• Aucun résultat trouvé

Two-dimensional almost-Riemannian structures with tangency points

N/A
N/A
Protected

Academic year: 2022

Partager "Two-dimensional almost-Riemannian structures with tangency points"

Copied!
15
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Two-dimensional almost-Riemannian structures with tangency points

A.A. Agrachev

a

, U. Boscain

b

, G. Charlot

c

, R. Ghezzi

a

, M. Sigalotti

d,e,

aSISSA, via Beirut 2-4, 34014 Trieste, Italy

bCentre de Mathématiques Appliquées, École Polytechnique Route de Saclay, 91128 Palaiseau Cedex, France cInstitut Fourier, UMR 5582, CNRS/Université Grenoble 1, 100 rue des Maths, BP 74, 38402 St Martin d’Hères, France

dINRIA Nancy – Grand Est, Équipe-projet CORIDA, France

eInstitut Élie Cartan, UMR CNRS/INRIA/Nancy Université, BP 239, 54506 Vandœuvre-lès-Nancy, France Received 17 August 2009; received in revised form 22 November 2009; accepted 23 November 2009

Available online 3 December 2009

Abstract

Two-dimensional almost-Riemannian structures are generalized Riemannian structures on surfaces for which a local orthonor- mal frame is given by a Lie bracket generating pair of vector fields that can become collinear. We study the relation between the topological invariants of an almost-Riemannian structure on a compact oriented surface and the rank-two vector bundle over the surface which defines the structure. We analyse the generic case including the presence of tangency points, i.e. points where two generators of the distribution and their Lie bracket are linearly dependent. The main result of the paper provides a classifi- cation of oriented almost-Riemannian structures on compact oriented surfaces in terms of the Euler number of the vector bundle corresponding to the structure. Moreover, we present a Gauss–Bonnet formula for almost-Riemannian structures with tangency points.

©2009 Elsevier Masson SAS. All rights reserved.

1. Introduction

LetMbe a two-dimensional smooth manifold. A Riemannian distance on Mcan be seen as the minimum-time function of an optimal control problem where admissible velocities are vectors of norm one. The control problem can be written locally as

˙

q=uX(q)+vY (q), u2+v21, (1)

by fixing an orthonormal frame(X, Y ).

Almost-Riemannian structures generalize Riemannian ones by allowingXandY to be collinear at some points. If the pair(X, Y )is Lie bracket generating, i.e., if

This work was partially supported by the ANR “GCM” program “Blanc – CSD 5”.

* Corresponding author at: Institut Élie Cartan, UMR CNRS/INRIA/Nancy Université, BP 239, 54506 Vandœuvre-lès-Nancy, France.

E-mail addresses:agrachev@sissa.it (A.A. Agrachev), ugo.boscain@cmap.polytechnique.fr (U. Boscain), Gregoire.Charlot@ujf-grenoble.fr (G. Charlot), ghezzi@sissa.it (R. Ghezzi), Mario.Sigalotti@inria.fr (M. Sigalotti).

0294-1449/$ – see front matter ©2009 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2009.11.011

(2)

span

X(q), Y (q),[X, Y](q),

X,[X, Y] (q), . . .

=TqM

at everyqM, then (1) is completely controllable and the minimum-time function still defines a continuous distance d onM. Notice that a Riemannian distance can be globally defined onM by a control problem as in (1) only if the Riemannian structure admits a global orthonormal frame, implying thatMis parallelizable. More in general,Xand Y are parallel on a setZM(calledsingular locus), which is generically a one-dimensional embedded submanifold ofM(possibly disconnected).

Metric structures that can be defined locally by a pair of vector fields (X, Y ) through (1) are called almost- Riemannian structures. An almost-Riemannian structureScan be seen as an Euclidean bundleEof rank two overM (i.e. a vector bundle whose fibre is equipped with a smoothly-varying scalar product·,·q) and a morphism of vector bundlesf:ET Msuch that the evaluation atq of the Lie algebra generated by

{fσ|σ section ofE}

is equal toTqMfor everyqM, see Section 2 and also [17].

IfE is orientable, we say thatS isorientable. IfEis isomorphic to the trivial bundle M×R2, we say that the almost-Riemannian structure is trivializable. The singular locusZ is the set of pointsq of M at whichf (Eq)is one-dimensional. An almost-Riemannian structure is Riemannian if and only ifZ= ∅, i.e.f is an isomorphism of vector bundles.

The first example of genuinely almost-Riemannian structure is provided by the Grushin plane, which is the almost- Riemannian structure on M=R2 withE=R2×R2,f ((x, y), (a, b))=((x, y), (a, bx)) and ·,· the canonical Euclidean structure onR2. The model was originally introduced in the context of hypoelliptic operator theory [10,11]

(see also [4,8]). Notice that the singular locus is indeed nonempty, being equal to they-axis. Another example of (trivializable) almost-Riemannian structure appeared in problems of control of quantum mechanical systems (see [6,7]).

The notion of almost-Riemannian structure was introduced in [1]. In that paper, an almost-Riemannian structure is defined as a locally finitely generated Lie bracket generatingC(M)-submoduleof Vec(M), the space of smooth vector fields onM, endowed with a bilinear, symmetric mapG:×C(M)which is positive definite (in a suitable sense). This definition is equivalent to the one given above in terms of Euclidean bundles of rank two overM.

A pair of vector fields(X, Y )inis said to be anorthonormal frame forGon some open setΩifG(X, Y )(q)=0 andG(X, X)(q)=G(Y, Y )(q)=1 for everyqΩ. Equivalently, there exists a local orthonormal frame(σ, ρ)for

·,·such thatX=fσ,Y =fρ.

