• Aucun résultat trouvé

Invertibility of Sobolev mappings under minimal hypotheses

N/A
N/A
Protected

Academic year: 2022

Partager "Invertibility of Sobolev mappings under minimal hypotheses"

Copied!
12
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Invertibility of Sobolev mappings under minimal hypotheses

Leonid V. Kovalev

a,1

, Jani Onninen

a,2

, Kai Rajala

b,,3

aDepartment of Mathematics, Syracuse University, Syracuse, NY 13244, USA

bDepartment of Mathematics and Statistics, University of Jyväskylä, PO Box 35 (MaD), FI-40014, University of Jyväskylä, Finland Received 15 May 2009; received in revised form 14 September 2009; accepted 14 September 2009

Available online 1 October 2009

Abstract

We prove a version of the Inverse Function Theorem for continuous weakly differentiable mappings. Namely, a nonconstant W1,nmapping is a local homeomorphism if it has integrable inner distortion function and satisfies a certain differential inclusion.

The integrability assumption is shown to be optimal.

©2009 Elsevier Masson SAS. All rights reserved.

MSC:primary 30C65; secondary 26B10, 26B25

Keywords:Local homeomorphism; Differential inclusion; Finite distortion

1. Introduction

Throughout this paper Ω is a bounded domain in Rn. The classical Inverse Function Theorem states that if f:Ω→Rnis continuously differentiable and the differential matrixDf (x)is invertible at some pointx, thenf is a homeomorphism in a neighborhood ofx. We are interested in a version of the Inverse Function Theorem for con- tinuous weakly differentiable mappings. In this context the invertibility of the differential matrix is not sufficient. As an example, consider the winding mappingf:R3→R3written in cylindrical coordinates asf (r, θ, z)=(r,2θ, z).

Althoughf is Lipschitz and its Jacobian determinantJ (x, f )equals 2 for a.e. x∈Rn, this mapping is not a local homeomorphism.

Let us introduce the following subset ofn×nmatrices.

M(δ)=

A∈Rn×n: Aξ, ξδ|||ξ|for allξ∈Rn ,

where−1δ1. Note thatδ= −1 imposes no condition on the matrix. When−1< δ <0, the setM(δ) is not convex and the differential inclusion

* Corresponding author.

E-mail addresses:lvkovale@syr.edu (L.V. Kovalev), jkonnine@syr.edu (J. Onninen), kirajala@maths.jyu.fi (K. Rajala).

1 Supported by the NSF grant DMS-0913474.

2 Supported by the NSF grant DMS-0701059.

3 Supported by the Academy of Finland.

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

doi:10.1016/j.anihpc.2009.09.010

(2)

Df (x)M(δ) for a.e.xΩ, (1.1) cannot be integrated to yield a pointwise inequality forf.

The winding mapping does not satisfy (1.1) for anyδ >−1. Even so, this differential inclusion does not by itself guarantee thatf is locally invertible, e.g., f (x1, x2)=(x1,0). There are also such examples with strictly positive Jacobian [14, Example 18]. To quantify the invertibility of a matrix A∈Rn×n, we introduce the inner distortion KI(A)∈ [1,∞].

KI(A)=

⎧⎪

⎪⎩

An

(detA)n1, detA >0,

1, A=0,

, otherwise.

(1.2)

HereAstands for the cofactor matrix ofAand·is the operator norm. To shorten the notation we writeKI(x, f )= KI(Df (x))and

KΩ[f] := 1

|Ω|

Ω

KI(x, f )dx,

where|Ω| is the Lebesgue measure ofΩ. IffW1,n(Ω,Rn)andKI(x, f ) <∞ a.e., thenf has a logarithmic modulus of continuity [4,9]; that is,

f (a)f (b) nC(n)

2BDfn

log(e+2 diamB|ab| ), a, bB, 2BΩ.

In this paper we always takef to be its continuous representative.

If moreoverKΩ[f]<∞andf is invertible, then the inverseh:=f1is aW1,n-mapping and

Ω

KI(x, f )dx=

f (Ω)

Dhn,

see [1, Theorem 9.1]. ThusKΩ[f]controls the modulus of continuity off1, should it exist. Our main result ad- dresses its existence.

Theorem 1.1.Suppose thatfWloc1,n(Ω,Rn)is a nonconstant mapping such thatKΩ[f]<. If there existsδ >−1 such thatDf (x)M(δ)for almost everyxΩ, thenf is a local homeomorphism.

This theorem is already known in the planar casen=2 [14, Theorem 4]. The assumptionKΩ[f]<∞cannot be replaced by

ΩKIq(x, f )dx <∞for anyq <1, see [14, Example 18] or [2, Example 1].

Our proof of Theorem 1.1 is based on two results of independent interest. The first step toward proving that a mapping is a local homeomorphism is to show that it is discrete and open; that is, preimages of points are discrete sets and images of open sets are open.

Theorem 1.2.Letf:Ω→Rnbe a mapping inWloc1,n(Ω,Rn)such thatJ (x, f ) >0a.e. If(Df )1L(Ω,Rn×n), thenf is discrete and open.

