• Aucun résultat trouvé

Main iteration

Dans le document arXiv:1606.02551v1 [math.AP] 8 Jun 2016 (Page 25-28)

We start by noticing that Theorem 2.1.2 is a “strict subset” of Theorem 2.1.4: if Σ is closed, then the limit set of any map is empty. Moreover, the following simple topological fact will be used several times:

Lemma 2.2.1. Let Σ be a differentiable n-dimensional manifold and {Vλ} an open cover of Σ. Then there is an open cover {U} with the properties that:

(a) each U is contained in some Vλ;

(b) the closure of each U is diffeomorphic to an n-dimensional ball;

(c) each U intersects at most finitely many other elements of the cover;

(d) each pointp∈Σhas a neighborhood contained in at mostn+ 1elements of the cover;

(e) {U} can be subdivided into n+ 1 classes Ci consisting of pairwise disjoint U’s.

Proof. By a classical theorem Σ can be triangulated (see [96]) and by locally refining the triangulation we can assume that each simplex is contained in some Vλ. Denote byS such triangulation and enumerate its vertices as {Si0}, its 1-dimensional edges as {Si1} and so on. Then take the barycentric subdivision of S and call it T.

• For each vertex Si0 consider the interior Ui0 of the star of Si0 in the triangulation T (recall that the star of Si0 is the union of all simplices of the triangulation which contain Si0, cf. [43, p. 178]. Observe that the Ui0 are pairwise disjoint.

• For each edgeSi1 consider the interiorUi1 of the star ofSi1 in the triangulationT. The Ui1 are pairwise disjoint. Additionally, observe that ifUi1∩Uj0 6=∅, then Sj0 ⊂Si1.

• Proceed likewise up to n−1. Complete the collection {Uit : 0≤t ≤n−1} with the interiors Uin of the n-dimensional simplices Sin of S.

(a) and (e) are obvious by construction. If two distinct elementsUis andUjthave nonempty intersection and s ≥t, then s > t and Sjt is a face of Sis: this gives directly (c). (d) is an obvious consequence of (e). EachU is diffeomorphic to the open Euclideann-dimensional ball, but its closure is only homeomorphic to the closed ball: however it suffices to choose an appropriate smaller open set for each U to achieve (b) while keeping all the other properties.

From now on we fix therefore a smooth manifold Σ as in Theorem 2.1.4 and a corre-sponding smooth atlas A = {U} (which is either finite or countably infinite) where the U’s have compact closure and satisfy the properties (b), (c) and (d) of Lemma 2.2.1.

Given any symmetric (0,2) tensor h on Σ we write h = hijdxi ⊗dxj and denote by khk0,U the supremum of the Hilbert-Schmidt norm of the matrices hij(p) for p ∈ U.

Similarly, if v : Σ→RN is a C1 map, we write kDvk0,U for the supremum of the Hilbert-Schmidt norms of the matrices Dv(p) = (∂1v(p), . . . , ∂nv(p)), where p ∈ U. Finally we set

khk0 := sup

khk0,U

kDvk0 := sup

kDvk0,U.

We are now ready to state the main inductive statement2 whose iteration will prove The-orem 2.1.4.

Proposition 2.2.2. Let (Σ, g) be as in Theorem 2.1.4 and w : Σ→RN a smooth strictly short immersion. For any choice of positive numbers η > 0 and any δ > 0 there is a smooth short immersion z: Σ→RN such that

kz−wk0,U < η, ∀ℓ (2.3)

kg−zek0 < δ , (2.4)

kDw−Dzk0 < Cp

kg−wek0, (2.5)

for some dimensional constant C. If w is injective, then we can choose z injective.

Note that the right hand side of (2.5) might be ∞, in which case the condition (2.5) is an empty requirement. We show first how to conclude Theorem 2.1.4 from the proposition above. Subsequently we close this section by proving Corollary 2.1.5. The rest of the chapter will then be dedicated to prove Proposition 2.2.2.

Proof of Theorem 2.1.4. Let v0 := v and ε be as in the statement and assume for the moment thatv is an immersion. For the first part we can assume without loss of generality that the map is strictly short, after multiplying it by a constant smaller (but close to) 1.

We will produce a sequence of maps vq by applying iteratively Proposition 2.2.2. Since the limit set of v is closed and v(U) compact, there is a positive number β such that any point ofv(U) is at distance at leastβ from the limit set ofv. We then define the numbers

¯

ηq,ℓ:=2−q−1min{ε, β,2−ℓ} δq:=4−q.

At each q ≥ 1 we then apply Proposition 2.2.2 with w = vq−1, η = ¯ηq,ℓ and δ = δq to produce z =:vq. We then conclude immediately that:

• kvq −vq−1k0 ≤ 2−q−1ε and thus vq converges uniformly to some u with ku−vk0 ≤ εP

q≥12−q−1 = 2ε;

• similarly kv−uk0,U ≤βP

q≥12−q−1 = β2;

2This is what Nash calls “a stage”, cf. [69, Page 391]

• again by a similar computationku−vk0,U ≤2−ℓand thus the limit set ofucoincides with the limit set ofv; combined with the estimate above, this implies that the limit set ofu does not intersect the image of u;

• kDvq−Dvq−1k0 ≤C2−q+1 for every q≥2 and thus u is a C1 map (observe that we claim no bound on kDv1−Dv0k0; on the other hand we do not need it!);

• since vq converges to u in C1, we have g−ue= limq(g−vqe) = 0 and thus u is an isometry, from which we also conclude that the differential of u has everywhere full rank and hence u is an immersion.

