition so that gg satisfies the following property: given two points xx and yy in a leaf LL of 𝒢n\mathcal{G}_{n}, the distance of xx and yy in LL is bounded by 2n2^{-n}.

We claim that \mathcal{F} is a non-splittable foliation. Indeed, consider a transverse section Σ=S1M0=S1×S1\Sigma=S^{1}\subset M_{0}=S^{1}\times S^{1} and let us identify Σ\Sigma with the interval [0,1][0,1]. Given a point xΣx\in\Sigma, denote by LxL_{x} the leaf passing by xx. First, consider the foliation 𝒢0\mathcal{G}_{0}, and note that, since f(x)=f(x+1/2)f(x)=f(x+1/2) and 𝒢0\mathcal{G}_{0} is a foliation by pants, x+1/2x+1/2 also belongs to the leaf LL, cf. figure 1(b). Since this argument can be iterated over any 𝒢n\mathcal{G}_{n}, we get that all points x+m/2nx+m/2^{n} with m,nm,n\in\mathbb{N} belong to LxL_{x}. Moreover, the distance on LxL_{x} between xx and x+m/2nx+m/2^{n} is bounded by:

2k=0n12k<42\cdot\sum_{k=0}^{n}\frac{1}{2^{k}}<4

since there exists a path between xx and x+m/2nx+m/2^{n}, contained in the leaf LxL_{x}, and which is contained in the union of VkV_{k} with k<nk<n, crossing each of these components at most twice. In conclusion, for every xΣx\in\Sigma, the intersection of LxL_{x}, the leaf passing through xx, with Σ\Sigma is a countable and dense set of points invariant by a countable subgroup of rotations, which are pairwise (,4)(\mathcal{F},4)-related. We infer that \mathcal{F} is a not splittable in (M,g)(M,g) because there is no measurable set EΣE\subset\Sigma with positive Lebesgue measure whose intersection with each LxL_{x} (with xΣx\in\Sigma) contains only one point.

3. Witness transverse sections

We follow the notation and we use results given in [4, Section 3.3 and 3.4]. The main goal of this section is to show the following result:

Theorem 3.1.

Let NN be a real-analytic manifold of dimension d1d\geq 1 equipped with a complete smooth Riemannian metric gg, Ω\Omega be a family (or, more generally, a sheaf) of analytic 11-forms which is integrable of generic corank rr, with singular set Σ\Sigma. Denote by 𝒦\vec{\mathcal{K}} the distribution associated to Ω\Omega and by {\mathcal{F}} the foliation on NΣN\setminus\Sigma associated to Ω\Omega. Let xNx\in N and >0\ell>0 be fixed. Then, there exist a relatively compact open neighborhood VV of xx in NN, a real-analytic function h:V[0,)h:V\to[0,\infty), a subanalytic set XVΣX\subset V\setminus\Sigma and C,ϵ>0C,\epsilon>0 such that the following properties are satisfied:

  • (i)

    The set h1(0)h^{-1}(0) is equal to ΣV\Sigma\cap V.

  • (ii)

    (uniform volume bound) dimXr+1\dim X\leq r+1 and for every 0<c<ϵ0<c<\epsilon the subanalytic set Xc:=Xh1(c)X^{c}:=X\cap h^{-1}(c) satisfies dimXcr\dim X^{c}\leq r and its rr-dimensional volume with respect to gg is bounded by CC.

  • (iii)

    (uniform intrinsic distance bound) For every 0<c<ϵ0<c<\epsilon and for every 𝔞h1(c)VΣ{\mathfrak{a}}\in h^{-1}(c)\subset V\setminus\Sigma, there is a smooth curve α:[0,1]VΣ\alpha:[0,1]\rightarrow V\setminus\Sigma contained in 𝔞h1(c)\mathcal{L}_{{\mathfrak{a}}}\cap h^{-1}(c), where 𝔞\mathcal{L}_{{\mathfrak{a}}} is the leaf of {\mathcal{F}} containing 𝔞{\mathfrak{a}}, such that

    α(0)=𝔞,α(1)Xc,andlengthg(α).\alpha(0)={\mathfrak{a}},\quad\alpha(1)\in X^{c},\quad\mbox{and}\quad\mbox{\em length}^{{g}}(\alpha)\leq\ell.
  • (iv)

    (generic tranversality) There is a decomposition X=YZX=Y\sqcup Z as the disjoint union of two subanalytic sets Y,ZY,Z such that: dimZr\dim Z\leq r, Y=iIYiY=\bigsqcup_{i\in I}Y_{i} is a finite disjoint union of smooth subanalytic sets YiY_{i} of dimension r+1r+1, and, moreover, for every 0<c<ϵ0<c<\epsilon, Zc:=Zh1(c)Z^{c}:=Z\cap h^{-1}(c) is of dimension <r<r, Yic:=Yih1(c)Y_{i}^{c}:=Y_{i}\cap h^{-1}(c) is smooth of dimension rr such that

    Yc=Yc¯YcZc and T𝔞N=T𝔞Yic+𝒦(𝔞)𝔞Yic,iI.\partial Y^{c}=\overline{Y^{c}}\setminus Y^{c}\subset Z^{c}\quad\text{ and }\quad T_{{\mathfrak{a}}}N=T_{{\mathfrak{a}}}Y_{i}^{c}+\vec{\mathcal{K}}({\mathfrak{a}})\qquad\forall{\mathfrak{a}}\in Y_{i}^{c},\,\forall i\in I.

First we show the following general theorem on the existence of a transverse section, that is of independent interest for foliation theory. Then Theorem 3.1 will be a corollary of this result.

Theorem 3.2.

Let NN be a real-analytic manifold of dimension d1d\geq 1 equipped with a complete smooth Riemannian metric gg, Ω\Omega be a family (or, more generally, a sheaf) of analytic 11-forms which is integrable of generic corank rr, with singular set Σ\Sigma. Denote by 𝒦\vec{\mathcal{K}} the distribution associated to Ω\Omega, and by {\mathcal{F}} the foliation on NΣN\setminus\Sigma associated to Ω\Omega. Then, for every xNx\in N there exist a relatively compact open subanalytic neighborhood VV of xx in NN, a subanalytic set XVΣX\subset V\setminus\Sigma, called witness transverse section, such that the following properties are satisfied:

  • (i)

    For every zVΣz\in V\setminus\Sigma there is a smooth curve α:[0,1]VΣ\alpha:[0,1]\rightarrow V\setminus\Sigma contained in a leaf of {\mathcal{F}} such that

    α(0)X,α(1)=zandlengthg(α)Cddiamg(V),\alpha(0)\in X,\quad\alpha(1)=z\quad\mbox{and}\quad\mbox{\em length}^{g}(\alpha)\leq C_{d}\,\mbox{\em diam}^{g}(V),

    where CdC_{d} is a constant depending only on dd.

  • (ii)

    XX is the disjoint union of finitely many locally closed smooth subanalytic sets X=iXiX=\bigcup_{i}X_{i} of dimension at most rr such that for every 𝔞Xi{\mathfrak{a}}\in X_{i} we have 𝒦(𝔞)T𝔞Xi={0}\vec{\mathcal{K}}({\mathfrak{a}})\cap T_{{\mathfrak{a}}}X_{i}=\{0\}. In particular, if YY is the union of XiX_{i} of maximal dimension rr and ZZ the union of those of dimension <r<r then X=YZX=Y\sqcup Z and YY is transverse to the leaves of {\mathcal{F}}.

We may assume, without loss of generality, that the metric gg is real-analytic (or even Euclidean with respect to a fixed local coordinate system). Indeed, it is enough to show the statement of Theorem 3.2 locally at xx, and any CC^{\infty} metric gg is locally bi-Lipschitz equivalent to the Euclidean metric and, moreover, the bi-Lipschitz constant may be taken arbitrarily close to 11.

Given a small open neighborhood VV of xNx\in N there is a finite family of analytic functions on VV, 𝒢={gi}\mathcal{G}=\{g_{i}\}, such that:

  1. (i)

    each gig_{i} is Lipschitz with constant 22 (with respect to the geodesic distance dgd^{g}),

  2. (ii)

    for every xVx\in V, for every smooth submanifold MVM\subset V and every vector vTNv\in T_{N} there is an index ii such that

    1. (1)

      |(gi|M)(x)|1/2|\nabla(g_{i|M})(x)|\geq 1/2,

    2. (2)

      (gi|M)(x),v0\langle\nabla(g_{i|M})(x),v\rangle\geq 0.

Indeed, first note that it suffices to show that existence of the family {gi}\{g_{i}\} satisfying (i)(i) and (1)(1) of (ii)(ii). The conditions (2)(2) of (ii)(ii) can be obtained by adding to the family {gi}\{g_{i}\} all the opposite functions {gi}\{-g_{i}\}. To get (i)(i) and (1)(1) of (ii)(ii), by the preceding remark, it suffices to consider only the case when N=nN=\mathbb{R}^{n} and gg is the Euclidean metric. In this case, let us identify 𝕊n1\mathbb{S}^{n-1} with the vectors vnv\in\mathbb{R}^{n} of norm one. For each fixed w𝕊n1w\in\mathbb{S}^{n-1} there exists an open neighborhood UwU_{w} of ww in 𝕊n1\mathbb{S}^{n-1} such that, for all vUwv\in U_{w}, w,v>1/2\langle w,v\rangle>1/2. By the compactness of 𝕊n1\mathbb{S}^{n-1}, consider a finite family wiw_{i} such that UwiU_{w_{i}} covers 𝕊n1\mathbb{S}^{n-1}, and take as 𝒢\mathcal{G} the family of functions gi(x)=w,xg_{i}(x)=\langle w,x\rangle. Since |gi(x)|=|wi|=1|\nabla g_{i}(x)|=|w_{i}|=1, the family clearly satisfies (i)(i). Next, fixed a submanifold MM and a vector vTxMv\in T_{x}M of norm 11, there exists i0i_{0} such that |(gi0|M)(x)||gi0(x),v|=|wi0,v|1/2|\nabla(g_{i_{0}|M})(x)|\geq|\langle\nabla g_{i_{0}}(x),v\rangle|=|\langle w_{i_{0}},v\rangle|\geq 1/2 by construction, proving (ii)(ii).

Remark 3.3.

There is a family 𝒢={gi}\mathcal{G}=\{g_{i}\} of functions defined on the entire NN that satisfies the above properties (i)(i) and (ii)(ii) at every point xNx\in N. Indeed, one may show it first in the class of C1C^{1} functions and then approximate them by real analytic ones in Whitney C1C^{1}-topology, see e.g. [9]. Therefore, in this case, a stronger version of Theorem 3.2 holds, where we may take V=NV=N and replace diamg(V)\mbox{\em diam}^{g}(V) by an arbitrary constant D>0D>0. We do not need this stronger result in this paper.

We now consider an auxiliary locally closed nonsingular subanalytic subset WW of NΣN\setminus\Sigma. Note that we do not fix its dimension; in fact, we will make inductive arguments based on its dimension. Later on, in the proof of Theorem 3.2, WW will be taken as a connected component of VΣV\setminus\Sigma. Following the notion introduced in [4, Section 3] we say that is 𝒦\vec{\mathcal{K}} is regular on W{W} if the restriction of 𝒦\vec{\mathcal{K}} to W{W} is a regular analytic distribution and 𝒦\vec{\mathcal{K}} has constant rank along W{W}. We denote by 𝒦W\vec{\mathcal{K}}_{W} this restriction and by rWr_{W} its corank. By [4, Remark 3.11], rWrr_{W}\leq r, 𝒦W\vec{\mathcal{K}}_{W} is integrable and induces a foliation that we denote by W{\mathcal{F}}_{W}. We start by studying what happens when 𝒦\vec{\mathcal{K}} is regular on WW:

Lemma 3.4.

Following the above notation, suppose that 𝒦\vec{\mathcal{K}} is regular on WW and rW<dimWr_{W}<\dim{W}. Then there exists a subanalytic (as a subset of NN) subset YWWY_{W}\subset{W} of dimension <dimW<\dim{W} such that for every zWz\in{W} there is a smooth curve α:[0,1]W\alpha:[0,1]\rightarrow{W}, contained entirely in a leaf of W{\mathcal{F}}_{W} such that

α(0)YW,α(1)=zandlengthg(α)4distg(z,α(0)).\alpha(0)\in Y_{W},\quad\alpha(1)=z\quad\mbox{and}\quad\mbox{\em length}^{g}(\alpha)\leq 4\,\mbox{\em dist}^{g}(z,\alpha(0)).
Proof.

We work locally in a relatively compact neighborhood VV of xW¯x\in\overline{W}. Let ff be a C2C^{2} subanalytic function such that f1(0)=(W¯W)(V¯WV)f^{-1}(0)=(\overline{W}\setminus{W})\cup(\overline{V}\cap{W}\setminus V). Such a function, even a function of class CpC^{p} for any fixed finite pp, always exists. It follows from a more general result valid in any o-minimal structure, see Theorem C11 of [22]. The subanalytic case, that we use here, was proven first by Bierstone, Milman and Pawłucki (unpublished). By replacing ff by f2f^{2} we may suppose f0f\geq 0. Fix a family 𝒢={gi}\mathcal{G}=\{g_{i}\} of analytic functions as above and define

Yi:=BdW({zW;|(gi|W)(z)|=1/2}) BdW({zW;|(f|W)(x),(gi|W)(x)=0}),Y_{i}:=\mbox{Bd}_{W}\left(\left\{z\in{W}\,;\,|\nabla(g_{i|{\mathcal{F}}_{W}})(z)|=1/2\right\}\right)\\ \cup\mbox{ Bd}_{W}\left(\left\{z\in{W}\,;\,|\langle\nabla(f_{|{\mathcal{F}}_{W}})(x),\nabla(g_{i|{\mathcal{F}}_{W}})(x)\rangle=0\right\}\right),

where by BdW\mbox{Bd}_{W} we mean the topological boundary in W{W}. Here by (f|W)(z)\nabla(f_{|{\mathcal{F}}_{W}})(z) we mean the gradient of the function: ff restricted to the leaf of W{\mathcal{F}}_{W} through zz. These leaves are of dimension 1\geq 1 by the assumption rW<dimWr_{W}<\dim{W}.

The sets YiY_{i} are subanalytic (the Riemannian metric gg is assumed real-analytic) and of dimension <dimW<\dim{W}. Then we take as YWY_{W} the union of all YiY_{i}.

Let z(WYW)Vz\in({W}\setminus Y_{W})\cap V be fixed. By the above property (ii), there is ii such that

|(gi|F)(z)|1/2and(f|F)(z),(gi|F)(z)0,|\nabla(g_{i|F})(z)|\geq 1/2\quad\mbox{and}\quad\langle\nabla(f_{|F})(z),\nabla(g_{i|F})(z)\rangle\geq 0,

where FF is the leaf of W{\mathcal{F}}_{W} containing zz. Let β:[0,t0)WYW\beta:[0,t_{0})\to{W}\setminus Y_{W} be the maximal integral curve of (gi|F)\nabla(g_{i|F}) with β(0)=z\beta(0)=z (the curve α\alpha from the Lemma will be later taken as a reparametrization of β\beta). It is of finite length. Indeed, for any t1[0,t0)t_{1}\in[0,t_{0}), we have (note that by construction of YWY_{W}, |(gi|F)(β(t)|1/2|\nabla(g_{i|F})(\beta(t)|\geq 1/2 for all t[0,t0)t\in[0,t_{0}))

(3.1) gi(β(t1))gi(z)=0t1|(gi|F)(β(t)|2dt120t1|(gi|F)(β(t)|dt=12lengthg(β([0,t1]).g_{i}(\beta(t_{1}))-g_{i}(z)=\int_{0}^{t_{1}}\left|\nabla(g_{i|F})(\beta(t)\right|^{2}\,dt\\ \geq\frac{1}{2}\int_{0}^{t_{1}}\left|\nabla(g_{i|F})(\beta(t)\right|\,dt=\frac{1}{2}\,\mbox{length}^{g}(\beta([0,t_{1}]).

Therefore, since gi|Vg_{i|V} is bounded, t0t_{0} is finite and limtt0β(t)\lim_{t\to t_{0}}\beta(t) exists. We denote it by β(t0)\beta(t_{0}). Because f(β(t))f(\beta(t)) is not decreasing it is not possible that β(t0)(W¯W)(V¯WV)\beta(t_{0})\in(\overline{W}\setminus{W})\cup(\overline{V}\cap{W}\setminus V) and therefore β(t0)YW\beta(t_{0})\in Y_{W}. Moreover, since gig_{i} is 22-Lipschitz, (3.1) yields

lengthg(β)2(gi(β(t0))gi(z))4dg(β(t0),z).\mbox{length}^{g}(\beta)\leq 2\left(g_{i}(\beta(t_{0}))-g_{i}(z)\right)\leq 4\,d^{g}(\beta(t_{0}),z).