The challenging Iwaniec–Šverák conjecture asserts even more: a nonconstant mapping fWloc1,n(Ω,Rn)with KΩ[f]<is discrete and open. So far this conjecture was proved only forn=2 in [10]. Partial results in this direction were recently obtained in [6–8,15,19,20].

Another crucial ingredient of our proof of Theorem 1.1 is an estimate for the multiplicity N (y, f, A) :=

#(f1(y)A) of a local homeomorphismf in terms of the integral ofKI(·, f )in dimensionsn3. This result (Theorem 5.1) continues the line of development that began in 1967 with the celebrated Global Homeomorphism Theorem of Zorich [24].

The proof of Theorem 1.1 proceeds as follows. The differential inclusion (1.1) allows us to approximate f by mappings fλ(x):=f (x)+λx to which Theorem 1.2 can be applied. The results of [14] yield that fλ is a local

(3)

homeomorphism. By virtue of Theorem 5.1 the mappingsfλ have uniformly bounded multiplicity, which leads to a bound for the essential multiplicity off. This additional information suffices to show thatf is discrete and open, see Proposition 2.2 below. Sincef is a limit of local homeomorphismsfλ, the conclusion follows.

Different approaches to the invertibility of Sobolev mappings were pursued in [2,3,5,16,18,22], see also references therein.

2. Background

In this section we collect necessary notation and preliminaries. An open ball with centeraand radiusris denoted byB(a, r):= {x∈Rn:|xa|< r}. Its boundary is the sphereS(a, r). Ifλ >0 andB=B(a, r), thenλB=B(a, λr) andλS=S(a, λr). In addition,B=B(0,1),Br=B(0, r),S=S(0,1)andSr=S(0, r).

LetHdstand for thed-dimensional Hausdorff measure which agrees with the Lebesgue measure whendcoincides with the space dimension. The Hausdorff distance dH(E, F )between nonempty bounded setsEandF is defined as the infimum of numbers >0 such that the-neighborhood ofEcontainsF and vice versa.

Given a continuous mappingf:Ω →Rn and a setEΩ, we denote by N (y, f, E)the cardinality (possibly infinite) of the setf1(y)E. Ify∈Rn\f (∂Ω), the local degree off aty with respect to a domainGΩ is denoted by deg(y, f, G). We writef:A−→hom Bto indicate thatf is a homeomorphism fromAontoB.

LetΓ be a family of paths (parametrized curves) inRn,n2. The image ofγΓ is denoted by|γ|. We letΥΓ be the set of all Borel functionsρ:Rn→ [0,∞]such that

γ

ρds1

for every locally rectifiable pathγΓ. The functions inΥΓ are called admissible forΓ. For a given weightω:Rn→ [0,∞]we define

MωΓ = inf

ρΥΓ

ρ(x)nω(x)dx,

and callMωΓ the weighted conformal modulus ofΓ. Here it suffices to haveωdefined on a Borel set containing

γΓ|γ|. Whenω≡1 we obtain the conformal modulusMΓ. We will also use the spherical modulus with respect to a sphereS,

MSΓ = inf

ρΥΓ

S

ρ(y)ndHn1(y).

The reader may wish to consult the monographs [21,23] for basic properties of moduli of path families. The following generalization of the Poletsky inequality relates the moduli ofΓ and of its image underf, denoted byf Γ.

Proposition 2.1.(See [12].) Suppose thatfW1,n(Ω,Rn)is a discrete and open mapping with KΩ[f]<∞. If Γ is a family of paths contained inΩ, then

Mf Γ MKI(·,f )Γ. (2.1)

We will use the following result, which establishes the Iwaniec–Šverák conjecture under an additional assumption on the multiplicity off.

Proposition 2.2.Suppose thatfWloc1,n(Ω,Rn)is a nonconstant mapping withKΩ[f]<∞. LetB be a ball such that2BΩ. If

ess lim sup

r0

r1n

S(a,r)

N (y, f, B)dHn1(y) <∞ (2.2)

for everya∈Rn, thenf is discrete and open inB.

(4)

This proposition is a consequence of [20, Theorem 2.2]. Although [20, Theorem 2.2] requires that ess sup

0<t <1

∂(t B)

Df (x)

|f (x)a|n1dHn1(x) <, this condition is only used to obtain (2.2).

3. Preliminary results

For the sake of brevity, the connected component of a set A that contains a pointxA will be called thex- component ofA.

Proposition 3.1.Suppose thatfWloc1,n(Ω,Rn)is a mapping such thatKΩ[f]<. LetxΩ andy=f (x). If thex-component off1(y)is{x}, thenf is discrete and open in some neighborhood ofx.

Proof. Pickr >0 such thatB(x, r)Ω and letUj be thex-component of(f1B(y,1/j ))∩B(x, r),j =1,2, . . .. Since the setsUj⊂Rnare nested, compact, and connected, their intersectionEis also connected. On the other hand, xEf1(y), henceE= {x}. It follows that diam(Uj)→0 asj→ ∞. Let us fixj such thatUjΩ. Note that Uj coincides with thex-component off1B(y,1/j ).