It remains to show that, if v is injective, then the iteration above can be arranged so to guarantee that u is also injective. To this aim, notice first that all the conclusions above certainly hold in case we implement the same iteration applying Proposition 2.2.2 with parameters ηq,ℓ smaller than ¯ηq,ℓ. Moreover the proposition guarantees the injectivity at each step: we just need to show that the limit map is also injective. For each q consider the compact set Vq :=∪ℓ≤qU and the positive numbers

i := min{|vi(x)−vi(y)|:d(x, y)≥2−i, x, y∈Vi} for i < q,

where d is the geodesic distance induced by the Riemannian metric g. We then set ηq,ℓ :=

min{η¯q,ℓ,2−q−1γ1,2−q−1γ2, . . . ,2−q−1γq−1}and apply the iteration as above withηq,ℓin place of ¯ηq,ℓ. We want to check that the resulting u is injective. Fix x 6= y in Σ and choose q such that 2−q ≤d(x, y) and x, y ∈Vq. We can then estimate

|u(x)−u(y)| ≥ |vq(x)−vq(y)| −X

k≥q

kvk+1−vkk0,Vq ≥2γq−X

k≥q

2−k−1γq ≥γq>0. Proof of Corollary 2.1.5. Recall that, according to Whitney’s embedding theorem in its strong form (see [98]), any smooth differentiable manifold Σ of dimension n can be em-bedded in R2n. If the manifold in addition is closed, then it suffices to multiply the corre-sponding map by a sufficiently small positive constant to make it short and the existence of a nearby C1 isometry with the desired property follows from Theorem 2.1.2.

The general case requires somewhat more care. Fix a smooth Riemannian manifold (Σ, g) of dimension n, not closed. Below we will produce a suitable smooth embedding z : Σ→RN for N = (n+ 1)(n+ 2), with the additional properties that

(i) z is a short map;

(ii) the limit set ofz is {0} and does not intersect the image ofz.

We then can follow the standard procedure of the proof of the Whitney’s embedding theorem in its weak form (cf. [97]): if we consider the Grassmannian of 2n+ 1 dimensional planes π of RN, we know that for a subset of full measure the projection Pπ onto π is injective and has injective differential on z(Σ). A similar argument shows that for a set of planes π of full measure Pπ(z(Σ)) does not contain the origin. Since clearly Pπ ◦z is

also short, the map v := Pπ ◦z satisfies the assumptions of Theorem 2.1.4. If we drop the injectivity assumption on π (namely we restrict to immersions), we can project on a suitable 2n-dimensional plane.

Coming to the existence of z, we use the atlas {U} of Σ given by Lemma 2.2.1 and we let Φ :U → Rn be the corresponding charts. Observe that, since Σ is not closed, the atlas is necessarily (countably) infinite. After further multiplying each Φ by a positive scalar we can assume, without loss of generality, that |Φ| ≤ 1. Recall the n+ 1 classes Ci of Lemma 2.2.1(e). Consider then a family of smooth functions ϕ, each supported in U, with 0 ≤ ϕ ≤ 1 and such that for any point p ∈ Σ there is at least one ϕ which is identically 1 in some neighborhood ofp. Finally, after numbering the elements of the atlas, we fix a vanishing sequence ε of strictly monotone positive numbers, whose choice will be specified in a moment.

We are now ready to define our map z, which will be done specifying each component zj. Fix p ∈Σ and i ∈ {1, . . . , n+ 1}. If p does not belong to any element of Ci, then we set z(i−1)(n+2)+1(p) =. . .=zi(n+2)(p) = 0. Otherwise there is a unique U ∈ Ci with p∈U

and we set:

z(i−1)(n+2)+j(p) =ε2ϕ(p)(Φ(p))j for j ∈ {1, . . . , n} (2.6)

z(i−1)(n+2)+n+1(p) =ε2ϕ(p) (2.7)

z(i−1)(n+2)+n+2(p) =εϕ(p). (2.8)

Now, for any pointpthere is at least oneℓfor whichϕis identically one in a neighborhood of p: this will have two effects, namely that the differential of z at p is injective and that z(p) 6= 0. Since the limit set of z is obviously 0, condition (ii) above is satisfied. To prove that z is an embedding we need to show that z is injective. Fix two points p and q and fix a U ∈ Ci for which ϕ(p) = 1. If q ∈ U, then either ϕ(q) 6= 1, in which case z(i−1)(n+2)+n+1(p) 6= z(i−1)(n+2)+n+1(q), or ϕ(q) = 1 and, by Φ(p) 6= Φ(q), we conclude z(q)6=z(p). If q 6∈U and ϕ(q) = 0 for any other U ∈ Ci, then z(i−1)(n+2)+n+1(q) = 0 6= z(i−1)(n+2)+n+1(p). Otherwise there is a U ∈ Ci distinct from U such that ϕ(q) 6= 0. In this case we have

z(i−1)(n+2)+n+1(p)

zi(n+2)(p) =ε 6=ε = z(i−1)(n+1)+n+1(q) zi(n+2)(q) . Thus z is injective.

Finally, by choosing the ε inductively appropriately small, it is easy to show that we can ensure the shortness ofz.

Dans le document arXiv:1606.02551v1 [math.AP] 8 Jun 2016 (Page 25-28)