The curve β\beta is analytic except, maybe, at t0t_{0}. In this case we reparametrize it to obtain a smooth curve α\alpha satisfying the statement. ∎

Remark 3.5.

Lemma 3.4 implies that every leaf of {\mathcal{F}} intersecting W{W} intersects YWY_{W}. A similar result was shown in the definable set-up in [21] under an additional assumption that the leaves of {\mathcal{F}} are Rolle, see [21, Proposition 2.2]. This extra assumption implies that the leaves are locally closed that is not the case in general.

More generally, recall that we say that a stratification 𝒮=(Sα)\mathcal{S}=(S_{\alpha}) of W{W} is compatible with the distribution 𝒦\vec{\mathcal{K}} if for every stratum SS of 𝒮\mathcal{S}, 𝒦\vec{\mathcal{K}} is regular on SS. In this case, for every stratum SS, denote the restriction of 𝒦\vec{\mathcal{K}} to SS by 𝒦S\vec{\mathcal{K}}_{S}, and its corank, which is constant on SS, by rSr_{S}.

Proposition 3.6.

Let W{W} be a locally closed relatively compact nonsingular subanalytic subset of NΣN\setminus\Sigma. There exists a subanalytic stratification 𝒮\mathcal{S} of W{W}, compatible with 𝒦\vec{\mathcal{K}}, satisfying the following property: let X0X_{0} denote the union of all strata SS^{\prime} of 𝒮\mathcal{S} for which dimS=rS\dim S^{\prime}=r_{S^{\prime}} (i.e. the leaves of S{\mathcal{F}}_{S^{\prime}} are points), then for every zWz\in{W} there is a smooth curve α:[0,1]WΣ\alpha:[0,1]\rightarrow{W}\setminus\Sigma, contained in the leaf of \mathcal{F} through zz such that

α(0)X0,α(1)=zandlengthg(α)Cddiamg(W).\alpha(0)\in X_{0},\quad\alpha(1)=z\quad\mbox{and}\quad\mbox{\em length}^{g}(\alpha)\leq C_{d}\,\,\mbox{\em diam}^{g}({W}).

Note that a refinement of a stratification satisfying the conclusion of the above proposition also satisfies this conclusion. Therefore, if we want this stratification to satisfy additional properties, to be Whitney for instance, we replace it by its refinement.

Proof.

Induction on dimW\dim W. Let 𝒮\mathcal{S} be a subanalytic stratification of WW compatible with 𝒦\vec{\mathcal{K}}. The cases of dimW=0\dim{W}=0 or rS=dimSr_{S}=\dim S for all strata SS are obvious. Therefore we may assume that there is a stratum SS such that rS<dimSr_{S}<\dim S and then we apply Lemma 3.4 to SS. Let YSY_{S} be the set given by this lemma, which has dimension smaller than dimW\dim W. We stratify YSY_{S} (it is not necessarily nonsingular) and apply the inductive assumption to every stratum. We repeat this procedure for all strata SS such that rS<dimSr_{S}<\dim S. The obtained stratification satisfies the statement for SS.

Indeed, let us prove the last property, namely the bound on the length of α\alpha. Let zSz\in S. By Lemma 3.4 we may connect zz and a point of yYSy\in Y_{S} by an arc in a leaf of {\mathcal{F}} of length 4diamg(S)\leq 4\mbox{diam}^{g}(S). The point yy belongs to a stratum of smaller dimension and we may use to it the inductive assumption. So finally we may connect zz to a point of X0X_{0} by an arc of 4dimS diamg(S)\leq 4^{\dim S}\mbox{ diam}^{g}(S). Since every leaf of \mathcal{F} is smooth, and this arc has at most dd non-smooth points, we can reparameterize it by an everywhere smooth arc without increasing its length. It shows that we may choose Cd=4dC_{d}=4^{d}. ∎

3.0.1. Proof of Theorem 3.2.

Let VV be a subanalytic open relatively compact connected subset of NN. Let X0V¯ΣX_{0}\subset\overline{V}\setminus\Sigma be the set given by Proposition 3.6 for W=VΣ{W}=V\setminus\Sigma. Then, X=X0X=X_{0} satisfies (i) of theorem. The condition (ii) of the theorem follows directly from the property that the leaves of S{\mathcal{F}}_{S^{\prime}} are points, for all strata SS^{\prime} contained in X0X_{0}. ∎

3.0.2. Proof of Theorem 3.1.

Let hh be an analytic function defined in a neighborhood of xx such that h1(0)=Σh^{-1}(0)=\Sigma. Denote the distribution defined by dhdh and ωi,iI\omega_{i},i\in I, by 𝒦h{\vec{\mathcal{K}}_{h}}. Its singular locus equals Σ1=ΣΣh\Sigma_{1}=\Sigma\cup\Sigma_{h}, where

Σh={xVΣ;dh(x)𝒦h},\Sigma_{h}=\left\{x\in V\setminus\Sigma;dh(x)\in\vec{\mathcal{K}}_{h}^{\perp}\right\},

and 𝒦h{\vec{\mathcal{K}}_{h}} is integrable of corank r+1r+1 in its complement. We denote the induced foliation by h{\mathcal{F}}_{h}. Apply Theorem 3.2 to h{\mathcal{F}}_{h} and denote the set satisfying its statement by X1X_{1}.

Next, consider the leaves of the foliation induced by {\mathcal{F}} on Σh\Sigma_{h}, more precisely we stratify Σh\Sigma_{h} by a stratification regular with respect to 𝒦\vec{\mathcal{K}}. Note that hh is constant on the leaves of this foliation. We apply to the strata of this stratification Proposition 3.6. Let SS be a stratum from the conclusion of Proposition 3.6. It is of dimension dimS=rSr\dim S=r_{S}\leq r. It is clear that the union of such sets and X1X_{1} satisfies the claim of the theorem except (ii) and (iv).

The point (ii) follows for cc small from a general result, the local uniform bound of the volume of relatively compact subanalytic sets in subanalytic families, see e.g. [10, page 261] or [14, Théorème 1].

The transversality of point (iv) follows from (ii) of Theorem 3.2 and the subanalytic Sard theorem applied to the function hh restricted to the sets YiY_{i}. The set of critical values, being subanalytic and of measure zero has to be finite. We choose ϵ\epsilon smaller that the smallest positive critical value. To have the condition Yc=Yc¯YcZc\partial Y^{c}=\overline{Y^{c}}\setminus Y^{c}\subset Z^{c} we just add Y¯Y\overline{Y}\setminus Y to Z. ∎

4. Proof of Theorem 1.1

We follow the notation and we use several results given in [4, Section 3]. In particular, recall that the cotagent bundle TMT^{\ast}M is equipped with a canonical symplectic form ω\omega.

Assume that MM (of dimension n3n\geq 3) and Δ\Delta (of rank m2m\geq 2) are analytic and suppose for the sake of contradiction that there is x¯M\bar{x}\in M such that the set

AbnΔm(x¯)={γ(1)|γΩΔx¯ s.t. rankΔ(γ)=m}\mbox{Abn}^{m}_{\Delta}(\bar{x})=\Bigl{\{}\gamma(1)\,|\,\gamma\in\Omega_{\Delta}^{\bar{x}}\mbox{ s.t. }\mbox{rank}^{\Delta}(\gamma)=m\Bigr{\}}

has positive Lebesgue measure in MM. We equip MM with a complete smooth Riemannian metric gg. Let us now recall the setting provided by [4, Theorem 1.1]: there exist a subanalytic Whitney stratification 𝒮=(𝒮α)\mathcal{S}=(\mathcal{S}_{\alpha}) of Δ\Delta^{\perp}, three subanalytic distributions

𝒦𝒥TΔ\vec{\mathcal{K}}\subset\vec{\mathcal{J}}\subset\vec{\mathcal{I}}\subset T\Delta^{\perp}

adapted to 𝒮\mathcal{S} satisfying properties (i)-(iv). Then, denoting by 𝒮0\mathcal{S}_{0} the essential domain, that is the union of all strata of 𝒮\mathcal{S} of maximal dimension, and by Σ\Sigma its complement in Δ\Delta^{\perp} of dimension strictly less than 2nm=dim(Δ)2n-m=\dim(\Delta^{\perp}), [4, Theorem 1.1] implies that 𝒮0\mathcal{S}_{0} is an open set in Δ\Delta^{\perp}, Σ\Sigma is an analytic set in Δ\Delta^{\perp} of codimension at least 1, and 𝒦|𝒮0=𝒥|𝒮0=|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}=\vec{\mathcal{J}}_{|\mathcal{S}_{0}}=\vec{\mathcal{I}}_{|\mathcal{S}_{0}} is isotropic and integrable on 𝒮0\mathcal{S}_{0} of rank m0m_{0} verifying m0m(2)m_{0}\equiv m\,(2) and m0m2m_{0}\leq m-2. Note, furthermore, that [4, Proposition 3.6] combined with the contradiction assumption implies that m0>0m_{0}>0, that is, the distribution 𝒦\vec{\mathcal{K}} yields a non-trivial foliation over 𝒮0\mathcal{S}_{0} (in particular, n4n\geq 4 and m3m\geq 3). For every 𝔞𝒮0{\mathfrak{a}}\in\mathcal{S}_{0}, we denote by 𝔞𝒮0\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0} the leaf of the foliation generated by 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}} containing 𝔞{\mathfrak{a}}.

We start by considering a subset of AbnΔm(x¯)\mbox{Abn}^{m}_{\Delta}(\bar{x}) of positive measure with two extra properties (recall that we have supposed for contradiction that AbnΔm(x¯)\mbox{Abn}^{m}_{\Delta}(\bar{x}) has positive Lebesgue measure in MM):

Lemma 4.1.

There exist ¯>0\bar{\ell}>0 and a subset 𝒜¯M\bar{\mathcal{A}}\subset M of positive measure such that, for every point y𝒜¯y\in\bar{\mathcal{A}}, the intersection π1(y)𝒮0\pi^{-1}(y)\cap\mathcal{S}_{0}\neq\emptyset and there exists a singular horizontal curve of minimal rank of length ¯\leq\bar{\ell} (w.r.t gg) which joins x¯\bar{x} to yy, for which all abnormal lifts intersect the set Σ\Sigma.

Proof of Lemma 4.1.

Denote by 𝒜x¯𝒮0\mathcal{A}_{\bar{x}}^{\mathcal{S}_{0}} the set of points yy in AbnΔm(x¯)\mbox{Abn}^{m}_{\Delta}(\bar{x}) for which there is a curve γΩΔx¯\gamma\in\Omega_{\Delta}^{\bar{x}} of minimal rank with γ(1)=y\gamma(1)=y which admits an abnormal lift ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} such that ψ([0,1])𝒮0\psi([0,1])\subset\mathcal{S}_{0}. By construction, the set 𝒜x¯𝒮0\mathcal{A}_{\bar{x}}^{\mathcal{S}_{0}} is contained in the set

 Abn0(x¯):=𝔞(𝒮0)x¯π(𝔞),\mbox{ Abn}_{0}(\bar{x}):=\bigcup_{{\mathfrak{a}}\in(\mathcal{S}_{0})_{\bar{x}}}\pi\left(\mathcal{L}_{{\mathfrak{a}}}\right),

so by [4, Theorem 1.3] it has Lebesgue measure zero in MM. We set 𝒜x¯:=AbnΔm(x¯)𝒜x¯𝒮0\mathcal{A}_{\bar{x}}:=\mbox{Abn}^{m}_{\Delta}(\bar{x})\setminus\mathcal{A}_{\bar{x}}^{\mathcal{S}_{0}} and note that, without loss of generality, we may assume that 𝒜x¯\mathcal{A}_{\bar{x}} has positive measure in MM and that there is ¯>0\bar{\ell}>0 such that for every y𝒜x¯y\in\mathcal{A}_{\bar{x}} there is a singular horizontal curve of minimal rank of length ¯\leq\bar{\ell} (w.r.t gg) which joins x¯\bar{x} to yy for which all abnormal lifts intersect the set Σ\Sigma. Next, recall that π:TMM\pi:T^{*}M\rightarrow M denotes the canonical projection and set

𝒜x¯Σ:={y𝒜x¯|π1(y)ΔΣ}.\mathcal{A}_{\bar{x}}^{\Sigma}:=\left\{y\in\mathcal{A}_{\bar{x}}\,|\,\pi^{-1}(y)\cap\Delta^{\perp}\subset\Sigma\right\}.

We observe that the set 𝒜x¯ΣM\mathcal{A}_{\bar{x}}^{\Sigma}\subset M has Lebesgue measure zero in MM since otherwise Σ\Sigma, which is an analytic set in Δ\Delta^{\perp} of codimension at least 1 would have positive measure in Δ\Delta^{\perp}. Then, we set

𝒜¯:=𝒜x¯𝒜x¯ΣM,\bar{\mathcal{A}}:=\mathcal{A}_{\bar{x}}\setminus\mathcal{A}^{\Sigma}_{\bar{x}}\subset M,

which by construction has positive Lebesgue measure in MM. ∎

We now make a short interlude to introduce three objects which are going to be used in the proof, namely a complete Riemannian metric g~\tilde{g} over Δ\Delta^{\perp}, locally defined 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}-normal forms and transition maps, and a 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}-transverse measure.

The metric g~\tilde{g} over Δ\Delta^{\perp}: we can extend the Riemannian metric gg over MM into a complete smooth metric g~\tilde{g} “compatible” with gg over Δ\vec{\Delta} on Δ\Delta^{\perp}. As a matter of fact, we can define for every 𝔞Δ{\mathfrak{a}}\in\Delta^{\perp},

g~𝔞(ξ1,ξ2):=gπ(𝔞)(d𝔞π(ξ1),d𝔞π(ξ2))ξ1,ξ2Δ(𝔞),\tilde{g}_{{\mathfrak{a}}}(\xi_{1},\xi_{2}):=g_{\pi({\mathfrak{a}})}(d_{{\mathfrak{a}}}\pi(\xi_{1}),d_{{\mathfrak{a}}}\pi(\xi_{2}))\qquad\forall\xi_{1},\xi_{2}\in\vec{\Delta}({\mathfrak{a}}),

which is nondegenerate because Δ\vec{\Delta} is always transverse to the vertical fiber of the canonical projection π:TMM\pi:T^{\ast}M\to M, c.f. [4, Section 3.2], and extend g~\tilde{g} to the missing directions to obtain a complete smooth Riemannian metric on Δ\Delta^{\perp}. In the sequel, we denote by ||g~|\cdot|^{\tilde{g}} the norm given by g~\tilde{g} and by dg~d^{\tilde{g}} the geodesic distance with respect to g~\tilde{g}. Then, we denote by lengthg~\mbox{length}^{\tilde{g}} the length of an absolutely continuous curve ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} with respect to g~\tilde{g} and note that if ψ\psi is a lift of a singular horizontal path γ:[0,1]M\gamma:[0,1]\rightarrow M then

lengthg~(ψ)=lengthg(γ).\mbox{length}^{\tilde{g}}(\psi)=\mbox{length}^{g}(\gamma).

Local normal form and transition map: Fix a density point y¯𝒜¯{x¯}\bar{y}\in\bar{\mathcal{A}}\setminus\{\bar{x}\} together with some 𝔞¯𝒮0\bar{{\mathfrak{a}}}\in\mathcal{S}_{0} such that π(𝔞¯)=y¯\pi(\bar{{\mathfrak{a}}})=\bar{y}. By considering a local set of coordinates in an open neighborhood 𝒰M\mathcal{U}\subset M of y¯\bar{y}, we may assume that we have coordinates (y,q)(y,q) in T𝒰=𝒰×(n)T^{*}\mathcal{U}=\mathcal{U}\times(\mathbb{R}^{n})^{*} in such a way that the restriction of π\pi to T𝒰T^{*}\mathcal{U} is given by π(y,q)=y\pi(y,q)=y for all (y,q)T𝒰(y,q)\in T^{*}\mathcal{U}. Then, we let q¯Ty¯𝒰\bar{q}\in T_{\bar{y}}^{*}\mathcal{U} such that 𝔞¯=(y¯,q¯)\bar{{\mathfrak{a}}}=(\bar{y},\bar{q}), we set r:=2nmm0,r:=2n-m-m_{0}, and, since 𝒦\vec{\mathcal{K}} defines a foliation of dimension m0>0m_{0}>0 in 𝒮0\mathcal{S}_{0}, as noted in the beginning of the section, we may consider a foliation chart (𝒲,φ)(\mathcal{W},\varphi) of 𝔞¯\bar{{\mathfrak{a}}} such that 𝔞¯𝒲𝒮0T𝒰\bar{{\mathfrak{a}}}\in\mathcal{W}\subset\mathcal{S}_{0}\cap T^{*}\mathcal{U} and for which there are two open sets W1m0W^{1}\subset\mathbb{R}^{m_{0}} and W2rW^{2}\subset\mathbb{R}^{r} such that φ=(φ1,φ2):𝒲W:=W1×W2\varphi=(\varphi_{1},\varphi_{2}):\mathcal{W}\rightarrow W:=W^{1}\times W^{2} is an analytic diffeomorphism satisfying a¯:=φ(𝔞¯)=0\bar{a}:=\varphi(\bar{{\mathfrak{a}}})=0 and

(4.1) d𝔞φ(𝒦(𝔞))=K:=m0×{0}𝔞𝒲.\displaystyle d_{{\mathfrak{a}}}\varphi\bigl{(}\vec{\mathcal{K}}({\mathfrak{a}})\bigr{)}=\vec{K}:=\mathbb{R}^{m_{0}}\times\{0\}\qquad\forall{\mathfrak{a}}\in\mathcal{W}.