We claim that f is quasilight in Uj; that is, the connected components off1(w)Uj are compact for all w∈Rn. If not, then there existszUj such that thez-component off1(f (z))intersects∂Ujat some pointb. Since f (b)=f (z)B(y,1/j ), there existst >0 such thatf B(b, t )B(y,1/j ). This contradicts the definition ofUj. Therefore,f is quasilight inUj. By [19, Theorem 1.1]f is discrete and open inUj. 2

For the convenience of the reader we state two preliminary results from [21, III.3].

Lemma 3.2. (See [21, III.3.1].) Letf:Ω →Rn be a local homeomorphism and letQbe a simply connected and locally pathwise connected set inRn. SupposeP is a component off1Qsuch thatPΩ. Thenf:P −→hom Q. If in additionQis relatively locally connected, thenf:P −→hom Q.

Lemma 3.3.(See [21, III.3.3].) Letf:Ω→Rn be a local homeomorphism and letA, BΩ be two sets such that f is homeomorphic inAand inB. IfAB= ∅and iff Af Bis connected, thenf is homeomorphic inAB.

Given a sphereS=S(a, r)⊂Rn, and a pointpS, letCS(p, φ)be the open spherical cap ofSwith centerpand opening angleφ(0, π],

CS(p, φ)=

yS: ya, pa> r2cosφ .

For instanceCS(p, π/2)is a hemisphere andCS(p, π )is a punctured sphere. For anyφ(0, π]the capCS(p, φ) contains the pointp.

The following topological lemma forms the main step of the proof of Zorich Global Homeomorphism Theorem, see [21, III.3].

Lemma 3.4.Letf:Ω→Rnbe a local homeomorphism,Ω⊂Rn,n3. Suppose we have the following:

(i) such thatf:G−→hom GwhereGis convex;

(ii) GDΩand there isa∂G∂D;

(iii) a ballB⊂Rnthat containsa=f (a)and such thatS=∂BmeetsGat some pointb.

Let b=f1(b)Gand denote byCS(b, φ)the component off1CS(b, φ)containingb. Then there exists0<

φ0< πsuch thatCS(b, φ0)Dand the closure ofCS(b, φ0)meets∂D.

(5)

Proof. Letφ0be the supremum of allφsuch thatCS(b, φ)D. First we observe thatφ0>0. Indeed, sincef is a local homeomorphism, there exists a neighborhoodVDofbsuch thatf:V −→hom f (V ). Ifφis sufficiently small, thenCS(b, φ)f (V ), henceCS(b, φ)VD. It remains to show thatφ0< π.

Suppose to the contrary that φ0=π. Since CS(b, π )D, it follows from Lemma 3.2 thatf:CS(b, π )−→hom CS(b, π )=S(here the assumptionn3 is used). SinceS:=CS(b, π )is homeomorphic toS, it separatesRninto two components. LetUbe the bounded component ofRn\S. Then the boundary off (U )is contained inSwhich impliesf (U )=B. Moreover,f:U−→hom Bby Lemma 3.2. SincebUGand sincef (U )f (G)=BGis convex (hence connected), Lemma 3.3 yields thatf is homeomorphic inUG.

This leads to a contradiction. SinceUGDit follows thatalies on the boundary ofUG. On the other hand, f (a)=af (U )is an interior point off (UG). 2

We shall use a geometric lemma which is essentially contained in [13].

Lemma 3.5.Suppose we are given a ballB(y0, r)⊂Rn, a pointy1S(y0, r)and a connected setEthat containsy0 and some pointy2S(y0, r). Then there existqB(y0, r)and0< σ <2rsuch that for everyσ < t <4σ/3,

(i) y1B(q, t );

(ii) S(q, t )E= ∅;

(iii) S(q, t )B(y0,2r)\B(y0, r/10).

Proof. Letα be the angle at the point(y0+y1)/2 formed by the line segments fromy0 to(y0+y1)/2 and from (y0+y1)/2 toy2. There are two cases possible.

Case 1.0α < π/2, or, equivalently,|y1y2|> r. In this case we choose q =(y0+y1)/2 and σ =3r/5. For σ < t <4σ/3 we haveB(y0, r/10)B(q, t )andy1B(q, t ). At the same time,y2/B(q, t )because

|y2q|>

√3 2 r= 5

2√ 3σ >4

3σ.

Thus, all conditions (i)–(iii) are satisfied.

Case 2.π/2απ, or, equivalently,|y1y2|r. This time we chooseq =(y1+y2)/2 andσ = |y1y2|/2.

Since|y0q|(

3/2)r, it follows thatB(q, t )B(y0, r/10)= ∅provided that t <

√ 3 2 − 1

10

r.

This is indeed the case, because 4

3σ2 3r <

√ 3 2 − 1

10

r.