Almost-Riemannian structures present very interesting phenomena. For instance, even in the case where the Gaus- sian curvature is everywhere negative (where it is defined, i.e., on M\Z), geodesics may have conjugate points.

This happens for instance on the Grushin plane (see [1] and also [5] in the case of surfaces of revolution). Moreover it is possible to define non-orientable almost-Riemannian structures on orientable manifolds and orientable almost- Riemannian structures on non-orientable manifolds (see [1]).

This paper is a continuation of [1], where we provided a characterization of generic almost-Riemannian structures by means of local normal forms, and we proved a generalization of the Gauss–Bonnet formula. (For generalizations of Gauss–Bonnet formula in related contexts, see [2,16].) Let us briefly recall these results.

Theflag of a submoduleof Vec(M)is the sequence of submodules=12⊂ · · · ⊂m⊂ · · ·defined through the recursive formula

k+1=k+ [, k].

Denote by m(q) the set {V (q)|Vm}. Under generic assumptions, the singular locus Z has the following properties: (i)Z is an embedded one-dimensional submanifold ofM; (ii) the pointsqMat which2(q)is one- dimensional are isolated; (iii) 3(q)=TqM for everyqM. We say that S satisfies (H0) if properties (i), (ii), (iii) hold true. If this is the case, a point q of M is called ordinary if (q)=TqM, Grushin point if (q) is one-dimensional and 2(q)=TqM, i.e. the distribution is transversal to Z, and tangency pointif 2(q) is one- dimensional, i.e. the distribution is tangent toZ. If(Ω, X, Y )is a local generator ofsuch thatΩ\Z has exactly two components, then (X, Y )has different orientations on each of them. Normal forms for local generators of around ordinary, Grushin and tangency points are recalled in Section 2.1.

The main result of [1] is an extension of the Gauss–Bonnet theorem for orientable almost-Riemannian structures on orientable manifolds, under the hypothesis that there are not tangency points. More precisely, denote byK:M\Z

(3)

Rthe Gaussian curvature and byωa volume form for the Euclidean structure onE. LetdAs be the two-form on M\Z given by the pushforward ofωalong f. Fix an orientation Ξ onMand let M+(respectively, M) be the subset ofM\ZwheredAs is a positive (respectively, negative) multiple ofΞ. The main goal of [1] was to prove the existence and to assign a value to the limit

εlim0

Mε

K dAs, (2)

whereMε= {qM|d(q,Z) > ε}andd(·,·)is the distance globally defined by the almost-Riemannian structure onM. IfMhas no tangency point, then the limit in (2) can be shown to exist and to be equal to 2π(χ (M+)χ (M)), where χ denotes the Euler characteristic. When the almost-Riemannian structure is trivializable, we proved that χ (M+)=χ (M) whence the limit in (2) is equal to zero. Once applied to the special subclass of Riemannian structures, such a result simply states that the integral of the curvature of a parallelizable compact oriented surface (i.e., the torus) is equal to zero. In a sense, in the standard Riemannian construction the topology of the surface gives a constraint on the total curvature through the Gauss–Bonnet formula, whereas for an almost-Riemannian structure induced by a single pair of vector fields the total curvature is equal to zero and the topology of the manifold constrains the metric to be singular on a suitable set.

The main objective of this paper is to complete the analysis in the more complicated case in whichShas tangency points.

The following result provides a classification of almost-Riemannian structures in terms of the Euler number of the vector bundle associated with it (for a definition of the Euler number see Section 2).

Theorem 1.LetM be a compact oriented two-dimensional manifold endowed with an oriented almost-Riemannian structure S =(E, f,·,·) satisfying the generic hypothesis (H0). Then χ (M+)χ (M)+τ (S)=e(E), where e(E)denotes the Euler number ofEandτ (S)is the number of revolutions ofonZcomputed with respect to the orientation induced byM+onZ.

For a detailed definition ofτ (S)see Section 3. Notice that the Euler numbere(E)measures how far the vector bundleEis from the trivial one. Indeed,Eis trivial if and only ife(E)=0. As a direct consequence of Theorem 1 we get thatS is trivializable if and only ifχ (M+)χ (M)+τ (S)=0.

For what concerns the notion of integrability of the curvature with respect to the Riemannian density onM\Z, it turns out that the hypothesis made in [1] about the absence of tangency points is not just technical. Indeed, in Sec- tion 5.1 we present some numerical simulations strongly hinting that the limit in (2) diverges, in general, if tangency points are present. One possible explanation of this fact is the interaction between different orders in the asymptotic expansion of the almost-Riemannian distance. In this paper, to avoid this interference, we define a 3-scale integral of the curvature. LetT be the set of tangency points ofS. Associate with everyqT, a two parameters “rectangular”

neighborhoodBδq

121andδ2playing the role of lengths of the sides of the rectangle) built as follows. We consider a smooth curve(−1,1)sw(s)passing through the tangency pointw(0)=q and transversal to the distribution atq. We then consider, for eachs(−1,1), the geodesicγs (parameterized by arclength) such thatγs(0)=w(s)and minimizing locally the distance from{w(s)|s(−1,1)}(such geodesic exists and is unique due to the transversality of the curvew; for details see [1]). Forδ1, δ2sufficiently small, the rectangleBδq

12is the subset ofMcontaining the tangency pointq and having as boundary

γδ2

[−δ1, δ1]

γ[−δ22]1)γδ2

[−δ1, δ1]

γ[−δ22](δ1) (see Fig. 1). LetMε,δ12=Mε\ q∈T Bδq

12. We say thatKis 3-scaleS-integrableif

δlim10 lim

δ20 lim

ε0

Mε,δ12

K dAs (3)

exists and is finite. In this case we denote such a limit by

2MK dAs. The following result, proved in Section 5.3, is a generalization of the classical Gauss–Bonnet formula for Riemannian structures to generic oriented two-dimensional almost-Riemannian structures. It concludes the analysis started in [1] including the presence of tangency points.