We note that, by construction, for every a=(a1,a2)Wa=(a_{1},a_{2})\in W, the plaque φ1(W1×{a2})\varphi^{-1}(W^{1}\times\{a_{2}\}) is contained in the leaf φ1(a)\mathcal{L}_{\varphi^{-1}(a)} of 𝒦\vec{\mathcal{K}} in 𝒮0\mathcal{S}_{0}. We also consider a family of local disjoint transverse sections to 𝒦\vec{\mathcal{K}} in 𝒲\mathcal{W} parametrized by the connected component of 𝔞¯𝒲\mathcal{L}_{\bar{{\mathfrak{a}}}}\cap\mathcal{W} containing 𝔞¯\bar{{\mathfrak{a}}} and given by (see Figure 2)

𝒯𝔞:=φ1({φ1(𝔞)}×W2)𝔞𝔞¯𝒲.\mathcal{T}_{{\mathfrak{a}}}:=\varphi^{-1}\left(\bigl{\{}\varphi_{1}({\mathfrak{a}})\bigr{\}}\times W^{2}\right)\qquad\forall{\mathfrak{a}}\in\mathcal{L}_{\bar{{\mathfrak{a}}}}\cap\mathcal{W}.
Refer to caption
Figure 2. Local foliation chart and transverse sections

Up to shrinking 𝒲\mathcal{W}, this family of sections allows us to define a local transition maps parametrized by the connected component of 𝔞¯𝒲\mathcal{L}_{\bar{{\mathfrak{a}}}}\cap\mathcal{W} containing 𝔞¯\bar{{\mathfrak{a}}}, that is, diffeomorphisms T𝔞¯,𝔞:𝒯𝔞¯𝒯𝔞T^{\bar{{\mathfrak{a}}},{\mathfrak{a}}}:\mathcal{T}_{\bar{{\mathfrak{a}}}}\to\mathcal{T}_{{\mathfrak{a}}} for all 𝔞𝔞¯𝒲{\mathfrak{a}}\in\mathcal{L}_{\bar{{\mathfrak{a}}}}\cap\mathcal{W} defined by

(4.2) T𝔞¯,𝔞(𝔟):=φ1({φ1(𝔞)}×π2(φ(𝔟)))𝔟𝒯𝔞¯\displaystyle T^{\bar{{\mathfrak{a}}},{\mathfrak{a}}}({\mathfrak{b}}):=\varphi^{-1}\Bigl{(}\left\{\varphi_{1}({\mathfrak{a}})\right\}\times\pi_{2}\left(\varphi({\mathfrak{b}})\right)\Bigr{)}\qquad\forall{\mathfrak{b}}\in\mathcal{T}_{\bar{{\mathfrak{a}}}}

Given a subset Γ𝔞¯\Gamma^{\bar{{\mathfrak{a}}}} of 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}}, we will sometimes abuse notation and write

Γ𝔞𝔞¯:=T𝔞¯,𝔞(Γ𝔞¯).\Gamma^{\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}}:=T^{\bar{{\mathfrak{a}}},{\mathfrak{a}}}(\Gamma^{\bar{{\mathfrak{a}}}}).

Transverse metric: We define a 2l2l-form η\eta on Δ\Delta^{\perp} by

η:=(ω)lwithl:=r2,\eta:=\bigl{(}\omega^{\perp}\bigr{)}^{l}\quad\mbox{with}\quad l:=\frac{r}{2},

where rr is the co-rank of 𝒦|𝒮0\vec{\mathcal{K}}|_{\mathcal{S}_{0}} in respect to Δ\Delta^{\perp}, that is, r=2nmm0r=2n-m-m_{0}. The following lemma follows essentially from [4, Proposition 3.2(ii)] and the assumption that 𝒦|S0\vec{\mathcal{K}}_{|S_{0}} is splittable, cf. §\S 2.

Lemma 4.2.

There are a 𝒦\vec{\mathcal{K}}-transverse section 𝒯𝔞¯𝒮0\mathcal{T}_{\bar{{\mathfrak{a}}}}\subset\mathcal{S}_{0} centered at 𝔞¯\bar{{\mathfrak{a}}} and a compact set 𝒜~𝔞¯𝒯𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}\subset\mathcal{T}_{\bar{{\mathfrak{a}}}} such that the following properties are satisfied:

  • (i)

    The set 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}} has positive measure with respect to the volume form η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}}.

  • (ii)

    For every 𝔞𝒜~𝔞¯{\mathfrak{a}}\in\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}, there is an absolutely continuous curve ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} such that

    ψ(0)=𝔞,ψ(1)Σ,lengthg~(ψ)¯+1andψ(t)𝔞𝒮0t[0,1).\psi(0)={\mathfrak{a}},\quad\psi(1)\in\Sigma,\quad\mbox{\em length}^{\tilde{g}}(\psi)\leq\bar{\ell}+1\quad\mbox{and}\quad\psi(t)\in\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0}\quad\forall t\in[0,1).
  • (iii)

    For any distinct points 𝔞,𝔞𝒜~𝔞¯{\mathfrak{a}},{\mathfrak{a}}^{\prime}\in\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}, 𝔞{\mathfrak{a}} and 𝔞{\mathfrak{a}}^{\prime} are not (𝒦,2¯+5)(\vec{\mathcal{K}},2\bar{\ell}+5)-related.

Proof of Lemma 4.2.

Recall that y¯\bar{y} is a density point of 𝒜¯{x¯}\bar{\mathcal{A}}\setminus\{\bar{x}\} and 𝔞¯𝒮0\bar{{\mathfrak{a}}}\in\mathcal{S}_{0} satisfies π(𝔞¯)=y¯\pi(\bar{{\mathfrak{a}}})=\bar{y}. Consider the notation of the local normal form above and for every y𝒰y\in\mathcal{U}, denote by 𝒱y\mathcal{V}_{y} the vertical fiber in Δ\Delta^{\perp} over yy given by

𝒱y:=π1({y})T𝔞¯Δ,\mathcal{V}_{y}:=\pi^{-1}(\{y\})\cap T_{\bar{{\mathfrak{a}}}}\Delta^{\perp},

which coincides with a vector space Vy\vec{V}_{y} of dimension nmn-m with the origin removed. Since 𝒦(𝔞¯)T𝔞¯𝒱y¯={0}\vec{\mathcal{K}}(\bar{{\mathfrak{a}}})\cap T_{\bar{{\mathfrak{a}}}}\mathcal{V}_{\bar{y}}=\{0\} (see [4, Theorem 1.1(i) and Equation (3.5)]), there is a vector space T𝔞¯Δ\vec{\mathcal{H}}\subset T_{\bar{{\mathfrak{a}}}}\Delta^{\perp} of dimension nn containing 𝒦(𝔞¯)\vec{\mathcal{K}}(\bar{{\mathfrak{a}}}) which is transverse to T𝔞¯𝒱y¯=Vy¯T_{\bar{{\mathfrak{a}}}}\mathcal{V}_{\bar{y}}=\vec{V}_{\bar{y}} in T𝔞¯ΔT_{\bar{{\mathfrak{a}}}}\Delta^{\perp}, that is, such that

Vy¯=T𝔞¯Δand𝒦(𝔞¯).\vec{\mathcal{H}}\oplus\vec{V}_{\bar{y}}=T_{\bar{{\mathfrak{a}}}}\Delta^{\perp}\quad\mbox{and}\quad\vec{\mathcal{K}}(\bar{{\mathfrak{a}}})\subset\vec{\mathcal{H}}.

Then, we consider a vector space 𝒫\vec{\mathcal{P}}\subset\vec{\mathcal{H}} such that

(4.3) 𝒦(𝔞¯)𝒫=\displaystyle\vec{\mathcal{K}}(\bar{{\mathfrak{a}}})\oplus\vec{\mathcal{P}}=\vec{\mathcal{H}}

and define the vector spaces 𝒬T𝔞¯Δ\vec{\mathcal{Q}}\subset T_{\bar{{\mathfrak{a}}}}\Delta^{\perp} and Q,H,P2nm\vec{Q},\vec{H},\vec{P}\subset\mathbb{R}^{2n-m} by

𝒬:=𝒫Vy¯,Q:=d𝔞¯φ(𝒬),H:=d𝔞¯φ(),P:=d𝔞¯φ(𝒫).\vec{\mathcal{Q}}:=\vec{\mathcal{P}}\oplus\vec{V}_{\bar{y}},\quad\vec{Q}:=d_{\bar{{\mathfrak{a}}}}\varphi\bigl{(}\vec{\mathcal{Q}}\bigr{)},\quad\vec{H}:=d_{\bar{{\mathfrak{a}}}}\varphi\bigl{(}\vec{\mathcal{H}}\bigr{)},\quad\vec{P}:=d_{\bar{{\mathfrak{a}}}}\varphi\bigl{(}\vec{\mathcal{P}}\bigr{)}.

By construction, \vec{\mathcal{H}} and H\vec{H} have dimension nn, 𝒫\vec{\mathcal{P}} and P\vec{P} have dimension nm0n-m_{0}, 𝒬\vec{\mathcal{Q}} and Q\vec{Q} have dimension r=2nmm0r=2n-m-m_{0} and, remembering (4.1)-(4.3), we have

(4.4) KP=H,𝒦(𝔞¯)𝒬=T𝔞¯Δ,KQ=2nm.\displaystyle\vec{K}\oplus\vec{P}=\vec{H},\quad\vec{\mathcal{K}}(\bar{{\mathfrak{a}}})\oplus\vec{\mathcal{Q}}=T_{\bar{{\mathfrak{a}}}}\Delta^{\perp},\quad\vec{K}\oplus\vec{Q}=\mathbb{R}^{2n-m}.

Then, we define two nn-dimensional open smooth manifolds HWH\subset W and 𝒲\mathcal{H}\subset\mathcal{W} by

H:=HWand:=φ1(H),H:=\vec{H}\cap W\quad\mbox{and}\quad\mathcal{H}:=\varphi^{-1}(H),

and note that the restriction of π\pi to \mathcal{H} is a submersion at 𝔞¯\bar{{\mathfrak{a}}}. Therefore, there is a smooth submanifold II of WW of dimension nn containing a¯=0\bar{a}=0 of the form (|||\cdot| stands for the Euclidean norm in m0\mathbb{R}^{m_{0}} or 2nm\mathbb{R}^{2n-m})

I={(a1,0)+p|a1W1,pP,|a1|<δ,|p|<δ}H,I=\left\{(a_{1},0)+p\,|\,a_{1}\in W^{1},\,p\in\vec{P},\,|a_{1}|<\delta,|p|<\delta\right\}\subset H,

with δ>0\delta>0, such that the mapping

F=π|::=φ1(I):=π()F=\pi_{|\mathcal{I}}\,:\,\mathcal{I}:=\varphi^{-1}(I)\,\longrightarrow\,\mathcal{E}:=\pi(\mathcal{I})

is a smooth diffeomorphism.

Refer to caption
Figure 3. A picture showing the sets K,H,P,Q,I\vec{K},\vec{H},\vec{P},\vec{Q},I and W¯1\bar{W}^{1}

Then, we denote by W¯1\bar{W}^{1} the set of a1W1a_{1}\in W^{1} with |a1|δ/2|a_{1}|\leq\delta/2 and for each a1W¯1a_{1}\in\bar{W}_{1}, we define the sets Pa1HP_{a_{1}}\subset H, 𝒫a1\mathcal{P}_{a_{1}}\subset\mathcal{H}, a1\mathcal{E}_{a_{1}}\subset\mathcal{E}, 𝒬a1𝒲\mathcal{Q}_{a_{1}}\subset\mathcal{W} and Qa1WQ_{a_{1}}\subset W by

Pa1:=((a1,0)+P)I,𝒫a1:=φ1(Pa1),a1:=F(𝒫a1).P_{a_{1}}:=\bigl{(}(a_{1},0)+\vec{P}\bigr{)}\cap I,\quad\mathcal{P}_{a_{1}}:=\varphi^{-1}\left(P_{a_{1}}\right),\quad\mathcal{E}_{a_{1}}:=F\left(\mathcal{P}_{a_{1}}\right).
𝒬a1:={(y,q)+(0,h)|(y,q)𝒫a1,hVy,|h|<δ}andQa1:=φ(𝒬a1).\mathcal{Q}_{a_{1}}:=\left\{(y,q)+(0,h)\,|\,(y,q)\in\mathcal{P}_{a_{1}},\,h\in\vec{V}_{y},\,|h|<\delta\right\}\quad\mbox{and}\quad Q_{a_{1}}:=\varphi\left(\mathcal{Q}_{a_{1}}\right).

By construction, for each a1W¯1a_{1}\in\bar{W}^{1}, the set Pa1P_{a_{1}} is an open smooth submanifold of WW of dimension nm0n-m_{0}, the set 𝒫a1\mathcal{P}_{a_{1}} is an open smooth submanifold of 𝒲\mathcal{W} of dimension nm0n-m_{0} and the set a1\mathcal{E}_{a_{1}} is an open smooth submanifold of 𝒰\mathcal{U} of dimension nm0n-m_{0}. In fact, the collection {𝒫a1}a1W¯1\{\mathcal{P}_{a_{1}}\}_{a_{1}\in\bar{W}_{1}} defines a collection of pairwise disjoint slices (or plaques to use the terminology of foliations recalled above) that cover \mathcal{I}\subset\mathcal{H} and projects to the collection of pairwise disjoint slices {a1}a1W¯1\{\mathcal{E}_{a_{1}}\}_{a_{1}\in\bar{W}_{1}} covering =π()M\mathcal{E}=\pi(\mathcal{I})\subset M. Furthermore, since by 𝒫Vy¯=d0φ1(P)T𝔞¯𝒱y¯={0}\vec{\mathcal{P}}\cap\vec{V}_{\bar{y}}=d_{0}\varphi^{-1}(\vec{P})\cap T_{\bar{{\mathfrak{a}}}}\mathcal{V}_{\bar{y}}=\{0\} the mapping

(𝔞,h){((y,q),h)|(y,q)𝒫0,hVy}𝔞+(0,h)Δ({\mathfrak{a}},h)\in\left\{((y,q),h)\,|\,(y,q)\in\mathcal{P}_{0},\,h\in\vec{V}_{y}\right\}\,\longmapsto\,{\mathfrak{a}}+(0,h)\in\Delta^{\perp}

is an immersion at (𝔞¯,0)(\bar{{\mathfrak{a}}},0) valued in 𝒬0\mathcal{Q}_{0} and since the mapping

aQ(0,π2(a))T0({0}×W2)a\in\vec{Q}\,\longmapsto\bigl{(}0,\pi^{2}(a)\bigr{)}\in T_{0}\bigl{(}\{0\}\times W^{2}\bigr{)}

is a linear isomorphism (by (4.4) we have KQ=2nm\vec{K}\oplus\vec{Q}=\mathbb{R}^{2n-m}), we may assume by taking δ>0\delta>0 small enough that, for each a1W¯1a_{1}\in\bar{W}^{1}, the sets 𝒬a1\mathcal{Q}_{a_{1}} and Qa1Q_{a_{1}} are open smooth manifolds of dimension rr and that the mapping

Ga1:aQa1(0,π2(a))T0({0}×W2)G_{a_{1}}\,:\,a\in Q_{a_{1}}\,\longmapsto\,\bigl{(}0,\pi^{2}(a)\bigr{)}\in T_{0}\bigl{(}\{0\}\times W^{2}\bigr{)}

is a smooth diffeomorphism from Qa1Q_{a_{1}} onto its image Ga1(Qa1)G_{a_{1}}(Q_{a_{1}}). Thus, {𝒬a1}a1W¯1\{\mathcal{Q}_{a_{1}}\}_{a_{1}\in\bar{W}_{1}} defines a collection of pairwise disjoint slices that extends the collection {𝒫a1}a1W¯1\{\mathcal{P}_{a_{1}}\}_{a_{1}\in\bar{W}_{1}} (for each a1W¯1a_{1}\in\bar{W}^{1}, 𝒫a1𝒬a1\mathcal{P}_{a_{1}}\subset\mathcal{Q}_{a_{1}}) and covers a neigborhood of 𝔞¯\bar{{\mathfrak{a}}} in Δ\Delta^{\perp}. Furthermore, we observe that for every a1W¯1a_{1}\in\bar{W}^{1} and every aQa1a\in Q_{a_{1}}, the two points aa and b=Ga1(a)b=G_{a_{1}}(a) have the same coordinate in W2W^{2} so that their images by φ1\varphi^{-1}, φ1(a)\varphi^{-1}(a) and φ1(b)\varphi^{-1}(b), belong to the same plaque and to the same leaf of the foliation defined by 𝒦\vec{\mathcal{K}} in 𝒲\mathcal{W}, so, by the construction made before the statement of the lemma, the points φ1(a)𝒲\varphi^{-1}(a)\in\mathcal{W} and φ1(b)𝒯𝔞¯\varphi^{-1}(b)\in\mathcal{T}_{\bar{{\mathfrak{a}}}} can be connected through a smooth curve horizontal with respect to 𝒦\vec{\mathcal{K}} of length (w.r.t g~\tilde{g}) less than 1. In other words, for every a1W¯1a_{1}\in\bar{W}^{1}, the mapping φ1Ga1φ\varphi^{-1}\circ G_{a_{1}}\circ\varphi acts as a projection from the slice 𝒬a1\mathcal{Q}_{a_{1}} to 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} along horizontal curves with respect to 𝒦\vec{\mathcal{K}} (and with length less than 1). We are now ready to conclude the proof of the Lemma which consists in applying Fubini’s Theorem to select a slice a~1\mathcal{E}_{\tilde{a}_{1}} in {a1}a1W¯1\{\mathcal{E}_{a_{1}}\}_{a_{1}\in\bar{W}_{1}} whose intersection with 𝒜¯\bar{\mathcal{A}} has positive (nm0)(n-m_{0})-dimensional Lebesgue measure, to lift the intersection to 𝒬a~1Δ\mathcal{Q}_{\tilde{a}_{1}}\subset\Delta^{\perp}, and to project it to 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} by action of φ1Ga~1φ\varphi^{-1}\circ G_{\tilde{a}_{1}}\circ\varphi.