All conditions (i)–(iii) are met. 2 4. Proof of Theorem 1.2

Let(Df )1=L. First we observe that the inner distortion off is locally integrable because

KI(x, f )=Df (x)1nJ (x, f )LnDfn for a.e.xΩ. (4.1) We may assume that B4=B(0,4)Ω. It suffices to show that f is discrete and open inB. We will do this by proving that (2.2) holds. Without loss of generality, a in (2.2) equals 0. Fix 1< t <2 and 3< T <4 so that Hn1(fSt) <∞andHn1(fST) <∞. By the area formula we have

(6)

Rn

N (y, f,BT)dy=

BT

J (x, f )dx <∞. Therefore, for almost every 0< R <∞we have

SR

N (y, f,BT)dHn1(y) <∞ and Hn1

f (ST)∩SR

=0. (4.2)

We fix suchR <1/(2L)so that (4.2) holds, and let M:=R1n

SR

N (y, f,BT)dHn1(y).

Our goal is to prove that r1n

Sr

N (y, f,B)dHn1(y)M for a.e. 0< r < R. (4.3)

Letr < Rbe such thatHn1(f (St)∩Sr)=0, and denote byE⊂Sthe set of unit vectorsvfor which

deg(Rv, f,BT) <deg(rv, f,Bt). (4.4)

LetIv:[r, R] →Rnbe the parametrized line segmentIv(s)=sv. By Proposition 3.1, eitherf1(sv)has a nontrivial component for somersR, orf is discrete and open in a neighborhood off1(Iv[r, R]), denoted byUv. By using the co-area formula as in [20, Lemma 2.4], we see that the former possibility only occurs forvF1where Hn1(F1)=0. The mappingf is discrete and open in the open setU:=

{Uv: vE\F1}. It follows from (4.4) and basic properties of path lifting [21, Section II.3] that for eachvE\F1the segmentIvhas a maximalf-liftingIv starting atBtand leavingBT.

Denote

f(x):=lim inf

zx

|f (z)f (x)|

|zx| .

By our assumption on(Df )1there exists a Borel null setFΩsuch thatf(x)1/LforxΩ\F. LetF2be the set ofvE\F1such that eitherIvis unrectifiable orH1(|Iv|∩F ) >0. Since the measure ofFis zero, it follows that the family of curvesΓF := {Iv: vF2}has zero weighted modulus for any locally integrable weight. In particular, MKIΓF =0. SinceΓFU we can apply (2.1) and obtainM{Iv: vF2} =0, which impliesHn1(F2)=0.

ForvE\(F1F2)we have H1

Iv

LH1(Iv) < LR <1

2, (4.5)

which contradicts the fact thatIvbegins atBt and leavesBT. ThusEF1F2. As a consequence,Hn1(E)=0, which means deg(rv, f,Bt)deg(Rv, f,BT)forHn1-a.e.v∈S. Since deg(y, f,Bt)=N (y, f,Bt)for a.e.y∈Rn [8, Proposition 2], inequality (4.3) follows. This completes the proof of Theorem 1.2 via Proposition 2.2.

5. Multiplicity of local homeomorphisms

In 1967 Zorich [24] proved that a local homeomorphismf:Rn→Rn,n3, withKI(·, f )L(Rn)must be a global homeomorphism. Martio, Rickman and Väisälä [16] gave a local version of this result. Namely, iff: 2B→Rn, n3, is a local homeomorphism with bounded distortionKI, then its radius of injectivity in B is bounded from below by a constant depending only onnand ess supKI. As a consequence, the multiplicityN (y, f, B)is bounded byC(n,ess supKI)for ally∈Rn.

The boundedness ofKI can be replaced by the condition exp

λKI1/(n1)

L1(2B),

(7)

but this cannot be relaxed any further [13,17]. Surprisingly, the multiplicity bound remains true under a much weaker condition, namelyKIL1. Example 7.2 below shows that KIqL1 withq <1 does not suffice. The mappings fj(z)=ej zshow that all results discussed here fail whenn=2.

Theorem 5.1.LetfWloc1,n(Ω,Rn),n3, be a local homeomorphism such thatKΩ[f]<∞. IfB is a ball such that4BΩ, thenN (y, f, B)C(n,K4B[f])for ally∈Rn.

Proof. We may assume thatBis the unit ballB. Letx1, . . . , xmf1(y)∩B. Moreover, letrjbe the largest radiusr so that thexj-componentU (xj, r)of f1B(y, r)satisfies U (xj, r)⊂B3. By Lemma 3.2f is a homeomorphism fromU (xj, rj)ontoB(y, rj). We denote bysj the largest radiusssuch thatB(xj, s)U (xj, rj). Thenf B(xj, sj) intersects bothy andS(y, rj). We notice that sincexj∈Band since the ballsB(xj, sj)are pairwise disjoint, there exist at mostN (n)indicesj for whichsj1. Thus we may assume thatB(xj, sj)⊂B2for every 1jm.