(4)

Fig. 1. The rectangular boxBqδ

12.

Theorem 2.LetMbe a compact oriented two-dimensional manifold. If an oriented almost-Riemannian structureS onMsatisfies the hypothesis(H0), thenKis3-scaleS-integrable and

2MK dAs=2π e(E).

Notice that the construction ofMε,δ12 is not intrinsic since it depends on the choice of a manifold transversal to Z at each tangency point and on its parameterization. The result, however, is. An interesting question is whether a canonical way of choosing these manifolds and their parameterizations exists. This is related to the problem of finding intrinsic local normal forms for almost-Riemannian structures.

As a consequence of Theorems 1 and 2 we get the following corollary.

Corollary 1.LetMbe a compact oriented two-dimensional manifold. For an oriented almost-Riemannian structure S onMsatisfying the generic hypothesis(H0)we have

2MK dAs=0if and only ifS is trivializable. In particular, ifShas not tangency points then

MK dAs=0if and only ifSis trivializable.

These results complete the analysis of the relation between the integral of the curvature and the topology of the manifold for two-dimensional almost-Riemannian structures.

The structure of the paper is the following. In Section 2 we introduce the notion ofn-dimensional rank-varying sub-Riemannian structure and recall some definitions and results given in [1]. In Section 3 we define the number of revolutions of a one-dimensional distribution along an oriented closed curve. In Section 4 we prove Theorem 1. First we construct a section of the sphere bundle on a tubular neighborhood of the singular locus having each tangency point as singularity. Then, we extend the section to the complement in the manifold of a finite set and show the sum of the indexes at all the singularities to be equal to χ (M+)χ (M)+τ (S). In Section 5, the relation between tangency points and integrability of the curvature with respect to the area form associated with an almost-Riemannian structure is discussed. In particular, we provide in Section 5.1 some numerical simulations strongly hinting that in presence of tangency points the integral

MεK dAs does not converge asεtends to zero. This leads us to introduce in Section 5.2 the notion of 3-scaleS-integrability. Thanks to a Gauss–Bonnet formula for almost-Riemannian surfaces with boundary given in [9], we compute in Section 5.3 the total curvature of a generic two-dimensional almost- Riemannian structure with tangency points, proving Theorem 2.

(5)

2. Basic definitions

LetMbe ann-dimensional manifold. Throughout the paper, unless specified, manifolds are smooth (i.e.C) and without boundary; vector fields and differential forms are smooth. Given a vector bundleEoverM, the set of smooth sections ofE, denoted byΓ (E), is naturally endowed with the structure ofC(M)-module. In the caseE=T Mwe denoteΓ (E)by Vec(M).

Given an oriented vector bundle of rankn over a compact connected orientedn-manifoldM, the Euler number ofE, denoted bye(E), is the self-intersection number of MinE, whereMis identified with the zero section. To computee(E), consider a smooth sectionσ:MEtransverse to the zero section. Then, by definition,

e(E)=

p|σ (p)=0

i(p, σ ),

wherei(p, σ )=1, respectively−1, ifdpσ:TpMTσ (p)Epreserves, respectively reverses, the orientation. Notice that if we reverse the orientation onMor onEthene(E)changes sign. Hence, the Euler number of an orientable vector bundleEis defined up to a sign, depending on the orientations of bothEandM. Since reversing the orientation onMalso reverses the orientation ofT M, the Euler number ofT Mis defined unambiguously and is equal toχ (M), the Euler characteristic ofM. We refer the reader to [12] for a more detailed discussion of this subject.

Remark 1.Assume thatσΓ (E)has only isolated zeros, i.e. the set{p|σ (p)=0}is finite. IfEis endowed with a smooth scalar product·,·, we defineσ˜ :M\ {p|σ (p)=0} →SE byσ (q)˜ = σ (q),σ (q)σ (q) . Then ifσ (p)=0, i(p,σ )˜ =i(p, σ )is equal to the degree of the map∂BSn1that associate with eachq∂Bthe valueσ (q), where˜ Bis a neighborhood ofpdiffeomorphic to an open ball inRnthat does not contain any other zero ofσ.

Notice that ifi(p, σ )=0, the limit limqpσ (q)˜ does not exist and, in this case, we say thatσ˜ has a singularity atp.

Definition 1.An (n, k)-rank-varying distributionon an n-dimensional manifold M is a pair (E, f ) whereE is a vector bundle of rankk overM andf :ET M is a morphism of vector bundles, i.e., (i) if π:T MM and πE:EMdenote the canonical projections, then the diagram

E f

πE

T M

π

M

commutes and (ii)f is linear on fibers.

Moreover, we require the mapσfσfromΓ (E)to Vec(M)to be injective.

Remark 2. Definition 1 is equivalent to the definition of (n, k)-rank-varying distribution given in [1], namely a submodule of Vec(M) locally generated byk (and not less than k) vector fields. Given an (n, k)-rank-varying distribution(E, f ), we define= {fσ|σΓ (E)}. Thanks to (i)is a submodule of Vec(M). The injectivity assumption on the mapσfσ implies thatcannot be generated by less thankvector fields. In the following, we use either definition, depending on our convenience.