By construction, the sets Pa1P_{a_{1}} as well as 𝒫a1\mathcal{P}_{a_{1}}, a1\mathcal{E}_{a_{1}}, with a1W¯1a_{1}\in\bar{W}^{1} are pairwise disjoint and satisfy

a1W¯1Pa1=Ia1W¯1𝒫a1=,a1W¯1a1=.\bigcup_{a_{1}\in\bar{W}^{1}}P_{a_{1}}=I\quad\bigcup_{a_{1}\in\bar{W}^{1}}\mathcal{P}_{a_{1}}=\mathcal{I},\quad\bigcup_{a_{1}\in\bar{W}^{1}}\mathcal{E}_{a_{1}}=\mathcal{E}.

Since y¯\bar{y} is a density point of 𝒜¯\bar{\mathcal{A}}, by Fubini’s Theorem, we infer that there is a~1W¯1\tilde{a}_{1}\in\bar{W}^{1} such that the (nm0)(n-m_{0})-dimensional Lebesgue measure of the set

Θ:=𝒜¯a~1a~1\Theta:=\bar{\mathcal{A}}\cap\mathcal{E}_{\tilde{a}_{1}}\subset\mathcal{E}_{\tilde{a}_{1}}

is positive. In fact, by taking a compact subset of Θ\Theta of positive measure, we may indeed assume that Θ\Theta is compact. By construction, for every θΘ\theta\in\Theta, there is an horizontal path γθΩΔx¯\gamma^{\theta}\in\Omega_{\Delta}^{\bar{x}} of length ¯\leq\bar{\ell} (w.r.t gg) such that γθ(1)=θ\gamma^{\theta}(1)=\theta, rankΔ(γ)=m\mbox{rank}^{\Delta}(\gamma)=m and for which all abnormal lifts meet the set Σ\Sigma. Hence, by [4, Proposition 3.4], for every pΔθp\in\Delta^{\perp}_{\theta}, γθ\gamma^{\theta} admits an abnormal lift ψθ,p:[0,1]Δ\psi^{\theta,p}:[0,1]\rightarrow\Delta^{\perp} such that ψθ,p(1)=(θ,p)\psi^{\theta,p}(1)=(\theta,p) and ψx,p([0,1])Σ\psi^{x,p}([0,1])\cap\Sigma\neq\emptyset. Thus, we obtain that any 𝔞{\mathfrak{a}} in the set

Θ~:={F1(θ)+(0,h)|θΘ,hVθ,|h|δ}\tilde{\Theta}:=\left\{F^{-1}(\theta)+(0,h)\,|\,\theta\in\Theta,\,h\in\vec{V}_{\theta},\,|h|\leq\delta\right\}

can be joined to Σ\Sigma by a curve of length ¯\leq\bar{\ell}. By Fubini’s Theorem, the set Θ~\tilde{\Theta} is a compact set of positive measure in the manifold 𝒬a1~\mathcal{Q}_{\tilde{a_{1}}}, thus its image by φ\varphi, φ(Θ~)\varphi(\tilde{\Theta}), is a compact set of positive measure in the manifold Qa1~Q_{\tilde{a_{1}}}, the image of φ(Θ~)\varphi(\tilde{\Theta}) by Ga~1G_{\tilde{a}_{1}}, Λ:=(Ga~1φ)(Θ~)\Lambda:=(G_{\tilde{a}_{1}}\circ\varphi)(\tilde{\Theta}) has positive measure in {0}×W2\{0\}\times W^{2} and the image of Λ\Lambda by φ1\varphi^{-1} has positive measure in 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}}. By construction, any point of φ1(Λ)\varphi^{-1}(\Lambda) can be joined to a point of Σ\Sigma by an absolutely continuous curve horizontal with respect to 𝒦\vec{\mathcal{K}} of length (w.r.t g~\tilde{g}) l¯+1\leq\bar{l}+1. By assumption of splittability, we can select in φ1(Λ)\varphi^{-1}(\Lambda) a compact subset 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}} of positive measure satisfying the same property and whose points are not (𝒦,2¯+5)(\vec{\mathcal{K}},2\bar{\ell}+5)-related. We complete the proof by applying [4, Proposition 3.2(ii)]. ∎

The next result combines the geometrical framework of Lemma 4.2 with a compactness argument and the witness section given by Theorem 3.1.

Lemma 4.3.

There are a point 𝔞^Σ\hat{{\mathfrak{a}}}\in\Sigma, a compact set 𝒜ˇ𝔞¯𝒜~𝔞¯𝒯𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}\subset\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}\subset\mathcal{T}_{\bar{{\mathfrak{a}}}}, a relatively compact open neighborhood VΔV\subset\Delta^{\perp} of 𝔞^\hat{{\mathfrak{a}}}, a compact set ΣˇΣV\check{\Sigma}\subset\Sigma\cap V, a real analytic function h:V[0,)h:V\to[0,\infty), a subanalytic set XVΣX\subset V\setminus\Sigma and C,ν,ϵ>0C,\nu,\epsilon>0 such that the following properties are satisfied:

  • (i)

    The set h1(0)h^{-1}(0) is equal to ΣV\Sigma\cap V.

  • (ii)

    For every 0<c<ϵ0<c<\epsilon, the subanalytic set Xc:=Xh1(c)X^{c}:=X\cap h^{-1}(c) has rr-dimensional volume with respect to g~\tilde{g} bounded by CC. In particular, XcX^{c} is a rr-dimensional set and XX is a (r+1)(r+1)-dimensional set.

  • (iii)

    For every 0<c<ϵ0<c<\epsilon and for every 𝔞h1(c)VΣ{\mathfrak{a}}\in h^{-1}(c)\subset V\setminus\Sigma, there is a smooth curve α:[0,1]VΣ\alpha:[0,1]\rightarrow V\setminus\Sigma which is contained in 𝔞h1(c)\mathcal{L}_{{\mathfrak{a}}}\cap h^{-1}(c) such that

    α(0)=𝔞,α(1)Xc,andlengthg~(α)1.\alpha(0)={\mathfrak{a}},\quad\alpha(1)\in X^{c},\quad\mbox{and}\quad\mbox{\em length}^{\tilde{g}}(\alpha)\leq 1.
  • (iv)

    For every 0<c<ϵ0<c<\epsilon, we can decompose XcX^{c} as the union of two disjoint subanalytic sets YcY^{c} and ZcZ^{c}, such that ZcZ^{c} has dimension <r<r, and YcY^{c} is the union of finitely many smooth subanalytic sets YicY_{i}^{c}, with iIci\in I^{c}, of dimension rr such that

    Yc=Yc¯YcZc and T𝔞Δ=T𝔞Yic+𝒦(𝔞)𝔞Yic,iIc.\partial Y^{c}=\overline{Y^{c}}\setminus Y^{c}\subset Z^{c}\quad\text{ and }\quad T_{{\mathfrak{a}}}\Delta^{\perp}=T_{{\mathfrak{a}}}Y_{i}^{c}+\vec{\mathcal{K}}({\mathfrak{a}})\qquad\forall{\mathfrak{a}}\in Y_{i}^{c},\,\forall i\in I_{c}.
  • (v)

    For all 𝔞𝒜ˇ𝔞¯{\mathfrak{a}}\in\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}, there is an absolutely continuous curve ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} such that

    ψ(0)=𝔞,ψ(1)Σˇ,lengthg~(ψ)¯+1andψ(t)𝔞𝒮0t[0,1).\psi(0)={\mathfrak{a}},\quad\psi(1)\in\check{\Sigma},\quad\mbox{\em length}^{\tilde{g}}(\psi)\leq\bar{\ell}+1\quad\mbox{and}\quad\psi(t)\in\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0}\quad\forall t\in[0,1).
  • (vi)

    The set 𝒜ˇ𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}} has measure ν\geq\nu with respect to the volume form η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}}.

Moreover, there is a continuous function δ:[0,)[0,)\delta:[0,\infty)\rightarrow[0,\infty) with δ(0)=0\delta(0)=0 such that for every 0<c<ϵ0<c<\epsilon and every 𝔞h1(c){\mathfrak{a}}\in h^{-1}(c),

(4.5) |η𝔞(ξ1,,ξd)|δ(c)|ξ1|g~|ξd|g~ξ1,,ξdT𝔞Δ.\displaystyle\left|\eta_{{\mathfrak{a}}}\left(\xi_{1},\ldots,\xi_{d}\right)\right|\leq\delta(c)|\xi_{1}|^{\tilde{g}}\cdots|\xi_{d}|^{\tilde{g}}\qquad\forall\xi_{1},\ldots,\xi_{d}\in T_{{\mathfrak{a}}}\Delta^{\perp}.
Proof of Lemma 4.3.

Let N:=ΔN:=\Delta^{\perp} be the real-analytic manifold of dimension 2nm2n-m equipped with the singular analytic foliation \mathcal{F} of generic corank r=2nmm0r=2n-m-m_{0} with singular set Σ\Sigma and Σ\mathcal{B}\subset\Sigma the set of 𝔞Σ{\mathfrak{a}}\in\Sigma for which there is an absolutely continuous curve ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} such that

ψ(0)𝒜~𝔞¯,ψ(1)=𝔞,lengthg~(ψ)¯+1andψ(t)ψ(0)𝒮0t[0,1),\displaystyle\psi(0)\in\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}},\quad\psi(1)={\mathfrak{a}},\quad\mbox{length}^{\tilde{g}}(\psi)\leq\bar{\ell}+1\quad\mbox{and}\quad\psi(t)\in\mathcal{L}_{\psi(0)}\subset\mathcal{S}_{0}\quad\forall t\in[0,1),

where 𝒜~𝔞¯𝒯𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}\subset\mathcal{T}_{\bar{{\mathfrak{a}}}} is the set provided by Lemma 4.2. The compactness of 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}} together with the closedness of Σ\Sigma and the upper bound on the length of curves (with the completeness of g~\tilde{g}) imply that \mathcal{B} is a compact subset of Σ\Sigma. By Theorem 3.1 applied with =1\ell=1, for every 𝔞{\mathfrak{a}}\in\mathcal{B}, there are a relatively compact open neighborhood V𝔞V_{{\mathfrak{a}}} of 𝔞{\mathfrak{a}} in N=ΔN=\Delta^{\perp}, a real-analytic function h𝔞:V𝔞[0,)h_{{\mathfrak{a}}}:V_{{\mathfrak{a}}}\to[0,\infty), a subanalytic set X𝔞V𝔞ΣX_{{\mathfrak{a}}}\subset V_{{\mathfrak{a}}}\setminus\Sigma and C𝔞>0C_{{\mathfrak{a}}}>0 such that the properties (i)-(iv) of Theorem 3.1 are satisfied. Pick for each 𝔞{\mathfrak{a}}\in\mathcal{B} a compact neighborhood Vˇ𝔞V𝔞\check{V}_{{\mathfrak{a}}}\subset V_{{\mathfrak{a}}} of 𝔞{\mathfrak{a}} and consider by compactness of \mathcal{B} a finite family {𝔞i}iI\{{\mathfrak{a}}_{i}\}_{i\in I} such that

(4.6) iIVˇ𝔞iiIV𝔞i.\displaystyle\mathcal{B}\subset\bigcup_{i\in I}\check{V}_{{\mathfrak{a}}_{i}}\subset\bigcup_{i\in I}V_{{\mathfrak{a}}_{i}}.

Then, for every iIi\in I, denote by 𝒜~i𝔞¯\tilde{\mathcal{A}}_{i}^{\bar{{\mathfrak{a}}}} the set of 𝔞𝒜~𝔞¯{\mathfrak{a}}\in\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}} for which there is an absolutely continuous curve ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} such that

ψ(0)=𝔞,ψ(1)Σˇi,lengthg~(ψ)¯+1andψ(t)𝔞𝒮0t[0,1),\displaystyle\psi(0)={\mathfrak{a}},\quad\psi(1)\in\check{\Sigma}_{i},\quad\mbox{length}^{\tilde{g}}(\psi)\leq\bar{\ell}+1\quad\mbox{and}\quad\psi(t)\in\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0}\quad\forall t\in[0,1),

with Σˇi:=ΣVˇ𝔞i\check{\Sigma}_{i}:=\Sigma\cap\check{V}_{{\mathfrak{a}}_{i}}\cap\mathcal{B}. We claim that each set 𝒜~i𝔞¯\tilde{\mathcal{A}}_{i}^{\bar{{\mathfrak{a}}}} is a Borel subset of 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}. As a matter of fact, for each iIi\in I, we can write

𝒜~i𝔞¯=k𝒜~i,k𝔞¯,\tilde{\mathcal{A}}_{i}^{\bar{{\mathfrak{a}}}}=\bigcap_{k\in\mathbb{N}^{*}}\tilde{\mathcal{A}}_{i,k}^{\bar{{\mathfrak{a}}}},

where for each kk\in\mathbb{N}^{*}, the set 𝒜~i,k𝔞¯\tilde{\mathcal{A}}_{i,k}^{\bar{{\mathfrak{a}}}} is defined as the set of 𝔞𝒜~𝔞¯{\mathfrak{a}}\in\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}} for which there is an absolutely continuous curve ψ:[0,1]𝔞𝒮0\psi:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}}\in\mathcal{S}_{0} such that

ψ(0)=𝔞,ψ(1)B1/kg~(Σˇi)(VΣ),lengthg~(ψ)<¯+1,\displaystyle\psi(0)={\mathfrak{a}},\quad\psi(1)\in B_{1/k}^{\tilde{g}}\left(\check{\Sigma}_{i}\right)\cap\left(V\setminus\Sigma\right),\quad\mbox{length}^{\tilde{g}}(\psi)<\bar{\ell}+1,

with

B1/kg~(Σˇi):={𝔞Δ|dg~(𝔞,Σˇi)<1k}.B_{1/k}^{\tilde{g}}\left(\check{\Sigma}_{i}\right):=\left\{{\mathfrak{a}}^{\prime}\in\Delta^{\perp}\,|\,d^{\tilde{g}}\left({\mathfrak{a}}^{\prime},\check{\Sigma}_{i}\right)<\frac{1}{k}\right\}.

By regularity of 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}, each set 𝒜~i,k𝔞¯\tilde{\mathcal{A}}_{i,k}^{\bar{{\mathfrak{a}}}} is open in 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}, so we infer that each 𝒜~i𝔞¯\tilde{\mathcal{A}}_{i}^{\bar{{\mathfrak{a}}}} is a Borel subset of 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}. Furthermore, by construction of \mathcal{B}, (4.6) and Lemma 4.2 (ii), we have

𝒜~𝔞¯=iI𝒜~i𝔞¯.\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}=\bigcup_{i\in I}\tilde{\mathcal{A}}_{i}^{\bar{{\mathfrak{a}}}}.