We now fix 1jmand a pointajU (xj, rj)∩S3. We apply Lemma 3.5 withB(y0, r)=B(y, rj),y1=f (aj) andE=f (B(xj, sj)), obtaining a pointqj and a numberσj>0. For σj< t <j/3 choosewtB(xj, sj)such thatf (wt)S(qj, t ). We apply Lemma 3.4 withG=U (xj, rj),D=B3,a=aj,B=B(qj, t )andb=f (wt). As a result we obtain 0< φt < π such that the spherical cap Ct :=CS(qj,t )(f (wt), φt)satisfiesCt⊂B3andCt ∩S3

contains some pointct. Consequently, for every pathγjoiningf (wt)andf (ct)inCt, the maximalf-liftingγofγ starting atwt starts fromB2and leavesB3. Following [23, 10.2], we will choose a particular familyΓt of such paths.

Let us say that a circular arc isshortif it is contained in a half-circle. The familyΓt will consist of all short circular arcs that connectf (wt)tof (ct)within Ct. More precisely, lethbe a Möbius transformation that mapsf (wt)to infinity andS(qj, t )\ {f (wt)}toRn1. Observe thath(Ct)is the complement of a ball inRn1. The convexity of Rn1\h(Ct)implies that there exists an(n−2)-hemisphereV such thath(f (ct))+svh(Ct)for everys >0 and vV.

Introduce a family of curvesIv:[0,∞)Ct, defined by Iv(s)=h1

h f (ct)

+s1v ,

and denote byIvthe maximalf-lifting ofIvstarting atwt. Now let 0< (v) <∞be the smallest number such that Iv((v))∈S3. Let

Γt=

Iv|[0,(v)]: vVt .

We writef Γt for the image ofΓtunderf.

There is a lower bound for the spherical modulus off Γt, namely [23, Theorem 10.2]

MS(f Γt)C(n)

t . (5.1)

Let

Γj= {γ: γf Γtfor someσj< t <j/3},

and letΓjbe the family of the corresponding liftsγstarting atwt. Then integrating (5.1) we obtain

MΓj

j/3 σj

C(n)

t dtC(n). (5.2)

As observed earlier, everyγΓjstarts atB2and leavesB3. We denote byEj the smallest closed subset ofB3\B2

that contains|γ| ∩(B3\B2)for allγΓj. Note that Ejf1

B(y,2rj)\B(y, rj/10)

(5.3) by part (iii) of Lemma 3.5. Since the characteristic functionχEj is an admissible function forΓj, we have

(8)

MKIΓj

Ej

KI(x, f )dx. (5.4)

The generalized Poletsky inequalityMΓjMKIΓj[14, Theorem 4.1], together with (5.2) and (5.4) yield mC(n)

m j=1

MΓj m j=1

Ej

KI(x)dx

sup

x∈B3\B2

m j=1

χEj(x)

×

3B

KI(x, f )dx. (5.5)

Claim 1.There existsM=M(n,K4B[f])such that m

j=1

χEj(x)M for everyx∈B3\B2. (5.6)

By virtue of (5.5), Theorem 5.1 follows from Claim 1. In the rest of this section we prove (5.6).

Letx ∈B3\B2be a point covered byM of the setsEj. After relabeling we havexEj for 1j M, and r1r2· · ·rM. Since disjoint sets have disjoint preimages, (5.3) impliesrM20r1.

Chooseτ >0 such thatB(x, τ )⊂B3andf is injective inB(x, τ ). For 1j Mthere existsγjΓjwhich meetsB(x, τ ). Letwj be the starting point ofγj, and letγj be the subcurve ofγjthat begins atwj and ends once it meetsB(x, τ ).

Claim 2.For1jMthere is a curveτj that joinsytof (wj)withinB(y, rj)in such a way that the union of|τj| and |fγj| can be mapped onto a line segment by anL-biLipschitz mappingg:Rn→Rn. HereLis a universal constant.

Proof. Note that the imagefγj is a short circular arc contained in the sphereS(q, t )of Lemma 3.5. Part (iii) of Lemma 3.5 implies

dist

y,|fγj| dist

y, S(q, t )

1

10rj 1

40diam|fγj|. (5.7)

There are two cases. IfyB(q, t ), thenτj is the line segment connectingytof (wj). By virtue of (5.7), the distance fromy toS(q, t )is comparable tot. Therefore, the angle betweenτj and the sphereS(q, t )is bounded from below by a universal constant, and the claim follows.

Suppose that y /B(q, t ). Letρj := |f (wj)y|. Note thatrj/10ρjrj. Letp be the point of the sphere S(y, ρj)that is farthest fromq, namely

p=yρj qy

|qy|.

We chooseτj as the union of the line segment connectingytopand the geodesic arc onS(y, ρj)fromptof (wj).

Once again, the angle betweenτj and the sphereS(q, t )is bounded from below by a universal constant. 2

Letηj, 1jM, be the curve obtained by concatenating(fγj)with−τj, where−indicates the reversal of orientation. Note thatηj begins inf B(x, τ ), proceeds along a circular arc tof (wj), and ends aty. Itsf-liftingηj starting inB(x, τ )is contained inB3and ends atxj.