Let(E, f )be an(n, k)-rank-varying distribution,= {fσ|σΓ (E)}be its associated submodule and denote by(q)the linear subspace{V (q)|V} =f (Eq)TqM. Let Lie()be the smallest Lie subalgebra of Vec(M) containing and Lieq()= {V (q)|V ∈Lie()}for every qM. We say that(E, f ) satisfies theLie bracket generating conditionif Lieq()=TqMfor everyqM.

A property(P )defined for(n, k)-rank-varying distributions is said to begenericif for every vector bundleEof rankkoverM,(P )holds for everyf in an open and dense subset of the set of morphisms of vector bundles fromE toT M, endowed with theC-Whitney topology. E.g., generically, an(n, k)-rank-varying distribution is Lie bracket generating, provided thatk >1.

We say that an(n, k)-rank-varying distribution(E, f )isorientableifEis orientable as a vector bundle. Similarly, (E, f )istrivializableifEis isomorphic to the trivial bundleM×Rk. In this case, by definition, we havee(E)=0.

(6)

A rank-varying sub-Riemannian structure is defined by requiring thatEis an Euclidean bundle.

Definition 2. An (n, k)-rank-varying sub-Riemannian structureis a triple S =(E, f,·,·)where(E, f ) is a Lie bracket generating(n, k)-rank-varying distribution on a manifold M and·,·q is a scalar product onEq smoothly depending onq.

Remark 3. Definition 2 is equivalent to Definition 4 in [1], with G:×C(M) defined byG(V , W )= σV, σWwhereσV, σW are the unique sections ofEsatisfyingfσV =V , fσW=W.

Definition 2 generalizes several classical structures. First of all, a Riemannian structure(M, g)is an(n, n)-rank- varying sub-Riemannian structure with E=T M,f =1T M and·,· =g(·,·). Classical sub-Riemannian structures (see [3,15]) are (n, k)-rank-varying sub-Riemannian structures such thatEis a proper Euclidean subbundle ofT M andf is the inclusion. Finally, we definen-dimensional almost-Riemannian structures(n-ARSs for short) as(n, n)- rank-varying sub-Riemannian structures.

LetS=(E, f,·,·)be an (n, k)-rank-varying sub-Riemannian structure. For everyqM and everyv(q) define

Gq(v)=inf

u, uquEq, f (u)=v .

Ifσ1, . . . , σk is an orthonormal frame for·,·on an open subsetΩ ofM, anorthonormal frame forGonΩ is given byfσ1, . . . , fσk. One easily proves that orthonormal frames are systems of local generators of. Notice that S is trivializable if and only if there exists a global orthonormal frame for·,·. HenceS is trivializable if and only if there exists a global orthonormal frame forG.

A curveγ: [0, T] →Mis said to beadmissibleforSif it is Lipschitz continuous (with respect to the differential structure onM) and there exists a measurable essentially bounded function

[0, T] tu(t )Eγ (t )

called control function, such that γ (t )˙ =f (u(t )) for almost every t ∈ [0, T]. Given an admissible curve γ : [0, T] →M, thelength ofγ is

(γ )= T

0

Gγ (t )

γ (t )˙ dt.

The function(γ )is invariant under reparameterization of the curveγ. Moreover, if an admissible curveγ minimizes theenergy functionalJ (γ )=T

0 Gγ (t )(γ (t )) dt˙ (with fixedT and fixed endpoints) thenv=

Gγ (t )(γ (t ))˙ is constant andγ is also a minimizer of(·). On the other hand a minimizerγ of(·)such thatv is constant is a minimizer of J (·)withT =(γ )/v. AgeodesicforS is an admissible curveγ: [0, T] →Msuch that for every sufficiently small interval[t1, t2] ⊂ [0, T],γ|[t1,t2]is a minimizer ofJ (·). A geodesic for whichGγ (t )(γ (t ))˙ is (constantly) equal to one is said to be parameterized by arclength.

The distance induced bySonMis defined as d(q0, q1)=inf

(γ )γ (0)=q0, γ (T )=q1, γ admissible

. (4)

The finiteness and the continuity of d(·,·) with respect to the topology of M are guaranteed by the Lie bracket generating assumption on the rank-varying sub-Riemannian structure (see [13]). The distanced(·,·)endowsMwith the structure of metric space compatible with the topology ofMas smooth manifold.

2.1. Normal forms for generic 2-ARSs

Given a 2-ARSS, itssingular locusis the set of pointsqwhere(q)=f (Eq)has not maximal rank, that is, Z=

qMdim (q)

=1 .

Notice that dim((q)) >0 for every pointqM, becauseis a bracket generating distribution.

(7)

We say that S satisfies condition (H0) if the following properties hold: (i)Z is an embedded one-dimensional submanifold ofM; (ii) the pointsqMat which2(q)is one-dimensional are isolated; (iii)3(q)=TqMfor every qM, where1=andk+1=k+ [, k].

Proposition 1.(See [1].) Property(H0)is generic for2-ARSs.

ARSs satisfying hypothesis (H0) admit the following local normal forms.

Theorem 3.(See [1].) Given a2-ARSSsatisfying(H0), for every pointqMthere exist a neighborhoodUofqand an orthonormal frame(X, Y )forGonU, such that up to a smooth change of coordinates defined onU,q=(0,0) and(X, Y )has one of the forms

(F1) X(x, y)=(1,0), Y (x, y)=

0, eφ (x,y) , (F2) X(x, y)=(1,0), Y (x, y)=

0, xeφ (x,y) , (F3) X(x, y)=(1,0), Y (x, y)=

0,

yx2ψ (x) eξ(x,y)

,

whereφ,ξ andψare smooth real-valued functions such thatφ(0, y)=0andψ (0) >0.