As a consequence, since 𝒜~𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}} has positive measure with respect to the volume form η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}} (Lemma 4.2 (i)), there is iIi\in I such that 𝒜~i𝔞¯\tilde{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{i} and a compact subset 𝒜ˇi𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{i} of it satisfy the same property. We conclude the proof of (i)-(vi) by setting 𝒜ˇ𝔞¯:=𝒜ˇi𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}:=\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{i}, V:=V𝔞iV:=V_{{\mathfrak{a}}_{i}}, Σˇ:=Σˇi\check{\Sigma}:=\check{\Sigma}_{i}, h:=h𝔞ih:=h_{{\mathfrak{a}}_{i}}, X:=X𝔞iX:=X_{{\mathfrak{a}}_{i}}, C:=C𝔞iC:=C_{{\mathfrak{a}}_{i}} and ν\nu the volume of 𝒜ˇ𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}} with respect to η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}}.

The second part of the proof (4.5) follows from [4, Proposition 3.2(ii)]. By [4, Theorem 1.1(iv)], we have

dim(ker(ω𝔞))m0+2𝔞Σ.\dim\left(\mbox{ker}\bigl{(}\omega^{\perp}_{{\mathfrak{a}}}\bigr{)}\right)\geq m_{0}+2\qquad\forall{\mathfrak{a}}\in\Sigma.

Therefore, by [4, Proposition 3.2(iii)], we have η𝔞=0\eta_{{\mathfrak{a}}}=0 for all 𝔞Σˇ{\mathfrak{a}}\in\check{\Sigma} and we can conclude by regularity of hh near the compact set ΣˇV\check{\Sigma}\subset V. ∎

The idea of our proof consists now in obtaining a contradiction from the construction of an homotopy sending smoothly the points of a small neighborhood of a set 𝒜ˇ𝔞¯,c𝒜ˇ𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}\subset\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}} in 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} to an open subset of YcY^{c} for c>0c>0 small enough. Since this homotopy has to preserve the leaves of 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}, we perform the construction by following the minimizing geodesics from 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} to YcY^{c} with respect to some complete metric on 𝒮0\mathcal{S}_{0} that needs to be built (note that g~\tilde{g} is not complete when restricted to 𝒮0\mathcal{S}_{0}). The next Lemma formalizes this framework:

Lemma 4.4.

For every 0<c<ϵ0<c<\epsilon, there are a smooth Riemannian metric g~c\tilde{g}^{c} on 𝒮0\mathcal{S}_{0} and a compact set 𝒜ˇ𝔞¯,c𝒜ˇ𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}\subset\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}} satisfying the following properties:

  • (i)

    The Riemannian manifold (𝒮0,g~c)(\mathcal{S}_{0},\tilde{g}^{c}) is complete.

  • (ii)

    For every 𝔞𝒜ˇ𝔞¯,c{\mathfrak{a}}\in\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}, there is an absolutely continuous curve ψ:[0,1]𝔞𝒮0\psi:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0} such that (where YcY^{c} is defined in Lemma 4.3(iv))

    ψ(0)=𝔞,ψ(1)Ycandlengthg~(ψ)=lengthg~c(ψ)<¯+2.\psi(0)={\mathfrak{a}},\quad\psi(1)\in Y^{c}\quad\mbox{and}\quad\mbox{\em length}^{\tilde{g}}(\psi)=\mbox{\em length}^{\tilde{g}^{c}}(\psi)<\bar{\ell}+2.
  • (iii)

    The set 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c} has measure ν/4\geq\nu/4 with respect to the volume form η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}}.

  • (iv)

    Let 𝒞c𝒮0\mathcal{C}^{c}\subset\mathcal{S}_{0} be the set of points 𝔞𝒮0{\mathfrak{a}}\in\mathcal{S}_{0} for which there is an absolutely continuous curve ψ:[0,1]𝔞\psi:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}} of length ¯+2\leq\bar{\ell}+2 with respect to g~c\tilde{g}^{c} joining 𝔞{\mathfrak{a}} to a point of Zc¯\overline{Z^{c}} (defined in Lemma 4.3(iv)). Then 𝒞c\mathcal{C}^{c} is closed and does not intersect 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}.

Proof of Lemma 4.4.

Since Σ\Sigma is a closed subset of Δ\Delta^{\perp}, we can pick a smooth function F:Δ[0,)F:\Delta^{\perp}\rightarrow[0,\infty) such that

Σ=F1({0})\Sigma=F^{-1}\left(\{0\}\right)

and fix some c>0c>0. Consider the function D:+[0,+]D:\mathbb{R}^{+}\rightarrow[0,+\infty] given by

D(λ):={1¯+2λ if λ<¯+2+ if λ¯+2λ+D(\lambda):=\left\{\begin{array}[]{rcl}\frac{1}{\bar{\ell}+2-\lambda}&\mbox{ if }&\lambda<\bar{\ell}+2\\ +\infty&\mbox{ if }&\lambda\geq\bar{\ell}+2\end{array}\right.\qquad\forall\lambda\in\mathbb{R}^{+}

and define the function Ψc:𝒯𝔞¯[0,]\Psi^{c}:\mathcal{T}_{\bar{{\mathfrak{a}}}}\rightarrow[0,\infty] by

Ψc(𝔞):=inf{D(lengthg~(ψ))+max(F(ψ([0,1]))1)|ψΩ(𝔞,Xc¯)}\Psi^{c}({\mathfrak{a}}):=\inf\Bigl{\{}D\left(\mbox{length}^{\tilde{g}}(\psi)\right)+\max\left(F\bigl{(}\psi([0,1])\bigr{)}^{-1}\right)\,|\,\psi\in\Omega\left({\mathfrak{a}},\overline{X^{c}}\right)\Bigr{\}}

where for every 𝔞𝒯𝔞¯{\mathfrak{a}}\in\mathcal{T}_{\bar{{\mathfrak{a}}}}, Ω(𝔞,Xc¯)\Omega\left({\mathfrak{a}},\overline{X^{c}}\right) stands for the set of absolutely continuous curves ψ:[0,1]𝒮0\psi:[0,1]\rightarrow\mathcal{S}_{0} such that ψ(0)=𝔞\psi(0)={\mathfrak{a}}, ψ(1)Xc¯\psi(1)\in\overline{X^{c}} and ψ\psi is almost everywhere tangent to 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}. Let 𝔞𝒜ˇ𝔞¯{\mathfrak{a}}\in\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}} be fixed, by Lemma 4.3 (v), there is an absolutely continuous curve ψ:[0,1]Δ\psi:[0,1]\rightarrow\Delta^{\perp} such that ψ(0)=𝔞\psi(0)={\mathfrak{a}}, ψ(1)Σˇ\psi(1)\in\check{\Sigma}, lengthg~(ψ)¯+1\mbox{length}^{\tilde{g}}(\psi)\leq\bar{\ell}+1 and ψ(t)𝔞𝒮0\psi(t)\in\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0} for all t[0,1)t\in[0,1). Thus, since ψ(t)\psi(t) belongs to VΣV\setminus\Sigma for tt close to 11, Lemma 4.3(iii) shows that 𝔞{\mathfrak{a}} can be joined to XcX^{c} by a curve tangent to 𝔞\mathcal{L}_{{\mathfrak{a}}} contained in 𝒮0\mathcal{S}_{0} of length <¯+2<\bar{\ell}+2. Therefore Ψc(𝔞)\Psi^{c}({\mathfrak{a}}) is finite for every 𝔞𝒜ˇ𝔞¯{\mathfrak{a}}\in\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}. Moreover, the function Ψc\Psi^{c} is lower semi-continuous on 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} (because we consider curves satisfying ψ(1)Xc¯\psi(1)\in\overline{X^{c}} and we may use foliation charts along ψ\psi), so we have

𝒜ˇ𝔞¯=k𝒜ˇk𝔞¯with𝒜ˇk𝔞¯:=(Ψc)1([0,k])𝒜ˇ𝔞¯,\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}=\bigcup_{k\in\mathbb{N}}\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{k}\quad\mbox{with}\quad\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{k}:=\left(\Psi^{c}\right)^{-1}\bigl{(}[0,k]\bigr{)}\cap\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}},

where each set of the above union is a compact subset of 𝒜ˇ𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}. Thus, there is kk\in\mathbb{N} such that the measure of 𝒜ˇk𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{k} with respect to the volume form η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}} is ν/2\geq\nu/2 and such that for every 𝔞𝒜ˇk𝔞¯{\mathfrak{a}}\in\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{k}, there is an absolutely continuous curve ψ:[0,1]𝔞𝒮0\psi:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}}\subset\mathcal{S}_{0} satisfying ψ(0)=𝔞\psi(0)={\mathfrak{a}}, ψ(1)Xc¯\psi(1)\in\overline{X^{c}} and

D(lengthg~(ψ))+max(F(ψ([0,1]))1)k,D\left(\mbox{length}^{\tilde{g}}(\psi)\right)+\max\left(F\bigl{(}\psi([0,1])\bigr{)}^{-1}\right)\leq k,

which implies

lengthg~(ψ)<¯+2andmin(F(ψ([0,1])))1k.\mbox{length}^{\tilde{g}}(\psi)<\bar{\ell}+2\quad\mbox{and}\quad\min\left(F\bigl{(}\psi([0,1])\bigr{)}\right)\geq\frac{1}{k}.

Fix a smooth complete metric g~c\tilde{g}^{c} on 𝒮0\mathcal{S}_{0} which coincides with g~\tilde{g} on the set F1([1/(k),))F^{-1}([1/(k),\infty)). Recall that the definition of YcY^{c} and ZcZ^{c} is given in Lemma 4.3(iv), and note that Xc¯=Yc¯Zc¯\overline{X^{c}}=\overline{Y^{c}}\cup\overline{Z^{c}}, where Zc¯\overline{Z^{c}} has dimension r1\leq r-1. By Lemma 4.3 (iv), the boundary Yc:=Yc¯Yc\partial Y^{c}:=\overline{Y^{c}}\setminus Y^{c} is contained in ZcZ^{c}, so the set of points of 𝒮0\mathcal{S}_{0} that can be joined to Zc¯\overline{Z^{c}} along absolutely continuous curves tangent to 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}} is a countable union of smooth submanifolds of dimension at most r1+m0=2nm1r-1+m_{0}=2n-m-1, so it has measure zero in 𝒮0\mathcal{S}_{0} and in fact since it is invariant by the foliation associated with 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}, its intersection with 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} has measure zero in 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} (by Fubini’s Theorem). Thus, we can consider a compact subset 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c} of 𝒜ˇk𝔞¯𝒜ˇ𝔞¯\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}}_{k}\subset\check{\mathcal{A}}^{\bar{{\mathfrak{a}}}} of measure ν/4\geq\nu/4 such that the properties (i)-(iii) are satisfied. Finally, the set 𝒞c\mathcal{C}^{c} is closed because Yc\partial Y^{c} is closed and (𝒮0,g~c)(\mathcal{S}_{0},\tilde{g}^{c}) is complete. We conclude that 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c} satisfies (iv) by construction. ∎

Given 0<c<ϵ0<c<\epsilon, we define the function Dc:𝒮0𝒞c[0,]D^{c}:\mathcal{S}_{0}\setminus\mathcal{C}^{c}\rightarrow[0,\infty] by

Dc(𝔞):=inf{lengthg~c(ψ)|ψ:[0,1]𝔞 abs. cont. ψ(0)=𝔞,ψ(1)Yc},D^{c}({\mathfrak{a}}):=\inf\Bigl{\{}\mbox{length}^{\tilde{g}^{c}}(\psi)\,|\,\psi:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}}\mbox{ abs. cont. }\,\psi(0)={\mathfrak{a}},\,\psi(1)\in Y^{c}\Bigr{\}},

for every 𝔞𝒮0𝒞c{\mathfrak{a}}\in\mathcal{S}_{0}\setminus\mathcal{C}^{c}, and we denote its domain, the set of points 𝔞𝒮0Cc{\mathfrak{a}}\in\mathcal{S}_{0}\setminus C^{c}, where Dc(𝔞)D^{c}({\mathfrak{a}}) is finite, by dom(Dc)\mbox{dom}(D^{c}). Then, we call 𝔞\mathcal{L}_{{\mathfrak{a}}}-geodesic a curve ψ:[0,1]𝔞\psi:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}} which is geodesic with respect to the metric g~c,𝔞\tilde{g}^{c,{\mathfrak{a}}} induced by g~c\tilde{g}^{c} on 𝔞\mathcal{L}_{{\mathfrak{a}}}, for any point 𝔞Yc{\mathfrak{a}}\in Y^{c} we denote by exp𝔞c:T𝔞𝔞𝔞\exp_{{\mathfrak{a}}}^{c}:T_{{\mathfrak{a}}}\mathcal{L}_{{\mathfrak{a}}}\rightarrow\mathcal{L}_{{\mathfrak{a}}} the exponential map from 𝔞{\mathfrak{a}} with respect to g~c,𝔞\tilde{g}^{c,{\mathfrak{a}}} and by considering 𝒦Yc\vec{\mathcal{K}}_{Y^{c}} as a subbundle of TΔT\Delta^{\perp} (that is, for every 𝔞Yc{\mathfrak{a}}\in Y^{c} we take 𝒦(𝔞)\vec{\mathcal{K}}({\mathfrak{a}})) we define the smooth mapping Expc:𝒦Yc𝒮0\mbox{Exp}^{c}:\vec{\mathcal{K}}_{Y^{c}}\rightarrow\mathcal{S}_{0} by

Expc(𝔞,ζ):=exp𝔞c(ζ)(𝔞,ζ)𝒦Yc.\mbox{Exp}^{c}({\mathfrak{a}},\zeta):=\exp_{{\mathfrak{a}}}^{c}(\zeta)\qquad\forall({\mathfrak{a}},\zeta)\in\vec{\mathcal{K}}_{Y^{c}}.

By completeness of (𝒮0,g~c)(\mathcal{S}_{0},\tilde{g}^{c}), see Lemma 4.4 (i), for every 𝔞dom(Dc){\mathfrak{a}}\in\mbox{dom}(D^{c}) any sequence {ψk:[0,1]𝔞}k\{\psi_{k}:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}}\}_{k\in\mathbb{N}} of absolutely continuous curves such that

ψk(0)=𝔞,ψk(1)Ycandlimklengthgc(ψk)=Dc(𝔞),\psi_{k}(0)={\mathfrak{a}},\quad\psi_{k}(1)\in Y^{c}\quad\mbox{and}\quad\lim_{k\rightarrow\infty}\mbox{length}^{g^{c}}(\psi_{k})=D^{c}({\mathfrak{a}}),

converges, up to taking a subsequence, to a 𝔞\mathcal{L}_{{\mathfrak{a}}}-geodesic ψ¯\bar{\psi}, called minimizing geodesic for Dc(𝔞)D^{c}({\mathfrak{a}}), satisfying ψ¯(0)=𝔞\bar{\psi}(0)={\mathfrak{a}}, ψ¯(1)Yc¯\bar{\psi}(1)\in\overline{Y^{c}} and lengthgc(ψk)=Dc(𝔞)\mbox{length}^{g^{c}}(\psi_{k})=D^{c}({\mathfrak{a}}). Moreover if in addition Dc(𝔞)<l¯+2D^{c}({\mathfrak{a}})<\bar{l}+2 then we have ψ¯(1)Yc\bar{\psi}(1)\in Y^{c} because 𝔞dom(Dc)𝒮0𝒞c{\mathfrak{a}}\in\mbox{dom}(D^{c})\subset\mathcal{S}_{0}\setminus\mathcal{C}^{c}. For every 𝔞dom(Dc){\mathfrak{a}}\in\mbox{dom}(D^{c}), we set

(4.7) Γc(𝔞)\displaystyle\Gamma^{c}({\mathfrak{a}}) the set of all minimizing geodesics forDc(𝔞),\displaystyle\quad\text{the set of all minimizing geodesics for}\,D^{c}({\mathfrak{a}}),
c(𝔞)\displaystyle\mathcal{I}^{c}({\mathfrak{a}}) :={ψ(t)|ψΓc(𝔞),t[0,1]}.\displaystyle=\Bigl{\{}\psi(t)\,|\,\psi\in\Gamma^{c}({\mathfrak{a}}),\,t\in[0,1]\Bigr{\}}.

By completeness of (𝒮0,g~c)(\mathcal{S}_{0},\tilde{g}^{c}) and regularity of the foliation given by 𝒦|𝒮0\vec{\mathcal{K}}_{|\mathcal{S}_{0}}, the mapping 𝔞dom(Dc)c(𝔞){\mathfrak{a}}\in\mbox{dom}(D^{c})\mapsto\mathcal{I}^{c}({\mathfrak{a}}) has closed graph. Moreover, by the above construction and properties (iii)-(iv) of Lemma 4.4, the set

c(𝒜ˇ𝔞¯,c):=𝔞𝒜ˇ𝔞¯,cc(𝔞)\mathcal{I}^{c}\left(\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}\right):=\bigcup_{{\mathfrak{a}}\in\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}}\mathcal{I}^{c}({\mathfrak{a}})

is a compact subset of 𝒮0\mathcal{S}_{0} which is contained in dom(Dc)\mbox{dom}(D^{c}). The following lemma follows from classical results on distance functions from submanifolds in Riemannian geometry (see Figure 4).