Claim 3.There exists=(n, M)such that→0asM→ ∞, and

1i<jminM dH

|ηi|,|ηj|

r1/L. (5.8)

Proof. We begin our proof of Claim 3 by observing that |ηj| ⊂B(y,2rM)B(y,40r1). For >0 let Z = {z1, . . . , zN}be an(r1/L)-net inB(y,40r1), whereN=N (, n). The set of all nonempty subsets ofZis an(r1/L)- net in the set of all nonempty closed subsets ofB(y,40r1)equipped with the Hausdorff metric. IfM >2N, then by

(9)

the pigeonhole principle there existi < jsuch that|ηi|and|ηj|are within the distance(r1/L)from the same subset ofZ. Claim 3 follows. 2

Fixi,j, andas in Claim 3, and letg:Rn→Rnbe theL-biLipschitz mapping from Claim 2. By replacingf withgf, which has a comparable distortion functionKI, we may assume that|ηj|is a line segment. Forδ >0 we denote byW (δ)the openδ-neighborhood of|ηj|. LetW(δ)be thexj-component off1W (δ).

Claim 4.Ifδ > r1, thenW(δ)∩S4= ∅.

Proof. Sinceδ > r1, we have|ηi| ⊂W (δ). Suppose to the contrary thatW(δ)⊂B4. ThenW(δ)Ω, which by Lemma 3.2 implies thatf:W(δ)W (δ)is a homeomorphism. This contradicts the fact that thef-liftings of ηi andηj starting inB(x, τ )end at different points, namelyxi andxj. 2

Letδ0 be the supremum of all numbersδ such that W(δ)⊂B4. Since f is a local homeomorphism,δ0>0.

By Lemma 3.2,f:W(δ)−→hom W (δ)for every 0< δ < δ0. By Claim 4 we haveδ0r1.

Choose a pointa∂W0)∩S4. Leta=f (a). Sincea∂W (δ0), there existsp∈ |ηj|such that|ap| =δ0. Forδ0< t <12diam|ηj|choosebt∈ |ηj| ∩S(p, t ). We apply Lemma 3.4 withG=W0),D=B4,a=a,B= B(p, t ) and b=bt. As a result we obtain 0< φt < π such that the spherical cap Ct :=CS(p,t )(bt, φt)satisfies Ct⊂B4andCt ∩S4contains some pointct. Consequently, for every pathγjoiningbtandf (ct)inCt, the maximal f-liftingγofγ starting atf1(bt)∩ |ηj|starts fromB3and leavesB4. LetΓ be the family of all such pathsγ and letΓbe the family of the liftsγ. From [23, Theorem 10.2] we have

MΓ C(n)

diam(η j)/2 r1

dt

t C(n)logdiam(ηj) 2r1

.

By (5.7) we have diam|ηj|cr1with a universal constantc >0. Therefore, MΓ C(n)log1

. (5.9)

On the other hand, since the characteristic functionχB4\B3 is an admissible function forΓj, we obtain MKIΓ

B4\B3

KI(x, f )dx.

Combining this with (5.9) and using the Poletsky inequality again, we have C(n,K4B[f]), hence M C(n,K4B[f]). This gives (5.6). The proof of Theorem 5.1 is complete. 2

6. Proof of Theorem 1.1

Denotefλ(x)=f (x)+λx,λ >0. ThenfλWloc1,n(Ω,Rn). Moreover, by [14, Lemma 10], KI

x, fλ

C(δ, n)KI(x, f ) and Dfλ1

(x)C(δ, λ) (6.1)

for almost every xΩ. Thus fλ is discrete and open for every λ >0 by Theorem 1.2. Furthermore, by [14, Lemma 13] fλ is a local homeomorphism. (Although [14, Lemma 13] imposes a stronger condition on the dis- tortion off, this condition is only used to ensure thatf is discrete and open.) Sincefλf locally uniformly, the following proposition implies thatf is a local homeomorphism, completing the proof of Theorem 1.1.

Proposition 6.1.Suppose that a mappingfWloc1,n(Ω,Rn)withKΩ[f]<can be uniformly approximated by local homeomorphismsfjWloc1,n(Ω,Rn)such thatsupjKΩ[fj]<∞. Thenf is a local homeomorphism.

(10)

Proof. By [14, Proposition 7] it suffices to show thatf is discrete and open. If n=2, this is due to Iwaniec and Šverák [10]. Thus we assume thatn3. LetB=B(x0, R)be a ball such that 8BΩ. We will show that

N (y, f, B)C for a.e.y∈Rn, (6.2)

whereC <∞does not depend ony. Proposition 2.2 will then imply thatf is discrete and open inB. Applying Theorem 5.1 tofj, we obtain

N (y, fj,2B)C for everyy∈Rn, whereCdepends only on supjKΩ[fj]andn.

We fix R < t <2R so thatHn1(f S(x0, t )) <∞, and a pointyf B\f S(x0, t ). Letd=dist(y, f S(x0, t )).