LetSbe a 2-ARS satisfying (H0). A pointqMis said to be anordinary pointif(q)=TqM, hence, ifSis locally described by (F1). We callq aGrushin pointif(q)is one-dimensional and2(q)=TqM, i.e. if the local description (F2) applies. Finally, if (q)=2(q)has dimension one and3(q)=TqM then we say thatq is a tangency pointandS can be described nearq by the normal form (F3). We define

T = {qZ|qtangency point ofS}. 2.2. A Gauss–Bonnet formula for 2-ARSs

LetM be an orientable two-dimensional manifold andS=(E, f,·,·)an orientable 2-ARS onM. Notice that

·,·defines a Riemannian structure onM\Z. Denote byK the Gaussian curvature of such a structure and byωa volume form for the Euclidean structure onE. LetdAs be the two-form onM\Z given by the pushforward ofω alongf. Once an orientation onMis fixed,M\Zis split into two open setsM+andMsuch that the orientation onMcoincides with the one defined bydAs onM+and with its opposite onM.

For everyε >0 letMε= {qM|d(q,Z) > ε}, whered(·,·)is the almost-Riemannian distance (see Eq. (4)). We say thatKisS-integrableif

εlim0

Mε

K dAs

exists and is finite. In this case we denote such limit by

MK dAs. WhenShas no tangency pointsKhappens to beS-integrable and

MK dAs is determined by the topology ofM+ andM. This result, recalled in Theorem 4, can be seen as a generalization of Gauss–Bonnet formula to ARSs.

Theorem 4.(See [1].) LetMbe a compact oriented two-dimensional manifold, endowed with an oriented2-ARSS for which condition(H0)holds true. Assume thatShas no tangency points. ThenKisS-integrable and

MK dAs= 2π(χ (M+)χ (M)).

3. Number of revolutions of

LetMbe a compact oriented two-dimensional manifold andWan oriented closed simple curve inM. SinceMis oriented,T M|W is isomorphic to the trivial bundleW×R2and its projectivization is isomorphic toW×S1. Every subbundle Υ ofT M|W of rank one can be seen as a section of the projectivization ofT M|W, i.e. a smooth map Υ :WW×S1such thatπ1Υ =IdW, whereπ1:W×S1W denotes the projection on the first component.

We defineτ (Υ, W ), thenumber of revolutionsofΥ alongW, to be the degree of the mapπ2Υ :WS1, where

(8)

Fig. 2. Tangency points with opposite contributions.

π2:W ×S1S1is the projection on the second component. Notice thatτ (Υ, W ) changes sign if we reverse the orientation ofW.

Remark 4.LetM, W, Υ be as above and let us show how to computeτ (Υ, W ). LetVΓ (T W )be a never-vanishing vector field alongW. Then span(V )is a subbundle ofT M|Wof rank one and there exists an isomorphismt:T M|WW×R2such thattV (q)=(q, (1,0)). The trivializationt induces an isomorphism, still denoted byt, between the projectivization of T M|W andW×S1such thatπ2tV :WS1is constant. To simplify notations, we omitt in the following. Let π2V (q)θ0. Assume thatθ0is a regular value ofπ2Υ (otherwise there exists a smooth sectionΥ˜ homotopic toΥ havingθ0as a regular value). By definition,

τ (Υ, W )=

q|π2Υ (q)=θ0

sign

dq2Υ ) ,

wheredq denotes the differential atq of a smooth map. Notice that a pointq satisfiesπ2Υ (q)=θ0if and only if Υ (q)is tangent toWatq.

LetM be a compact oriented two-dimensional manifold andS be an oriented 2-ARS satisfying (H0). Then the singular locusZ is the boundary ofM+. Fix onZ the orientation induced byM+and let C(Z)denote the set of connected component ofZ. Sinceis one-dimensional alongZ, we can defineτ (S)=

WC(Z)τ (, W ).

Remark 5.LetSbe a 2-ARS satisfying hypothesis (H0). LetV , θ0, π2be as in Remark 4 withΥ=andWC(Z).

Since3(q)=TqMfor every pointqM,θ0is a regular value ofπ2. Moreover, the set of pointsqW such thatπ2(q)=θ0is the set of tangency points ofSbelonging toW. Hence,

τ (, W )=

qW∩T

sign

dq2) .

We defineτq=sign(dqπ2)(see Fig. 2). Clearly,τ (S)=

q∈Tτq. 4. Proof of Theorem 1

The idea of the proof is to find a sectionσ ofSEwith isolated singularitiesp1, . . . , pmsuch thatm

j=1i(pj, σ )= χ (M+)χ (M)+τ (S). In the sequel, we considerZto be oriented with the orientation induced byM+.

We start by definingσon a neighborhood ofZ. LetW be a connected component ofZ. SinceMis oriented, there exists an open tubular neighborhoodWofW and a diffeomorphismΨ :S1×(−1,1)→Wthat preserves the orien- tation andΨ|S1×{0}is an orientation-preserving diffeomorphism betweenS1andW. Remark thatf :E|W+TW+ is an orientation-preserving isomorphism of vector bundles, whilef:E|WTWis an orientation-reversing iso- morphism of vector bundles, whereW±=WM±. For everys=0, lift the tangent vector toθΨ (θ, s)toE usingf1, rotate it by the angleπ/2 and normalize it:σis defined as this unit vector (belonging toEΨ (θ,s)) ifs >0, its opposite ifs <0. In other words,σ:W\WSEis given by

(9)

σ (q)=sign(s) Rπ/2f1(∂Ψ∂θ(θ, s))

f1(∂Ψ∂θ(θ, s)), f1(∂Ψ∂θ(θ, s)), (θ, s)=Ψ1(q), (5)

whereRπ/2denotes the rotation (with respect to the Euclidean structure) inEby angleπ/2 in the counterclockwise sense. The following lemma shows thatσ can be extended to a continuous section fromW\T toSE.