Lemma 4.5.

For every 0<c<ϵ0<c<\epsilon, there are a relatively compact open subset 𝒱c,𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}} of 𝒯𝔞¯\mathcal{T}_{\bar{{\mathfrak{a}}}} containing 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}, an open neighborhood c\mathcal{H}^{c} of 𝔞¯\bar{{\mathfrak{a}}} in 𝔞¯\mathcal{L}_{\bar{{\mathfrak{a}}}}, an open set 𝒰c𝒲\mathcal{U}^{c}\subset\mathcal{W} and a set Fc𝒰cF^{c}\subset\mathcal{U}^{c} satisfying the following properties:

  • (i)

    The set Fc𝒰cF^{c}\subset\mathcal{U}^{c} is closed with respect to the induced topology on 𝒰c\mathcal{U}^{c}.

  • (ii)

    The set FcF^{c} has Lebesgue measure zero in 𝒰c\mathcal{U}^{c}.

  • (iii)

    The function DcD^{c} is smooth on the open set (recall the notation introduced for local transition maps (4.2))

    𝒰cFcwith𝒰c:=𝔞cT𝔞¯,𝔞(𝒱c,𝔞¯)=𝔞c𝒱𝔞c,𝔞¯𝒲\mathcal{U}^{c}\setminus F^{c}\quad\mbox{with}\quad\mathcal{U}^{c}:=\bigcup_{{\mathfrak{a}}\in\mathcal{H}^{c}}T^{\bar{{\mathfrak{a}}},{\mathfrak{a}}}(\mathcal{V}^{c,\bar{{\mathfrak{a}}}})=\bigcup_{{\mathfrak{a}}\in\mathcal{H}^{c}}\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}}\subset\mathcal{W}

    and for every 𝔞𝒰cFc{\mathfrak{a}}\in\mathcal{U}^{c}\setminus F^{c} the set Γc(𝔞)\Gamma^{c}({\mathfrak{a}}) given in (4.7) is a singleton {ψc,𝔞}\{\psi^{c,{\mathfrak{a}}}\}, where ψc,𝔞:[0,1]𝔞\psi^{c,{\mathfrak{a}}}:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}} is the 𝔞\mathcal{L}_{{\mathfrak{a}}}-geodesic (uniquely) defined by the initial conditions

    ψc,𝔞(0)=𝔞andψ˙c,𝔞(0)=D𝔞c(𝔞)\psi^{c,{\mathfrak{a}}}(0)={\mathfrak{a}}\quad\mbox{and}\quad\dot{\psi}^{c,{\mathfrak{a}}}(0)=-\nabla D^{c}_{{\mathfrak{a}}}({\mathfrak{a}})

    (D𝔞c\nabla D^{c}_{{\mathfrak{a}}} stands for the gradient of D𝔞cD^{c}_{{\mathfrak{a}}} with respect to gc,𝔞g^{c,{\mathfrak{a}}}).

  • (iv)

    For every 𝔞c{\mathfrak{a}}\in\mathcal{H}^{c}, the mapping

    Hc:((𝒰cFc)𝒯𝔞)×[0,1]𝒮0(𝔞,t)ψc,𝔞(t)\begin{array}[]{rcl}H^{c}\,:\,\left(\left(\mathcal{U}^{c}\setminus F^{c}\right)\cap\mathcal{T}_{{\mathfrak{a}}}\right)\times[0,1]&\longrightarrow&\,\mathcal{S}_{0}\\ ({\mathfrak{a}}^{\prime},t)&\longmapsto&\,\psi^{c,{\mathfrak{a}}^{\prime}}(t)\end{array}

    is a smooth diffeomorphism onto its image.

Refer to caption
Figure 4. A picture to illustrate Lemma 4.5
Proof of Lemma 4.5.

The set c(𝒜ˇ𝔞¯,c)Yc¯\mathcal{I}^{c}\left(\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}\right)\cap\overline{Y^{c}} is a compact set which does not intersect Yc\partial Y^{c}, so there is an open set 𝒪𝒮0\mathcal{O}\subset\mathcal{S}_{0} which contains Yc\partial Y^{c} and such that c(𝒜ˇ𝔞¯,c)YcYc𝒪¯\mathcal{I}^{c}\left(\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}\right)\cap Y^{c}\subset Y^{c}\setminus\overline{\mathcal{O}}. As a consequence, by regularity of the mapping 𝔞dom(Dc)c(𝔞){\mathfrak{a}}\in\mbox{dom}(D^{c})\mapsto\mathcal{I}^{c}({\mathfrak{a}}), there is an open set 𝒰c𝒲\mathcal{U}^{c}\subset\mathcal{W} containing 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c} such that

Dc(𝔞)<l¯+2andc(𝔞)Yc¯Yc𝒪¯𝔞𝒰c.D^{c}({\mathfrak{a}})<\bar{l}+2\quad\mbox{and}\quad\mathcal{I}^{c}({\mathfrak{a}})\cap\overline{Y^{c}}\subset Y^{c}\setminus\overline{\mathcal{O}}\qquad\forall{\mathfrak{a}}\in\mathcal{U}^{c}.

In fact, for every 𝔞𝒰c{\mathfrak{a}}\in\mathcal{U}^{c}, the restriction of DcD^{c} to the local leaf 𝔞𝒰c\mathcal{L}_{{\mathfrak{a}}}\cap\mathcal{U}^{c}, let us denote it by D𝔞cD^{c}_{{\mathfrak{a}}}, coincides with the distance function to the set Y~c,𝔞:=𝔞(Yc𝒪¯)\widetilde{Y}^{c,{\mathfrak{a}}}:=\mathcal{L}_{{\mathfrak{a}}}\cap(Y^{c}\setminus\overline{\mathcal{O}}) which, by the transversality property given by Lemma 4.3 (iv) and compactness of Yc𝒪YcY^{c}\setminus\mathcal{O}\subset Y^{c}, is the union of finitely many points. So, as a distance function from a smooth submanifold (of dimension zero) on a complete Riemannian manifold, for every 𝔞𝒰c{\mathfrak{a}}\in\mathcal{U}^{c} the function D𝔞cD^{c}_{{\mathfrak{a}}} satisfies the following properties:

  • (P1)

    The function D𝔞cD^{c}_{{\mathfrak{a}}} is locally lipschitz on ^𝔞:=𝔞𝒰c\hat{\mathcal{L}}_{{\mathfrak{a}}}:=\mathcal{L}_{{\mathfrak{a}}}\cap\mathcal{U}^{c} and its singular set Σ(D𝔞c)\Sigma(D^{c}_{{\mathfrak{a}}}), defined as the set of points in ^𝔞\hat{\mathcal{L}}_{{\mathfrak{a}}} where D𝔞cD^{c}_{{\mathfrak{a}}} is not differentiable, has measure zero in ^𝔞\hat{\mathcal{L}}_{{\mathfrak{a}}}.

  • (P2)

    Denoting by D𝔞c\nabla D^{c}_{{\mathfrak{a}}} the gradient of D𝔞cD^{c}_{{\mathfrak{a}}} with respect to gc,𝔞g^{c,{\mathfrak{a}}}, define the limiting-gradient of D𝔞cD^{c}_{{\mathfrak{a}}} at some point 𝔞^𝔞{\mathfrak{a}}^{\prime}\in\hat{\mathcal{L}}_{{\mathfrak{a}}}, denoted by LD𝔞c(𝔞)T𝔞𝔞\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}^{\prime})\subset T_{{\mathfrak{a}}^{\prime}}\mathcal{L}_{{\mathfrak{a}}}, as the set of all limits in T𝔞𝔞T_{{\mathfrak{a}}^{\prime}}\mathcal{L}_{{\mathfrak{a}}} of sequences of the form {D𝔞c(𝔞k)}kT𝔞k𝔞\{\nabla D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}_{k})\}_{k\in\mathbb{N}}\in T_{{\mathfrak{a}}_{k}}\mathcal{L}_{{\mathfrak{a}}} where {ak}k\{a_{k}\}_{k\in\mathbb{N}} is a sequence of points in ^𝔞Σ(D𝔞c)\hat{\mathcal{L}}_{{\mathfrak{a}}}\setminus\Sigma(D^{c}_{{\mathfrak{a}}}) converging to 𝔞{\mathfrak{a}}^{\prime} (note that by (P1) such sequences do exist). Then a point 𝔞^𝔞{\mathfrak{a}}^{\prime}\in\hat{\mathcal{L}}_{{\mathfrak{a}}} belongs to Σ(D𝔞c)\Sigma(D^{c}_{{\mathfrak{a}}}) if and only if LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}^{\prime}) is not a singleton. Moreover, for every 𝔞^𝔞{\mathfrak{a}}^{\prime}\in\hat{\mathcal{L}}_{{\mathfrak{a}}}, there is a one-to-one correspondence between LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}^{\prime}) and Γc(𝔞)\Gamma^{c}({\mathfrak{a}}^{\prime}) (the set of all minimizing geodesics for Dc(𝔞)D^{c}({\mathfrak{a}}^{\prime})), namely a vector ζT𝔞^𝔞\zeta^{\prime}\in T_{{\mathfrak{a}}^{\prime}}\hat{\mathcal{L}}_{{\mathfrak{a}}^{\prime}} belongs to LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}^{\prime}) if and only if the 𝔞\mathcal{L}_{{\mathfrak{a}}}-geodesic ψ𝔞,ζ:[0,1]𝔞\psi^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}:[0,1]\rightarrow\mathcal{L}_{{\mathfrak{a}}} (uniquely) defined by the initial conditions

    ψ𝔞,ζ(0)=𝔞andψ˙𝔞,ζ(0)=ζ\psi^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(0)={\mathfrak{a}}^{\prime}\quad\mbox{and}\quad\dot{\psi}^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(0)=-\zeta^{\prime}

    is a minimizing geodesic for Dc(𝔞)D^{c}({\mathfrak{a}}^{\prime}). Moreover, every such geodesic satisfies

    ψ𝔞,ζ(1)Zc,𝔞andψ𝔞,ζ(t)=β𝔞,ζ(1t)t[0,1]\psi^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(1)\in Z^{c,{\mathfrak{a}}}\quad\mbox{and}\quad\psi^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(t)=\beta^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(1-t)\quad\forall t\in[0,1]

    where β𝔞,ζ\beta^{{\mathfrak{a}}^{\prime},\zeta^{\prime}} is the 𝔞\mathcal{L}_{{\mathfrak{a}}}-geodesic given by

    β(t):=expP(𝔞,ζ)c(V(𝔞,ζ))t[0,1]\beta(t):=\exp_{P({\mathfrak{a}}^{\prime},\zeta^{\prime})}^{c}\left(V({\mathfrak{a}}^{\prime},\zeta^{\prime})\right)\quad\forall t\in[0,1]

    with

    P(𝔞,ζ):=ψ𝔞,ζ(1)Zc,𝔞andV(𝔞,ζ):=ψ˙𝔞,ζ(1)TP(𝔞,ζ)𝔞.P({\mathfrak{a}}^{\prime},\zeta^{\prime}):=\psi^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(1)\in Z^{c,{\mathfrak{a}}}\quad\mbox{and}\quad V({\mathfrak{a}}^{\prime},\zeta^{\prime}):=-\dot{\psi}^{{\mathfrak{a}}^{\prime},\zeta^{\prime}}(1)\in T_{P({\mathfrak{a}}^{\prime},\zeta^{\prime})}\mathcal{L}_{{\mathfrak{a}}}.
  • (P3)

    Let Conjc,𝔞(𝔞^)𝔞^\mbox{Conj}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}})\subset\hat{\mathcal{L}_{{\mathfrak{a}}}} be the set of points 𝔞𝔞^{\mathfrak{a}}^{\prime}\in\hat{\mathcal{L}_{{\mathfrak{a}}}} for which there is ζLD𝔞c(𝔞)\zeta^{\prime}\in\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}^{\prime}), called conjugate limiting-gradient of D𝔞cD^{c}_{{\mathfrak{a}}} at 𝔞{\mathfrak{a}}^{\prime}, such that the tangent vector V(𝔞,ζ)V({\mathfrak{a}}^{\prime},\zeta^{\prime}) is a critical point of the exponential map expP(𝔞,ζ)c\exp_{P({\mathfrak{a}}^{\prime},\zeta^{\prime})}^{c}. Then we have

    Cutc,𝔞(𝔞^):=Σ(D𝔞c)¯=Σ(D𝔞c)Conjc,𝔞(𝔞^).\mbox{Cut}^{c,{\mathfrak{a}}}\left(\hat{\mathcal{L}_{{\mathfrak{a}}}}\right):=\overline{\Sigma(D^{c}_{{\mathfrak{a}}})}=\Sigma(D^{c}_{{\mathfrak{a}}})\cup\mbox{Conj}^{c,{\mathfrak{a}}}\left(\hat{\mathcal{L}_{{\mathfrak{a}}}}\right).
  • (P4)

    The set Cutc,𝔞(𝔞^)\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}), called cut locus in 𝔞^\hat{\mathcal{L}_{{\mathfrak{a}}}}, has Lebesgue measure zero in ^𝔞\hat{\mathcal{L}}_{{\mathfrak{a}}} and the function D𝔞cD^{c}_{{\mathfrak{a}}} is smooth on ^𝔞Cutc,𝔞(𝔞^)\hat{\mathcal{L}}_{{\mathfrak{a}}}\setminus\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}). In particular, for every 𝔞^𝔞Cutc,𝔞(𝔞^){\mathfrak{a}}^{\prime}\in\hat{\mathcal{L}}_{{\mathfrak{a}}}\setminus\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}), the set LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}^{\prime}) is a singleton {ζ(𝔞)}\{\zeta({\mathfrak{a}}^{\prime})\} and moreover the exponential map expP(𝔞,ζ)c:TP(𝔞,ζ)𝔞𝔞\exp_{P({\mathfrak{a}}^{\prime},\zeta^{\prime})}^{c}:T_{P({\mathfrak{a}}^{\prime},\zeta^{\prime})}\mathcal{L}_{{\mathfrak{a}}}\rightarrow\mathcal{L}_{{\mathfrak{a}}} is a submersion at V(𝔞,ζ)V({\mathfrak{a}}^{\prime},\zeta^{\prime}).

The property (P1) follows from Rademacher’s Theorem, (P2) may be found in [18, Lemma 11] (where the result is stated with Hamiltonian viewpoint) and (P3)-(P4) may be found in [7, 15, 20].

To conclude the proof of the lemma, we define the set Fc𝒰cF^{c}\subset\mathcal{U}^{c} by

Fc:=𝔞𝒰cCutc,𝔞(𝔞^).F^{c}:=\bigcup_{{\mathfrak{a}}\in\mathcal{U}^{c}}\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}).

Let us prove (i), that is, FcF^{c} is closed in the topological subspace 𝒰c\mathcal{U}^{c}. Let {𝔞k}k\{{\mathfrak{a}}_{k}\}_{k\in\mathbb{N}} be a sequence of points of FcF^{c} converging to some 𝔞𝒰c{\mathfrak{a}}\in\mathcal{U}^{c}. Let us distinguish two cases:

Case 1: There is a constant δ>0\delta>0 such that the diameters (with respect to g~c\tilde{g}^{c}) of the sets LD𝔞kc(𝔞k)\nabla^{L}D^{c}_{{\mathfrak{a}}_{k}}({\mathfrak{a}}_{k}) are all larger than δ\delta (so that for all kk\in\mathbb{N}, 𝔞k{\mathfrak{a}}_{k} belongs to Σ(D𝔞kc)\Sigma(D^{c}_{{\mathfrak{a}}_{k}})).
Then 𝔞{\mathfrak{a}} admits two minimizing geodesics ψ1,ψ2\psi_{1},\psi_{2} for Dc(𝔞)D^{c}({\mathfrak{a}}) such that

|ψ˙1(0)ψ˙1(0)|g~cδ>0\left|\dot{\psi}_{1}(0)-\dot{\psi}_{1}(0)\right|^{\tilde{g}^{c}}\geq\delta>0

so LD𝔞kc(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}_{k}}({\mathfrak{a}}) is not a singleton (by (P2)) and 𝔞{\mathfrak{a}} belongs to Σ(D𝔞c)Fc\Sigma(D^{c}_{{\mathfrak{a}}})\subset F^{c}.