Sincefjf locally uniformly, there existsj0such that|fj(x)f (x)|< d/2 for allj j0and allxS(x0, t ).

Consequently, the restrictions offj andf toS(x0, t )are homotopic via the straight-line homotopy that takes values inRn\ {y}. It follows that

deg

y, f, B(x0, t )

=deg

y, fj, B(x0, t )

N (y, fj,2B)C

for alljj0. SinceN (y, f, B)N (y, f, B(x0, t ))=deg(y, f, B(x0, t ))for almost everyy∈Rn, we conclude that (6.2) indeed holds. The proof is complete. 2

7. Concluding remarks

Corollary 7.1. Suppose that fWloc1,n(Rn,Rn)is a nonconstant mapping such thatKI(·, f )L1loc(Rn). If there existsδ >−1such thatDf (x)M(δ)for almost everyx∈Rn, thenf is a homeomorphism.

Proof. As in the proof of Theorem 1.1 we have thatfλ(x)=f (x)+λx is a local homeomorphism for allλ >0.

Since(Dfλ)1L(Rn), it follows from [14, Lemma 12] that lim inf

xa

|fλ(x)fλ(a)|

|xa| λ

2 >0 (7.1)

for alla∈Rn. By a theorem of John [11, p. 87],fλis a homeomorphism. Sincef is discrete and open by Theorem 1.1, we can apply [14, Proposition 7] and conclude thatf is a homeomorphism. 2

Sharpness of Theorem 5.1 is demonstrated by the following example which combines the ideas from [2] and [13].

Example 7.2.For anyq <1 there exists a sequence of mappingsfjW1,3(B,R3)such that sup

j

B

KIq(x, fj)dx <∞ and N

0, fj, B(0,1/4)

→ ∞.

Proof. By a version of Zorich’s construction (see [9,21]) there exists a mapping φW1,3(R3,R3) such that KI(·, φ)L(R3),φ is a local homeomorphism outside of R×(2Z+1)2, and φ is 4-periodic in the last two variables. Therefore, it suffices for us to construct biLipschitz homeomorphismsfj:B→R3such that

(i) supj

BKIq(x, fj)dx <∞;

(ii) fj(B)⊂D×R(hereD⊂R2is the unit disc);

(iii) fj(B1/4)contains a line segment{0} × [−L, L] ⊂R2×RwhereL→ ∞asj→ ∞.

The compositionsφfjwill be mappings with large multiplicity.

Fory∈R3lets(y)=

y12+y22. Forα >2 we define a mappingx=g(y)by xi=s(y)α1yi, i=1,2;

x3=s(y)y3.

(11)

Sinces(x)=s(y)α, the inverse mappingy=f (x)outside the set{s(x)=0}is given by yi=s(x)1/α1xi, i=1,2;

y3=s(x)1/αx3, s(x)=0.

Let Ω= {x∈R3: s(x) <1, |x3|<1}and Ω=f (Ω). We restrict our attention to yΩ, where in particular s(y) <1. Elementary computations show that

Dg(y)Cmax

s(y),|y3| and J (y, g)Cs(y)2(α1)+1.

Therefore, Dg(y)3

J (y, g) Cs(y)2(1α)1max

s(y)3,|y3|3

. (7.2)

Since

Dg(y)3

J (y, g) =KI(x, f ),

inequality (7.2) can be used to estimateKI(x, f )as follows.

KI(x, f )Cs(x)(2(1α)1)/αmax

s(x)3/α, s(x)3/α|x3|3

Cs(x)(2α+2)/α, where at the last step we used|x3|<1. We achieve

ΩKI(x, f )qdx <∞by choosingαlarge enough so that 2α+2

α q <2.

The mappingf constructed thus far is not inW1,3, and is not even continuous. However, this can be corrected by replacings(y)withsj(y)=

y12+y22+1/j2. The mappingx=gj(y)given by xi=sj(y)α1yi, i=1,2;

x3=sj(y)y3,

is biLipschitz; we denote the inverse byfj. The computation ofDgjandJ (·, gj)goes through exactly as before and shows that the integral ofKIq(·, fj)is bounded independently ofj. Sincegj(0,0, y3)=(0,0, y3/j ), we have fj(0,0, x3)=(0,0, j x3). Thus, this mappingfj fulfills the requirements (i)–(iii). 2

Acknowledgements

The research was partially carried out during Rajala’s visit to Syracuse University. He wishes to thank the De- partment of Mathematics for its hospitality. Part of this research was done when Kovalev and Onninen were visiting Princeton University. Also, the authors are thankful to the referee for careful reading of the manuscript and several valuable comments.

References

[1] K. Astala, T. Iwaniec, G.J. Martin, J. Onninen, Extremal mappings of finite distortion, Proc. London Math. Soc. (3) 91 (3) (2005) 655–702.

[2] J.M. Ball, Global invertibility of Sobolev functions and the interpenetration of matter, Proc. Roy. Soc. Edinburgh Sect. A 88 (3–4) (1981) 315–328.