Lemma 1.σ can be continuously extended to every pointqW\T.

Proof. LetqW\T,U be a neighborhood ofq inM and(x, y)be a system of coordinates onU centered atq such the almost-Riemannian structure has the form (F2) (see Theorem 3). Assume, moreover, thatU is a trivializing neighborhood of bothEandT M and the pair of vector fields(X, Y )is the image underf of a positively-oriented local orthonormal frame ofE. ThenWU= {(x, y)|x=0}. Since∂Ψ∂θ(θ,0)is non-zero and tangent toW,∂Ψ∂θ(θ,0) is tangent to the y-axis. Hence, thanks to the Preparation Theorem [14], there existh2:R→R,h1, h3:R2→R smooth functions such thath2(y)=0 for everyy∈Rand forΨ (θ, s)U

∂Ψ

∂θ(θ, s)=

xh1(x, y), h2(y)+xh3(x, y) ,

where(x, y)are the coordinates of the pointΨ (θ, s). Let us computeσ at a pointp(WU )\W. Since

∂Ψ

∂θ(θ, s)=xh1(x, y)X(x, y)+h2(y)+xh3(x, y)

xeφ (x,y) Y (x, y), then

f1 ∂Ψ

∂θ(θ, s)

=xh1(x, y)ζ (x, y)+h2(y)+xh3(x, y)

xeφ (x,y) ρ(x, y),

where(ζ, ρ)is the unique local orthonormal basis ofE|U such thatfζ =Xandfρ=Y. Notice thatUM+= {(x, y)|x >0}andUM= {(x, y)|x <0}. Using formula (5), for(x, y)=Ψ (θ, s)U\Wone easily gets

σ (x, y)=sign(x) l(x, y)

h2(y)+xh3(x, y)

xeφ (x,y) ζ+xh1(x, y)ρ

, where

l(x, y)=

x2h1(x, y)2+(h2(y)+xh3(x, y))2 x2e2φ (x,y) . Since

xlim0

sign(x)(h2(y)+xh3(x, y))

l(x, y)xeφ (x,y) = h2(y)

|h2(y)| and lim

x0

sign(x)xh1(x, y) l(x, y) =0, σ can be continuously extended to the set{x=0} =WU. 2

The next step of the proof is to show that for everyqWT,i(q, σ )=τq.

Lemma 2.Letσ:W\TSEbe the continuous section obtained in Lemma1. Then, for everyqWT the index ofσatqis equal toτqand, consequently,

q∈TW

i(q, σ )=τ (, W ). (6)

Proof. LetqWT,Ube a neighborhood ofqinMand(x, y)be a system of coordinates onUcentered atqsuch the almost-Riemannian structure has the form (F3) (see Theorem 3), i.e. a local orthonormal frame(X, Y )is given by

X=(1,0), Y= 0,

yx2ψ (x) eξ(x,y)

.

(10)

Define α=1, respectively −1, if (X, Y ) is the image under f of a positively-oriented, respectively negatively- oriented, local orthonormal frame ofE. One can check thatτq= −α. Let us make the following change of coordinates

˜

x=x, y˜=α

yx2ψ (x) . In these new coordinates,XandY become

X= 1,−α

2xψ (˜ x)˜ + ˜x2ψ(x)˜

, Y =

0,ye˜ ξ(x,α˜ y˜+ ˜x2ψ(x))˜

andWUis thex˜-axis. In the following, to simplify notations, we omit the tildes and denote the functionξ(x, α˜ y˜+

˜

x2ψ (x))˜ byξ(x, y). Since ∂ψ∂θ(θ,0)is tangent toW, by the Preparation Theorem [14] there existh1:R→R,h2, h3: R2→Rsmooth functions such thath1(x)=0 for everyx∈Rand forΨ (θ, s)U

∂ψ

∂θ(θ, s)=

h1(x)+yh2(x, y), yh3(x, y) ,

where(x, y)are the coordinates of the pointΨ (θ, s). This implies that

∂ψ

∂θ(θ, s)=

h1(x)+yh2(x, y) X(x, y)

+yh3(x, y)+α(h1(x)+yh2(x, y))(2xψ (x)+x2ψ(x))

yeξ(x,y) Y (x, y).