Case 2: There is not a constant δ>0\delta>0 such that the diameters (with respect to g~c\tilde{g}^{c}) of the sets LD𝔞kc(𝔞k)\nabla^{L}D^{c}_{{\mathfrak{a}}_{k}}({\mathfrak{a}}_{k}) are all larger than δ\delta (so that for all kk\in\mathbb{N}, 𝔞k{\mathfrak{a}}_{k} belongs to Σ(D𝔞kc)\Sigma(D^{c}_{{\mathfrak{a}}_{k}})).
Then we have

lim infkdiamg~cLD𝔞kc(𝔞k)=0.\liminf_{k\rightarrow\infty}\mbox{diam}^{\tilde{g}^{c}}\nabla^{L}D^{c}_{{\mathfrak{a}}_{k}}({\mathfrak{a}}_{k})=0.

Let us again distinguish between two cases.

Subcase 2.1: There are infinitely many kk\in\mathbb{N} for which 𝔞k{\mathfrak{a}}_{k} belongs to Conjc,𝔞k(^𝔞k)\mbox{Conj}^{c,{\mathfrak{a}}_{k}}(\hat{\mathcal{L}}_{{\mathfrak{a}}_{k}}).
Then, by considering a subsequence of {V(𝔞k,ζk)}k\{V({\mathfrak{a}}_{k},\zeta_{k})\}_{k\in\mathbb{N}} with ζk\zeta_{k} a conjugate limiting-gradient of D𝔞kcD^{c}_{{\mathfrak{a}}_{k}} at 𝔞k{\mathfrak{a}}_{k}, there is a tangent vector V(𝔞,ζ)V({\mathfrak{a}},\zeta) which is a critical point of the exponential map expP(𝔞,ζ)c\exp_{P({\mathfrak{a}},\zeta)}^{c} as limit of the sequence of critical vectors {V(𝔞k,ζk)}k\{V({\mathfrak{a}}_{k},\zeta_{k})\}_{k\in\mathbb{N}} (with respect to expP(𝔞k,ζk)c\exp_{P({\mathfrak{a}}_{k},\zeta_{k})}^{c}). Therefore, 𝔞{\mathfrak{a}} belongs to Conjc,𝔞(^𝔞)Fc\mbox{Conj}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}}_{{\mathfrak{a}}})\subset F^{c} by (P3).

Subcase 2.2: The set of kk\in\mathbb{N} for which 𝔞k{\mathfrak{a}}_{k} belongs to Conjc,𝔞k(^𝔞k)\mbox{Conj}^{c,{\mathfrak{a}}_{k}}(\hat{\mathcal{L}}_{{\mathfrak{a}}_{k}}) is finite.
If 𝔞Cutc,𝔞(𝔞^){\mathfrak{a}}\notin\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}), then by (P2) the limiting-gradient LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}) is equal to a singleton {ζ}\{\zeta\} and there is only one minimizing geodesics for Dc(𝔞)D^{c}({\mathfrak{a}}) given by ψ𝔞,ζ\psi^{{\mathfrak{a}},\zeta}. Thus, by (P3), up to considering a subsequence, we may assume without loss of generality that for all kk\in\mathbb{N} there are ζk1,ζk2\zeta_{k}^{1},\zeta_{k}^{2} in LD𝔞kc(𝔞k)\nabla^{L}D^{c}_{{\mathfrak{a}}_{k}}({\mathfrak{a}}_{k}) such that

ζk1ζk2andlimk|ζk1ζk2|g~c=0\zeta_{k}^{1}\neq\zeta_{k}^{2}\quad\mbox{and}\quad\lim_{k\rightarrow\infty}\left|\zeta_{k}^{1}-\zeta_{k}^{2}\right|^{\tilde{g}^{c}}=0

and for i=1,2i=1,2

limkζik=ζ,limkP(𝔞k,ζik)=P(𝔞,ζ),limkV(𝔞k,ζik)=V(𝔞,ζ).\lim_{k\rightarrow\infty}\zeta_{i}^{k}=\zeta,\quad\lim_{k\rightarrow\infty}P({\mathfrak{a}}_{k},\zeta_{i}^{k})=P({\mathfrak{a}},\zeta),\quad\lim_{k\rightarrow\infty}V({\mathfrak{a}}_{k},\zeta_{i}^{k})=V({\mathfrak{a}},\zeta).

Since 𝔞Cutc,𝔞(𝔞^){\mathfrak{a}}\notin\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}), (P4) shows that the exponential map expP(𝔞,ζ)c:TP(𝔞,ζ)𝔞𝔞\exp_{P({\mathfrak{a}},\zeta)}^{c}:T_{P({\mathfrak{a}},\zeta)}\mathcal{L}_{{\mathfrak{a}}}\rightarrow\mathcal{L}_{{\mathfrak{a}}} is a submersion at V(𝔞,ζ)V({\mathfrak{a}},\zeta). So the mappings expP(𝔞k,ζk)c:TP(𝔞k,ζk)𝔞k𝔞k\exp_{P({\mathfrak{a}}_{k},\zeta_{k})}^{c}:T_{P({\mathfrak{a}}_{k},\zeta_{k})}\mathcal{L}_{{\mathfrak{a}}_{k}}\rightarrow\mathcal{L}_{{\mathfrak{a}}_{k}} are submersions at V(𝔞k,ζk)V({\mathfrak{a}}_{k},\zeta_{k}) for kk large enough but this is impossible because

expP(𝔞k,ζk)c(V(𝔞k,ζ1k))=expP(𝔞k,ζk)c(V(𝔞k,ζ2k))k.\exp_{P({\mathfrak{a}}_{k},\zeta_{k})}^{c}\left(V({\mathfrak{a}}_{k},\zeta_{1}^{k})\right)=\exp_{P({\mathfrak{a}}_{k},\zeta_{k})}^{c}\left(V({\mathfrak{a}}_{k},\zeta_{2}^{k})\right)\qquad\forall k\in\mathbb{N}.

So we have 𝔞Cutc,𝔞(𝔞^){\mathfrak{a}}\in\mbox{Cut}^{c,{\mathfrak{a}}}(\hat{\mathcal{L}_{{\mathfrak{a}}}}).

To prove (ii), we just notice that 𝒰c\mathcal{U}^{c} is foliated by the leaves ^𝔞\hat{\mathcal{L}}_{{\mathfrak{a}}} whose intersection with FcF^{c} has measure zero by (P4). So, we get the result by a Fubini argument.

The point (iii) is a consequence of the fact that LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}) is a singleton {ζ𝔞}\{\zeta^{{\mathfrak{a}}}\} for all 𝔞{\mathfrak{a}} in the open set 𝒰cFc\mathcal{U}^{c}\setminus F^{c} together with the fact that expP(𝔞,ζ𝔞)c:TP(𝔞,ζ𝔞)𝔞𝔞\exp_{P({\mathfrak{a}},\zeta^{{\mathfrak{a}}})}^{c}:T_{P({\mathfrak{a}},\zeta^{{\mathfrak{a}}})}\mathcal{L}_{{\mathfrak{a}}}\rightarrow\mathcal{L}_{{\mathfrak{a}}} is a submersion at V(𝔞,ζ𝔞)V({\mathfrak{a}},\zeta^{{\mathfrak{a}}}) which implies that the mapping Expc\mbox{Exp}^{c} is a submersion at (P(𝔞,ζ𝔞),V(𝔞,ζ𝔞))(P({\mathfrak{a}},\zeta^{{\mathfrak{a}}}),V({\mathfrak{a}},\zeta^{{\mathfrak{a}}})). As a matter of fact, if 𝔞𝒰cFc{\mathfrak{a}}\in\mathcal{U}^{c}\setminus F^{c} is fixed, then there is an open neighborhood NN of (P(𝔞,ζ𝔞),V(𝔞,ζ𝔞))(P({\mathfrak{a}},\zeta^{{\mathfrak{a}}}),V({\mathfrak{a}},\zeta^{{\mathfrak{a}}})) in 𝒦YcTΔ\vec{\mathcal{K}}_{Y^{c}}\subset T\Delta^{\perp} such that the image Expc(N)\mbox{Exp}^{c}(N) is an open neighborhood of 𝔞{\mathfrak{a}} and we have necessarily for every (P,V)N(P,V)\in N,

LDAc(A)={ζA}={ψ˙(P,V)(0)}withA=A(P,V):=Expc(P,V),\nabla^{L}D^{c}_{A}(A)=\left\{\zeta^{A}\right\}=\left\{-\dot{\psi}^{(P,V)}(0)\right\}\quad\mbox{with}\quad A=A(P,V):=\mbox{Exp}^{c}(P,V),

where ψ(P,V):[0,1](P,V)\psi^{(P,V)}:[0,1]\rightarrow\mathcal{L}_{(P,V)} is the (P,V)\mathcal{L}_{(P,V)}-geodesic given by

ψ(P,V)(t):=Expc(P,(1t)V)=exp(P,V)c((1t)V)t[0,1],\psi^{(P,V)}(t):=\mbox{Exp}^{c}(P,(1-t)V)=\exp_{(P,V)}^{c}((1-t)V)\qquad\forall t\in[0,1],

because ψ(P,V)\psi^{(P,V)} is the only (P,V)\mathcal{L}_{(P,V)}-geodesic closed to ψ𝔞,ζ𝔞\psi^{{\mathfrak{a}},\zeta^{{\mathfrak{a}}}} joining AA to Zc,AZ^{c,A}. Since the mapping

Addt{ψ(Expc)1(A)}(0)A\,\longmapsto\,-\frac{d}{dt}\left\{\psi^{\bigl{(}\mbox{Exp}^{c}\bigr{)}^{-1}(A)}\right\}(0)

is smooth we infer that DcD^{c} is smooth on 𝒰cFc\mathcal{U}^{c}\setminus F^{c} and that Γc(𝔞)\Gamma^{c}({\mathfrak{a}}) is a singleton for 𝔞𝒰cFc{\mathfrak{a}}\in\mathcal{U}^{c}\setminus F^{c} because LD𝔞c(𝔞)\nabla^{L}D^{c}_{{\mathfrak{a}}}({\mathfrak{a}}) is always a singleton (see (P1)).

To prove (iv), we first notice that, up to shrinking 𝒰c\mathcal{U}^{c}, we may assume that for every 𝔞c{\mathfrak{a}}\in\mathcal{H}^{c}, the mapping HcH^{c} is injective. As a matter of fact, suppose for contradiction that there are 𝔞c{\mathfrak{a}}\in\mathcal{H}^{c} and

(𝔞1,t1),(𝔞2,t2)((𝒰cFc)𝒯𝔞)×[0,1]({\mathfrak{a}}_{1},t_{1}),({\mathfrak{a}}_{2},t_{2})\in\left(\left(\mathcal{U}^{c}\setminus F^{c}\right)\cap\mathcal{T}_{{\mathfrak{a}}}\right)\times[0,1]

such that

Hc(𝔞1,t1)=ψc,𝔞1(t1)=ψc,𝔞2(t2)=Hc(𝔞2,t2).H^{c}({\mathfrak{a}}_{1},t_{1})=\psi^{c,{\mathfrak{a}}_{1}}(t_{1})=\psi^{c,{\mathfrak{a}}2}(t_{2})=H^{c}({\mathfrak{a}}_{2},t_{2}).

Since ψc,𝔞1\psi^{c,{\mathfrak{a}}_{1}} and ψc,𝔞2\psi^{c,{\mathfrak{a}}_{2}} are minimizing the length (among curves with are horizontal with respect to the foliation) we have either 𝔞1=𝔞2{\mathfrak{a}}_{1}={\mathfrak{a}}_{2} and ψc,𝔞1=ψc,𝔞2\psi^{c,{\mathfrak{a}}_{1}}=\psi^{c,{\mathfrak{a}}2} (because Γc(𝔞1)=Γc(𝔞2)\Gamma^{c}({\mathfrak{a}}_{1})=\Gamma^{c}({\mathfrak{a}}_{2}) is a singleton), or we have 𝔞1𝔞2{\mathfrak{a}}_{1}\neq{\mathfrak{a}}_{2} and t1=t2=1t_{1}=t_{2}=1. In the latter case, we infer that 𝔞1{\mathfrak{a}}_{1} and 𝔞2{\mathfrak{a}}_{2} belong to the same leaf 𝔞1=𝔞2\mathcal{L}_{{\mathfrak{a}}_{1}}=\mathcal{L}_{{\mathfrak{a}}_{2}} and can be connect by a curve horizontal (with respect to 𝔞1\mathcal{L}_{{\mathfrak{a}}_{1}}) of length <2¯+4<2\bar{\ell}+4. By Lemma 4.2 (iii), this cannot occur if the open neighborhood 𝒰c𝒲\mathcal{U}^{c}\subset\mathcal{W} of 𝒜ˇ𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c} is sufficiently small. The smoothness of HcH^{c} follows from (iii) and the property of diffeomorphism is a consequence of the fact that all minimizing curves from 𝒰c\mathcal{U}^{c} to Y¯c𝒪\bar{Y}^{c}\setminus\mathcal{O} have no conjugate times.

We conclude easily the construction of 𝒱c,𝔞¯𝒯𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}}\subset\mathcal{T}_{\bar{{\mathfrak{a}}}} and c\mathcal{H}^{c} of 𝔞¯𝔞¯\bar{{\mathfrak{a}}}\subset\mathcal{L}_{\bar{{\mathfrak{a}}}}. ∎

The following lemma will allow us to conclude the proof of Theorem 1.1, it follows easily from Lemma 4.5.

Lemma 4.6.

For every 0<c<ϵ0<c<\epsilon, there are 𝔞c𝔞¯𝒲{\mathfrak{a}}^{c}\in\mathcal{L}_{\bar{{\mathfrak{a}}}}\cap\mathcal{W}, a finite set JcJ^{c}, two collections of sets {𝒪jc,0}jJc\{\mathcal{O}_{j}^{c,0}\}_{j\in J^{c}}, {𝒪jc,1}jJc\{\mathcal{O}_{j}^{c,1}\}_{j\in J^{c}} and a collection of functions {Φjc:𝒪jc,0×[0,1]𝒮0}jJc\{\Phi_{j}^{c}:\mathcal{O}_{j}^{c,0}\times[0,1]\rightarrow\mathcal{S}_{0}\}_{j\in J^{c}} satisfying the following properties:

  • (i)

    The sets 𝒪jc,0\mathcal{O}_{j}^{c,0} (with jJcj\in J^{c}) are pairwise disjoint.

  • (ii)

    The sets 𝒪jc,1\mathcal{O}_{j}^{c,1} (with jJcj\in J^{c}) are pairwise disjoint.

  • (iii)

    For every jJcj\in J^{c}, 𝒪jc,0\mathcal{O}_{j}^{c,0} is a compact, connected and oriented, smooth submanifold with boundary of 𝒯𝔞c\mathcal{T}_{{\mathfrak{a}}^{c}} of dimension rr.

  • (iv)

    For every jJcj\in J^{c}, 𝒪jc,1\mathcal{O}_{j}^{c,1} is a compact, connected and oriented, smooth submanifold with boundary of YcXcY^{c}\subset X^{c} of dimension rr.

  • (v)

    For every jJcj\in J^{c}, Φjc:𝒪jc,0×[0,1]𝒮0\Phi_{j}^{c}:\mathcal{O}_{j}^{c,0}\times[0,1]\rightarrow\mathcal{S}_{0} is smooth and for every t[0,1]t\in[0,1], the restriction of Φjc\Phi_{j}^{c} to 𝒪jc,0×{t}\mathcal{O}_{j}^{c,0}\times\{t\} is a diffeomorphism from 𝒪jc,0×{t}\mathcal{O}_{j}^{c,0}\times\{t\} to its image

    𝒪jc,t:=Φjc(𝒪jc,0×{t}).\mathcal{O}_{j}^{c,t}:=\Phi_{j}^{c}\left(\mathcal{O}_{j}^{c,0}\times\{t\}\right).

    Moreover, Φjc(𝔞,0)=𝔞\Phi_{j}^{c}({\mathfrak{a}},0)={\mathfrak{a}} for every 𝔞𝒪jc,0{\mathfrak{a}}\in\mathcal{O}_{j}^{c,0} and 𝒪jc,1\mathcal{O}_{j}^{c,1} is the diffeomorphic image of 𝒪jc,0×{1}\mathcal{O}_{j}^{c,0}\times\{1\} by Φjc\Phi_{j}^{c}.

  • (vi)

    For every jJcj\in J^{c} and any t,t[0,1]t,t^{\prime}\in[0,1] with ttt\neq t^{\prime}, 𝒪jc,t𝒪jc,t=\mathcal{O}_{j}^{c,t}\cap\mathcal{O}_{j}^{c,t^{\prime}}=\emptyset.

  • (vii)

    For every jJcj\in J^{c} and every 𝔞𝒪ic,0{\mathfrak{a}}\in\mathcal{O}_{i}^{c,0}, the smooth curve t[0,1]Φic(𝔞,t)t\in[0,1]\rightarrow\Phi_{i}^{c}({\mathfrak{a}},t) is a 𝔞\mathcal{L}_{{\mathfrak{a}}}-geodesic with non zero speed.