[3] V.M. Gol’dshtein, The behavior of mappings with bounded distortion when the distortion coefficient is close to one, Sibirsk. Mat. Zh. 12 (1971) 1250–1258.

[4] V.M. Gol’dshtein, S.K. Vodopyanov, Quasiconformal mappings, and spaces of functions with first generalized derivatives, Sibirsk. Mat.

Zh. 17 (3) (1976) 515–531.

[5] J. Heinonen, T. Kilpeläinen, BLD-mappings inW2,2are locally invertible, Math. Ann. 318 (2) (2000) 391–396.

[6] J. Heinonen, P. Koskela, Sobolev mappings with integrable dilatations, Arch. Ration. Mech. Anal. 125 (1) (1993) 81–97.

(12)

[7] S. Hencl, P. Koskela, Mappings of finite distortion: Discreteness and openness for quasi-light mappings, Ann. Inst. H. Poincaré Anal. Non Linéaire 22 (3) (2005) 331–342.

[8] S. Hencl, J. Malý, Mappings of finite distortion: Hausdorff measure of zero sets, Math. Ann. 324 (3) (2002) 451–464.

[9] T. Iwaniec, G. Martin, Geometric Function Theory and Non-Linear Analysis, Oxford Univ. Press, New York, 2001.

[10] T. Iwaniec, V. Šverák, On mappings with integrable dilatation, Proc. Amer. Math. Soc. 118 (1) (1993) 181–188.

[11] F. John, On quasi-isometric mappings. I, Comm. Pure Appl. Math. 21 (1968) 77–110.

[12] P. Koskela, J. Onninen, Mappings of finite distortion: Capacity and modulus inequalities, J. Reine Angew. Math. 599 (2006) 1–26.

[13] P. Koskela, J. Onninen, K. Rajala, Mappings of finite distortion: Injectivity radius of a local homeomorphism, in: Future Trends in Geometric Function Theory, in: Rep. Univ. Jyväskylä Dep. Math. Stat., vol. 92, Univ. Jyväskylä, Jyväskylä, 2003, pp. 169–174.

[14] L.V. Kovalev, J. Onninen, On invertibility of Sobolev mappings, preprint, 2008, arXiv:0812.2350.

[15] J.J. Manfredi, E. Villamor, An extension of Reshetnyak’s theorem, Indiana Univ. Math. J. 47 (3) (1998) 1131–1145.

[16] O. Martio, S. Rickman, J. Väisälä, Topological and metric properties of quasiregular mappings, Ann. Acad. Sci. Fenn. Ser. A I 488 (1971).

[17] J. Onninen, Mappings of finite distortion: Minors of the differential matrix, Calc. Var. Partial Differential Equations 21 (4) (2004) 335–348.

[18] K. Rajala, The local homeomorphism property of spatial quasiregular mappings with distortion close to one, Geom. Funct. Anal. 15 (5) (2005) 1100–1127.

[19] K. Rajala, Reshetnyak’s theorem and the inner distortion, Pure Appl. Math. Q., in press, University of Jyväskylä preprint, No. 336, 2007.

[20] K. Rajala, Remarks on the Iwaniec–Šverák conjecture, University of Jyväskylä preprint, No. 377, 2009.

[21] S. Rickman, Quasiregular Mappings, Springer-Verlag, Berlin, 1993.

[22] Q. Tang, Almost-everywhere injectivity in nonlinear elasticity, Proc. Roy. Soc. Edinburgh Sect. A 109 (1–2) (1988) 79–95.

[23] J. Väisälä, Lectures onn-Dimensional Quasiconformal Mappings, Lecture Notes in Math., vol. 229, Springer-Verlag, Berlin, 1971.

[24] V.A. Zorich, M.A. Lavrentyev’s theorem on quasiconformal space maps, Mat. Sb. (N.S.) 74 (116) (1967) 417–433.

Références

Documents relatifs

De ce fait, on pourra plus tard, à la fin (problématique) de la présente guerre, faire un tableau précis de Ia tenue de nos industries pendant la durée du conflit, période

This foliation plus Sing(&lt;o) will be called the singular foliation of co. A natural problem to consider is the search for a significant class of integrable i-forms co for which

In contrast with the construction in Theorem 1.2 from [1] we do not put some basic function into each subcube that intersects our set but only into some of them. In the proof it

Then, also the restriction of σ to almost every straight line parallel to the coordinate axes is absolutely continuous, since it is the composition of two absolutely

We extend to quasi- conformal mappings in the unit ball of R&#34; a theorem of Pommcrenke concerning conformal mappings in the unit disk of the

Toute utilisation commerciale ou impression systématique est constitutive d’une infrac- tion pénale.. Toute copie ou impression de ce fichier doit contenir la présente mention

Toute utilisation commerciale ou impression systématique est constitutive d’une infrac- tion pénale.. Toute copie ou impression de ce fichier doit contenir la présente mention

Ouahab, Existence results for frac- tional order functional differential inclusions with infinite delay and applications to con- trol theory.. Colombo, Extensions and selections of