Let(ζ, ρ)be the local orthonormal frame ofEsuch thatX=fζ andY =fρ. From Eq. (5), it follows that σ (x, y)= −αsign(y)

l(x, y)

yh3(x, y)+α(h1(x)+yh2(x, y))(2xψ (x)+x2ψ(x))

yeξ(x,y) ζ

+αsign(y) l(x, y)

h1(x)+yh2(x, y) ρ, where

l(x, y)=

h1(x)+yh2(x, y)2

+

yh3(x, y)+α(h1(x)+yh2(x, y))(2xψ (x)+x2ψ(x)) yeξ(x,y)

2

. Notice that forx=0, y=0 we have

σ (0, y)= αsign(y)

(h1(0)+yh2(0, y))2+h3(0, y)2e2ξ(0,y)

eξ(0,y)h3(0, y)ζ+

h1(0)+yh2(0, y) ρ

,

whence the limit ofσ as(x, y)tends to(0,0)does not exist. Let us compute the index ofσatq=(0,0). Using Taylor expansions of the components ofσ in the basis(ζ, ρ)we find

σ (x, y)=αsign(y) l(x, y)

yh3(0,0)+2αxh1(0)ψ (0)

yeξ(0,0) +O

x2+y2 ζ+

h1(0)+O

x2+y2 ρ

. (7) Take a circlet(rcost, rsint )of radiusrcentered at(0,0)and assumerso small that(0,0)is the unique singu- larity ofσ on the closed disk of radiusr. By definition,i((0,0), σ )is half the degree of the map from the circleSr to RZthat associates to each point the angle between span(σ )and theζ. Using (7), this angle is

a(x, y)= −arctan

yh1(0)eξ(0,0)

yh3(0,0)+2αxψ (0)h1(0)+O

x2+y2 . Computingaalong the curvex(t )=rcost,y(t )=rsint, we find

a(rcost, rsint )= −arctan

sin(t )h1(0)eξ(0,0)

sin(t )h3(0,0)+2αcos(t )ψ (0)h1(0)+O(r)

.

Hence, by lettingrgo to zero, we are left to compute the degree of the mapa˜: [0,2π )→ [0, π )where

˜

a(t )= −arctan

sin(t )h1(0)eξ(0,0)

sin(t )h3(0,0)+2αcos(t )ψ (0)h1(0)

.

(11)

Since zero is a regular value ofa˜, the degree ofa˜is

t∈[0,2π )a(t )=0

sign

˜ a(t )

=sign

˜ a(0)

+sign

˜ a(π )

= −2α,

where the last equality follows froma˜(0)= ˜a(π )= −αeξ(0,0)/(2ψ (0)). Hence,i(q, σ )= −α. Sinceτq= −α, the lemma is proved (see Remark 5). 2

Let C(Z) denote the set of connected components of Z. Let Z˜ =

WC(Z)S1 and consider an orientation- preserving diffeomorphismΨ : ˜Z×(−1,1)→

WC(Z)Wsuch thatΨ|Z×{˜ 0}is an orientation-preserving diffeo- morphism ontoZ. Applying Lemma 1 to everyWC(Z)and reducing, if necessary, the cylindersW, we can assume that the set of singularities ofσonU=

WC(Z)WisT. Thenσ:U\TSEis continuous. Moreover, by Eq. (6),

q∈T

i(q, σ )=τ (S).

Extendσ toM\U. By a transversality argument, we can assume that the extended section has only isolated singular- ities{p1, . . . , pk} ∈M\Z. Since

e(E)= k

j=1

i(pj, σ )+

q∈T

i(q, σ )= k

j=1

i(pj, σ )+τ (S),

we are left to prove that k

j=1

i(pj, σ )=χ M+

χ M

. (8)

To this aim, consider the vector fieldF =fσ.F satisfiesG(F, F )≡1, whereG(·,·)is defined as in Remark 3 and the set of singularities ofF|M\Z is exactly{p1, . . . , pk}. Let us compute the index ofF at a singularityp∈ {p1, . . . , pk}. Sincef :E|M+T M+preserves the orientation andf :E|MT Mreverses the orientation, it follows thati(p, F )= ±i(p, σ ), ifpM±. Therefore,

k

j=1

i(pj, σ )=

j|pjM+

i(pj, F )

j|pjM

i(pj, F ). (9)

The theorem is proved if we show that

j|pjM+

i(pj, F )=χ M+

,

j|pjM

i(pj, F )=χ M

. (10)

To deduce Eq. (10), defineN+=M+\Ψ (Z˜×(0,1/2)). Notice that, by construction,σ|Ψ (Z×{˜ 1/2})is non-singular, hence the same is true forF|Ψ (Z×{˜ 1/2}). Moreover, the almost-Riemannian angle betweenTq(∂N+)and span(F (q))is constantly equal toπ/2. HenceF|∂N+points towardsN+and applying the Hopf’s Index Formula to every connected component ofN+we conclude that

j|pjM+

i(pj, F )=

j|pjN+

i(pj, F )=χ N+

=χ M+

.

Similarly, we find

j|pjM

i(pj, F )=χ M

.

(12)

Fig. 3. Graph ofKfor=span((1,0), (0, yx2)).

5. S-integrability in presence of tangency points

5.1. Numerical simulations

In this section we provide some numerical simulations hinting that, whenT = ∅,

Mε

K dAs

does not converge, in general, asεtends to zero.

From the proof of Theorem 4 we know that far from tangency points the integral of the geodesic curvature along

∂Mε+ and∂Mε offset each other forε going to zero. Hence, to understand whether the presence of a tangency point may lead to non-S-integrability ofK it is sufficient to compute the geodesic curvature of∂Mε+and∂Mεin a neighborhood of such a point. More precisely consider the almost-Riemannian structure(E, f,·,·)onM=R2 for whichE=R2×R2,f ((x, y), (a, b))=((x, y), (a, b(yx2)))and·,·is the canonical scalar product. For this system one has

K=−2(3x2+y) (x2y)2

.

The graph ofKis illustrated in Fig. 3. Notice that lim supq(0,0)K(q)= +∞and lim infq(0,0)K(q)= −∞. This situation is different from the Grushin case whereK(q)diverges to−∞asq approachesZ.