  • (viii)

    The set 𝒪c,0:=jJc𝒪jc,0\mathcal{O}^{c,0}:=\cup_{j\in J^{c}}\mathcal{O}_{j}^{c,0} has measure ν/16\geq\nu/16 with respect to the volume form η|𝒯𝔞¯\eta_{|\mathcal{T}_{\bar{{\mathfrak{a}}}}}.

Proof of Lemma 4.6.

Fix 0<c<ϵ0<c<\epsilon and consider the sets 𝒱c,𝔞¯𝒯𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}}\subset\mathcal{T}_{\bar{{\mathfrak{a}}}}, c𝔞¯\mathcal{H}^{c}\subset\mathcal{L}_{\bar{{\mathfrak{a}}}}, Fc𝒮0F^{c}\subset\mathcal{S}_{0} and 𝒰c𝒲\mathcal{U}^{c}\subset\mathcal{W} given by Lemma 4.5. The set 𝒰c\mathcal{U}^{c} is foliated by the leaves 𝒱𝔞c,𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}} with 𝔞c{\mathfrak{a}}\in\mathcal{H}^{c} and by Lemma 4.5 (ii), the set Fc𝒰cF^{c}\cap\mathcal{U}^{c} has Lebesgue measure zero. Hence Fubini’s Theorem implies that there is 𝔞cc{\mathfrak{a}}^{c}\in\mathcal{H}^{c} such that the set Fc𝒱𝔞cc,𝔞¯𝒯𝔞cF^{c}\cap\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}\subset\mathcal{T}_{{\mathfrak{a}}^{c}} has measure zero. Without loss of generality, up to shrinking 𝒱𝔞cc,𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}} in 𝒯𝔞c\mathcal{T}_{{\mathfrak{a}}^{c}} we may assume that the relatively compact open set 𝒱𝔞cc,𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}} satisfies

𝒱𝔞cc,𝔞¯¯𝒯𝔞c,\overline{\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}}\subset\mathcal{T}_{{\mathfrak{a}}^{c}},

and moreover, by Lemma 4.4 (iii), we may assume by taking 𝔞c{\mathfrak{a}}^{c} sufficiently close to 𝔞¯\bar{{\mathfrak{a}}} that the compact set

𝒜ˇ𝔞c𝔞¯,c𝒱𝔞cc,𝔞¯𝒯𝔞c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}_{{\mathfrak{a}}^{c}}\subset\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}\subset\mathcal{T}_{{\mathfrak{a}}^{c}}

has measure ν/8\geq\nu/8 with respect to the volume form η|𝒯𝔞c\eta_{|\mathcal{T}_{{\mathfrak{a}}^{c}}}. Consider a smooth function Gc:𝒯𝔞c[0,)G_{c}:\mathcal{T}_{{\mathfrak{a}}^{c}}\rightarrow[0,\infty) such that

Gc1({0})=(Fc𝒱𝔞cc,𝔞¯)𝒱𝔞cc,𝔞¯G^{-1}_{c}\left(\{0\}\right)=\left(F^{c}\cap\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}\right)\cup\partial\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}

and define for every ϵ>0\epsilon>0 the compact set (note that the set is compact because it does not intersect 𝒱𝔞cc,𝔞¯\partial\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}} where GcG_{c} is vanishing)

Ωϵc:=Gc1([ϵ,))𝒱𝔞cc,𝔞¯.\Omega^{c}_{\epsilon}:=G^{-1}_{c}\bigl{(}[\epsilon,\infty)\bigr{)}\cap\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}.

By Sard’s Theorem, GcG_{c} admits a decreasing sequence {ϵk}k\{\epsilon_{k}\}_{k\in\mathbb{N}} of regular values converging to 0. Thus we have

𝒱𝔞cc,𝔞¯=kΩϵkcwithΩϵkcΩϵk+1ck\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}=\bigcup_{k\in\mathbb{N}}\Omega^{c}_{\epsilon_{k}}\quad\mbox{with}\quad\Omega^{c}_{\epsilon_{k}}\subset\Omega^{c}_{\epsilon_{k+1}}\quad\forall k\in\mathbb{N}

and for every kk\in\mathbb{N} the set Ωϵkc\Omega^{c}_{\epsilon_{k}} is a compact, oriented, smooth submanifold with boundary of 𝒯𝔞c\mathcal{T}_{{\mathfrak{a}}^{c}} of dimension rr. As a consequence, since the measure of 𝒱𝔞cc,𝔞¯\mathcal{V}^{c,\bar{{\mathfrak{a}}}}_{{\mathfrak{a}}^{c}}, which contains 𝒜ˇ𝔞c𝔞¯,c\check{\mathcal{A}}^{\bar{{\mathfrak{a}}},c}_{{\mathfrak{a}}^{c}}, with respect to the volume form η|𝒯𝔞c\eta_{|\mathcal{T}_{{\mathfrak{a}}^{c}}} is ν/8\geq\nu/8, there is k¯\bar{k}\in\mathbb{N} large enough such that the measure (with respect to the volume form η|𝒯𝔞c\eta_{|\mathcal{T}_{{\mathfrak{a}}^{c}}}) of the set

𝒪c,0:=Ωϵk¯c\mathcal{O}^{c,0}:=\Omega^{c}_{\epsilon_{\bar{k}}}

is ν/16\geq\nu/16. By construction, 𝒪c,0\mathcal{O}^{c,0} is the union of finitely many components 𝒪jc,0\mathcal{O}^{c,0}_{j} satisfying properties (i), (iii), (viii) of the statement, where jj varies in a finite set JcJ^{c}. Then, for every jJcj\in J^{c}, we define Φjc:𝒪jc,0×[0,1]𝒮0\Phi_{j}^{c}:\mathcal{O}_{j}^{c,0}\times[0,1]\rightarrow\mathcal{S}_{0} by

Φjc:=H|𝒪jc,0×[0,1]c\Phi_{j}^{c}:=H^{c}_{|\mathcal{O}_{j}^{c,0}\times[0,1]}

and we set

𝒪jc,1:=Φjc(𝒪jc,0×{1}).\mathcal{O}_{j}^{c,1}:=\Phi_{j}^{c}\left(\mathcal{O}_{j}^{c,0}\times\{1\}\right).

The properties (ii), (iv), (v), (vi) and (vii) are satisfied by the construction together with Lemma 4.5 (iv). ∎

We are now ready to complete the proof of Theorem 1.1. Let us temporarily fix 0<c<ϵ0<c<\epsilon. By Lemma 4.6, there are a finite set JcJ^{c}, two collections of sets {𝒪jc,0}jJc\{\mathcal{O}_{j}^{c,0}\}_{j\in J^{c}}, {𝒪jc,1}jc\{\mathcal{O}_{j}^{c,1}\}_{j\in^{c}} and a collection of functions {Φjc:𝒪jc,0×[0,1]𝒮0}jJc\{\Phi_{j}^{c}:\mathcal{O}_{j}^{c,0}\times[0,1]\rightarrow\mathcal{S}_{0}\}_{j\in J^{c}} such that the properties (i)-(viii) are satisfied. Set for every jJcj\in J^{c} (see Figure 5)

jc:={Φjc(𝔞,t)|(𝔞,t)𝒪jc×[0,1]}.\mathcal{M}^{c}_{j}:=\Bigl{\{}\Phi_{j}^{c}({\mathfrak{a}},t)\,|\,({\mathfrak{a}},t)\in\mathcal{O}_{j}^{c}\times[0,1]\Bigr{\}}.
Refer to caption
Figure 5. The sets 𝒪jc,0,𝒪jc,1\mathcal{O}_{j}^{c,0},\mathcal{O}_{j}^{c,1} and jc\mathcal{M}_{j}^{c}

By properties (iii)-(vii), it is a topological manifold (with boundary) of dimension r+1r+1 whose boundary can be written as

jc=𝒪jc,0𝒪jc,1𝒞jc\partial\mathcal{M}_{j}^{c}=\mathcal{O}_{j}^{c,0}\cup\mathcal{O}_{j}^{c,1}\cup\mathcal{C}^{c}_{j}

where both 𝒪jc,0\mathcal{O}_{j}^{c,0} and 𝒪jc,1\mathcal{O}_{j}^{c,1} are compact, connected, oriented, smooth submanifolds with boundary (Lemma 4.6 (iii)-(iv)) and where the cylindrical part 𝒞jc\mathcal{C}^{c}_{j} given by

𝒞jc:={Φjc(𝔞,t)|(𝔞,t)𝒪jc,0×(0,1)}\mathcal{C}^{c}_{j}:=\Bigl{\{}\Phi_{j}^{c}({\mathfrak{a}},t)\,|\,({\mathfrak{a}},t)\in\partial\mathcal{O}_{j}^{c,0}\times(0,1)\Bigr{\}}

is a smooth open oriented submanifold of dimension r=2lr=2l satisfying

η|𝒞jc=0,\eta_{|\mathcal{C}^{c}_{j}}=0,

because any point of 𝒞jc\mathcal{C}^{c}_{j} has the form Φjc(𝔞,t)\Phi_{j}^{c}({\mathfrak{a}},t) with 𝔞𝒪jc,0{\mathfrak{a}}\in\partial\mathcal{O}_{j}^{c,0} and (by Lemma 4.6 (vii))

0Φjct(𝔞,t)(TΦjc(𝔞,t)𝒞jc)(TΦjc(𝔞,t)𝔞)0\neq\frac{\partial\Phi_{j}^{c}}{\partial t}({\mathfrak{a}},t)\in\left(T_{\Phi_{j}^{c}({\mathfrak{a}},t)}\mathcal{C}_{j}^{c}\right)\cap\left(T_{\Phi_{j}^{c}({\mathfrak{a}},t)}\mathcal{L}_{{\mathfrak{a}}}\right)
withTΦjc(𝔞,t)𝔞=𝒦(Φjc(𝔞,t))=ker(ωΦjc(𝔞,t)).\mbox{with}\quad T_{\Phi_{j}^{c}({\mathfrak{a}},t)}\mathcal{L}_{{\mathfrak{a}}}=\vec{\mathcal{K}}\left(\Phi_{j}^{c}({\mathfrak{a}},t)\right)=\mbox{ker}\left(\omega^{\perp}_{\Phi_{j}^{c}({\mathfrak{a}},t)}\right).

As a consequence, by applying Stokes’ Theorem we have for every jJcj\in J^{c},

𝒪jc,0η=𝒪jc,1η,\int_{\mathcal{O}_{j}^{c,0}}\eta=\int_{\mathcal{O}_{j}^{c,1}}\eta,

which imply (because, by Lemma 4.6 (i)-(ii), the sets 𝒪jc,0\mathcal{O}_{j}^{c,0} (resp. 𝒪jc,1\mathcal{O}_{j}^{c,1}) are pairwise disjoint)

(4.8) 𝒪c,0η=jJc𝒪jc,0η=jJc𝒪jc,0η=jJc𝒪jc,1η=jJc𝒪jc,1η=𝒪c,1η.\displaystyle\int_{\mathcal{O}^{c,0}}\eta=\int_{\cup_{j\in J^{c}}\mathcal{O}_{j}^{c,0}}\eta=\sum_{j\in J^{c}}\int_{\mathcal{O}_{j}^{c,0}}\eta=\sum_{j\in J^{c}}\int_{\mathcal{O}_{j}^{c,1}}\eta=\int_{\cup_{j\in J^{c}}\mathcal{O}_{j}^{c,1}}\eta=\int_{\mathcal{O}^{c,1}}\eta.

But, on the one hand, by Lemma 4.6 (viii), we have

𝒪c,0ην16,\int_{\mathcal{O}^{c,0}}\eta\geq\frac{\nu}{16},

and, on the other hand Lemma 4.3 (ii) together with equation (4.5) yield (g~|Xc\tilde{g}_{|X^{c}} denotes the metric induced by g~\tilde{g} on XcX^{c})

|𝒪c,1η|𝒪c,1|η|𝒪c,1δ(c)𝑑volg~|XcXcδ(c)𝑑volg~|Xcδ(c)C,\left|\int_{\mathcal{O}^{c,1}}\eta\right|\leq\int_{\mathcal{O}^{c,1}}|\eta|\leq\int_{\mathcal{O}^{c,1}}\delta(c)\,d\mbox{vol}^{\tilde{g}_{|X^{c}}}\leq\int_{X^{c}}\delta(c)\,d\mbox{vol}^{\tilde{g}_{|X^{c}}}\leq\delta(c)\,C,

which tends to zero as cc tends to zero. Thus (4.8) cannot be satisfied for all c>0c>0, this is a contradiction.

References

  • [1] Andrei A. Agrachev. Some open problems. In Geometric control theory and sub-Riemannian geometry, volume 5 of Springer INdAM Ser., pages 1–13. Springer, Cham, 2014.
  • [2] A. Belotto da Silva, A. Figalli, A. Parusiński, and L. Rifford. Strong Sard conjecture and regularity of singular minimizing geodesics for analytic sub-Riemannian structures in dimension 3. Invent. Math., 229(1):395–448, 2022.
  • [3] A. Belotto da Silva, A. Parusiński, and L. Rifford. Abnormal singular foliations and the sard conjecture for generic co-rank one distributions. arXiv 2310.20284, 2023.
  • [4] A. Belotto da Silva, A. Parusiński, and L. Rifford. Abnormal subanalytic distributions in sub-riemannian geometry. hal-04881557, 2025.
  • [5] André Belotto da Silva and Ludovic Rifford. The Sard conjecture on Martinet surfaces. Duke Math. J., 167(8):1433–1471, 2018.
  • [6] Francesco Boarotto and Davide Vittone. A dynamical approach to the Sard problem in Carnot groups. J. Differential Equations, 269(6):4998–5033, 2020.
  • [7] Marco Castelpietra and Ludovic Rifford. Regularity properties of the distance functions to conjugate and cut loci for viscosity solutions of Hamilton-Jacobi equations and applications in Riemannian geometry. ESAIM Control Optim. Calc. Var., 16(3):695–718, 2010.
  • [8] Isaac Chavel. Riemannian geometry, volume 98 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, second edition, 2006. A modern introduction.
  • [9] M. Golubitsky and V. Guillemin. Stable mappings and their singularities, volume Vol. 14 of Graduate Texts in Mathematics. Springer-Verlag, New York-Heidelberg, 1973.
  • [10] Robert M. Hardt. Some analytic bounds for subanalytic sets. In Differential geometric control theory (Houghton, Mich., 1982), volume 27 of Progr. Math., pages 259–267. Birkhäuser Boston, Boston, MA, 1983.
  • [11] Mw Hirsch. A stable analytic foliation with only exceptional minimal sets. Lecture Notes in Mathematics, 468:9–10, 1975.
  • [12] Enrico Le Donne, Richard Montgomery, Alessandro Ottazzi, Pierre Pansu, and Davide Vittone. Sard property for the endpoint map on some Carnot groups. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 33(6):1639–1666, 2016.
  • [13] Antonio Lerario, Luca Rizzi, and Daniele Tiberio. Quantitative approximate definable choices, 2024.
  • [14] J.-M. Lion and J.-P. Rolin. Intégration des fonctions sous-analytiques et volumes des sous-ensembles sous-analytiques. Ann. Inst. Fourier (Grenoble), 48(3):755–767, 1998.
  • [15] Carlo Mantegazza and Andrea Carlo Mennucci. Hamilton-Jacobi equations and distance functions on Riemannian manifolds. Appl. Math. Optim., 47(1):1–25, 2003.
  • [16] Richard Montgomery. A tour of subriemannian geometries, their geodesics and applications, volume 91 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.
  • [17] Alessandro Ottazzi and Davide Vittone. On the codimension of the abnormal set in step two Carnot groups. ESAIM Control Optim. Calc. Var., 25:Paper No. 18, 17, 2019.
  • [18] Ludovic Rifford. On viscosity solutions of certain Hamilton-Jacobi equations: regularity results and generalized Sard’s theorems. Comm. Partial Differential Equations, 33(1-3):517–559, 2008.
  • [19] Ludovic Rifford. Singulières minimisantes en géométrie sous-Riemannienne. Number 390, pages Exp. No. 1113, 277–301. 2017. Séminaire Bourbaki. Vol. 2015/2016. Exposés 1104–1119.
  • [20] Takashi Sakai. Riemannian geometry, volume 149 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 1996. Translated from the 1992 Japanese original by the author.
  • [21] Patrick Speissegger. The Pfaffian closure of an o-minimal structure. J. Reine Angew. Math., 508:189–211, 1999.
  • [22] Lou van den Dries and Chris Miller. Geometric categories and o-minimal structures. Duke Math. J., 84(2):497–540, 1996.
  • [23] I. Zelenko and M. Zhitomirski˘i. Rigid paths of generic 22-distributions on 33-manifolds. Duke Math. J., 79(2):281–307, 1995.