For everyε >0, the sets∂Mε+and∂Mεare smooth manifolds except at their intersections with the vertical axis x=0, which is the cut locus for the problem of minimizing the distance fromZ= {(x, x2)|x∈R}. Fix 0< a <1 and consider the two geodesics starting from the point (a, a2)and minimizing (locally) the distance from Z. Let P+ andP be the two points along these geodesics at distanceε fromZ. Denote byγ+ andγ the portions of

∂Mε+and∂Mεconnecting the vertical axis to the pointsP+andP, oriented as in Fig. 4. It is easy to approximate numericallyγ+andγby broken lines, but the evaluation of the integral of their geodesic curvatures is very unstable since its computation involves the second derivative of the curve parameterized by arclength. To avoid this problem, we rather apply the Riemannian Gauss–Bonnet formula on the regionsΩ+andΩintroduced in Fig. 4. This works better since the integral of the Gaussian curvature on Ω+ andΩ is numerically stable, and the integral of the geodesic curvature on horizontal and vertical segments can be computed analytically (in particular it is always zero on horizontal segments). Fig. 5 shows the value of

(13)

Fig. 4. RegionsΩ±where to apply Riemannian Gauss–Bonnet formula.

Fig. 5. Divergence of theS-integral ofK.

ε

γ+

Kgds

γ

Kgds

fora=0.1 andεvarying in the interval[0.01,0.04]. The graph seems to converge asεtends to zero to a non-zero constant, strongly hinting at the divergence of

MεK dAs. 5.2. More general notion ofS-integrability

The simulations of the previous section lead us to introduce the following alternative notion of integrability.

(14)

Definition 3(3-scaleS-integrability).LetqT andUq be a neighborhood ofqsuch that an orthonormal frame for GonUq is given by the normal form (F3). Forδ1,δ2>0 sufficiently small the rectangle[−δ1, δ1] × [−δ2, δ2]is a subset ofUqdenoted byBδq

12. For everyε >0, define Mε,δ12=Mε\

q∈T

Bδq

12. (11)

We say thatKis 3-scaleS-integrable if

δlim10 lim

δ20 lim

ε0

Mε,δ12

K dAs (12)

exists, is finite and does not depend on the choice of the normal form. In this case we denote such limit by

2MK dAs. Remark 6.Notice that ifT = ∅, then the concepts ofS-integrability and 3-scaleS-integrability coincide.

The order in which the limits are taken in (12) is important. Indeed, if the order is permuted, then the result given in Theorem 2 does not hold anymore.

Recall that the normal form (F3) is not totally intrinsic, since the functions ψ andφ depend on the choice of a parameterized smooth curve passing through the tangency point and transversal to the distribution at the point.

5.3. Proof of Theorem 2

Let us recall the following Gauss–Bonnet-like formula for domains whose boundary isC2in a neighborhood ofZ.

Theorem 5.(See [9, Theorem 5.2].) LetU be an open bounded connected subset ofMsuch that (i) Ucontains only ordinary and Grushin points,

(ii) ∂Uis piecewiseC2,

(iii) ∂UisC2in a neighborhood ofZ,

(iv) ∂Uis the union of the supports of a finite set of curvesγ1, . . . , γmthat are admissible forand of finite length.

DefineUε±=Mε±U. Then the following limits exist and are finite

U

K dAs:=lim

ε0

Uε+Uε

K dAs, (13)

∂U

kgs:=lim

ε0

∂U∂Uε+

kg

∂U∂Uε

kg

, (14)

where we interpret each integral

∂U∂Uε±kg as the sum of the integrals along theC2portions of∂U∂Uε±, plus the sum of the angles at the points of∂U∂Uε±where∂Uis notC1. Moreover, we have

U

K dAs+

∂U

kgs=2π χ

U+

χ U

. (15)

Fixδ1andδ2in such a way that the rectanglesBδq

12 are pairwise disjoint andZ∂Bδq

12⊂ [−δ1, δ1] × {δ2}, for everyqT. By construction,∂Bδq

12 is the union of the supports of four admissible curves for everyqT. Hence we can takeMε\ q∈T Bδq

12 asUin Theorem 5. As a consequence, we have

εlim0

Mε,δ12

K dAs+

q∈T

∂Bδq

12

kgs=2π

χ

M+\

q∈T

Bδq

12

χ

M\

q∈T

Bδq

12

=2π χ

M+

χ M

=2π

e(E)τ (S) ,

Références

Documents relatifs

— The leaf space M/F of a closed codimension two rie- mannian foliation on a simply connected closed manifold M is a 2-sphere riemannian orbifold with at most two singular

From this we can deduce analo- gous results for timelike immersions of Lorentzian surfaces in space forms of corresponding signature, as well as for spacelike and timelike immer-

For the local biH¨ older equivalence of E to a minimal cone, we do not need so much more than this, but for the C 1,α equivalence, we shall also need to know that the distance in

An almost-Riemannian structure (ARS in short) on an n-dimensional differential manifold can be defined, at least locally, by a set of n vector fields, considered as an

Theorem B can be combined with Sunada's method for constructing isospectral pairs of riemannian manifolds to produce a « simple isospectral pair»: a pair (Mi.gi), (Mg,^) with spec

In this section we use the techniques and results from [7], developped in the sub- Riemannian Martinet case, to compute asymptotics of the front from the tangency point for the

On an n-dimensional connected Lie group the simplest ARSs are defined by a set of n − 1 left-invariant vector fields and one linear vector field, the rank of which is equal to n on

The central limit theorem by Salem and Zygmund [30] asserts that, when properly rescaled and evaluated at a uniform random point on the circle, a generic real trigonometric