• Aucun résultat trouvé

Maximal metric surfaces and the Sobolev-to-Lipschitz property

N/A
N/A
Protected

Academic year: 2021

Partager "Maximal metric surfaces and the Sobolev-to-Lipschitz property"

Copied!
34
0
0

Texte intégral

(1)

https://doi.org/10.1007/s00526-020-01843-0

Calculus of Variations

Maximal metric surfaces and the Sobolev-to-Lipschitz

property

Paul Creutz1· Elefterios Soultanis2

Received: 7 February 2020 / Accepted: 31 July 2020 / Published online: 20 September 2020 © The Author(s) 2020

Abstract

We find maximal representatives within equivalence classes of metric spheres. For Ahlfors regular spheres these are uniquely characterized by satisfying the seemingly unrelated notions of Sobolev-to-Lipschitz property, or volume rigidity. We also apply our construction to solu-tions of the Plateau problem in metric spaces and obtain a variant of the associated intrinsic disc studied by Lytchak–Wenger, which satisfies a related maximality condition.

Mathematics Subject Classification 30L10· 49Q05 · 53B40

1 Introduction

1.1 Main result

A celebrated result due to Bonk and Kleiner [3] states that an Ahlfors 2-regular metric sphere Z is quasisymetrically equivalent to the standard sphere if and only if it is linearly locally connected. Recently, it was shown in [35] that the quasisymmetric homeomorphism

uZ : S2→ Z may be chosen to be of minimal energy E+2(uZ), and in this case is unique up to a conformal diffeomorphism of S2.

The map uZgives rise to a measurable Finsler structure on S2, defined by the approximate metric differential apmd uZ; cf. [33] and Sect.6below. When Z is a smooth Finsler surface, the approximate metric differential carries all the metric information of Z . In the present

Communicated by J. Jost.

The first author was partially supported by the DFG Grant SPP 2026. The second author was partially supported by the Vilho, Yrjö ja Kalle Väisalä Foundation (postdoc pool) and by the Swiss National Science Foundation Grant 182423.

B

Elefterios Soultanis

elefterios.soultanis@gmail.com Paul Creutz

pcreutz@math.uni-koeln.de

1 University of Cologne, Weyertal 86-90, 50931 Cologne, Germany

(2)

generality, however, apmd uZis defined only almost everywhere and thus does not determine the length of every curve.

Definition 1.1 Let Y and Z be Ahlfors 2-regular, linearly locally connected metric spheres. We say that Y and Z are analytically equivalent if there exist energy minimizing parametriza-tions uY and uZ such that

apmd uY = apmd uZ (1.1)

almost everywhere.

By the aformentioned uniqueness result, analytic equivalence defines an equivalence rela-tion on the class of linearly locally connected, Ahlfors 2-regular spheres. The main result of this paper states that the equivalence class of such a sphere contains a maximal representative, unique up to an isometry.

Theorem 1.2 Let Z be a linearly locally connected, Ahlfors 2-regular sphere. Then there is

linearly locally connected, Ahlfors 2-regular sphere Z which is analytically and bi-Lipschitz equivalent to Z and satisfies the following properties.

(1) Sobolev-to-Lipschitz property: If f ∈ N1,2(Z) has weak upper gradient 1, then f has a 1-Lipschitz representative.

(2) Thick geodecity: For arbitrary measurable subsets E, F ⊂ Z of positive measure and C > 1, one has Mod2(E, F; C) > 0.

(3) Maximality: If Y is analytically equivalent to Z , there exists a 1-Lipschitz homeomor-phism f : Z→ Y .

(4) Volume rigidity: If Y is a linearly locally connected, Ahlfors 2-regular sphere, and f :

Y → Z is a 1-Lipschitz area preserving map which is moreover cell-like, then f is an isometry.

Moreover Z is characterized uniquely, up to isometry, by any of the listed properties.

In the setting of Ahlfors regular metric spheres, Theorem 1.2links the essential metric investigated in [1,7,8], the concept of volume rigidity studied under curvature bounds in [27– 29], the Sobolev-to-Lipschitz property arising in the context of RCD spaces in [15,16], and the notion of thick quasiconvexity whose connection to Poincaré inequalities is investigated in [10–12].

More generally, an Ahlfors 2-regular disc Z with finite boundary length admits a quasisym-metric parametrization by the standard discD if and only if it is linearly locally connected. The parametrization may again be chosen to be of minimal Reshetynak energy and is then unique up to a conformal diffeomorphism, see [35]. A suitable variant of Theorem1.2also holds in this setting. In this case in the formulation of volume rigidity one has to assume that

f preserves the boundary curves and the restriction f : ∂Y → ∂ Z is monotone, instead of

assuming that the entire map f is cell-like.

Other parametrization results for metric surfaces have been obtained for example in [23, 36,39,40,43]. These results however do not include a uniqueness part. Thus it seems unclear how to define a suitable notion of analytic equivalence in the settings therein.

1.2 Construction of essential metrics

Our construction of essential metrics relies on the p-essential infimum over curve families and extends the construction of the essential metric in [1] to the case p< ∞; cf. (1.3) and

(3)

Definition5.1. For a metric measure space X , p ∈ [1, ∞], and a family of curves  in X, we define its p-essential length essp() ∈ [0, ∞] as the essential infimum of the length function on with respect to p-modulus. The p-essential distance dp : X × X → [0, ∞] is defined by

dp(x, y) := lim

δ→0essp(B(x, δ), B(y, δ)), x, y ∈ X, (1.2) where(E, F) denotes the family of curves in X joining two given measurable subsets

E, F ⊂ X. This quantity does not automatically satisfy the triangle inequality, and we

con-sider instead the largest metric dp ≤ dp; see the discussion after Definition5.3. In general,

dp might take infinite values, and its finiteness is related to an abundance of curves of uni-formly bounded length connecting given disjoint sets. The condition of thick quasiconvexity, introduced in [12], quantifies the existence of an abundance of quasiconvex curves. Definition 1.3 Let(X, d, μ) be a metric measure space and p ∈ [1, ∞]. We say that X is p-thick quasiconvex with constant C≥ 1 if, for all measurable subsets E, F ⊂ X of positive measure, we have

Modp(E, F; C) > 0.

We say that X is p-thick geodesic if X is p-thick quasiconvex with constant C for every

C> 1.

Here(E, F; C) denotes the family of curves γ : [0, 1] → X joining E and F such that

(γ ) ≤ C · d(γ (0), γ (1)). Note that this is equivalent to the original definition in [12] when

X is infinitesimally doubling.

Theorem 1.4 Let(X, d, μ) be an infinitesimally doubling metric measure space which is

p-thick quasiconvex with constant C and 1≤ p ≤ ∞. Then there exists a metric dp on X ,

for which Xp:= (X, dp, μ) is p-thick geodesic and d ≤ dp≤ Cd. Moreover dpis minimal

among metricsρ ≥ d for which Xρ := (X, ρ, μ) is p-thick geodesic.

It is known that doubling p-Poincaré spaces are p-thick quasiconvex for some constant depending only on the data of X , though the converse need not be true unless p = ∞; see [11,12]. For p < q, the assumption of p-thick geodecity is strictly stronger than that of q-thick geodecity, see Example3.3below. However, under the assumption of a suitable Poincaré inequality the particular value of q is immaterial. Indeed, for Poincaré spaces the essential metrics of all indices coincide with the essential metric dessintroduced in [1]. The essential metric dessis given by

dess(x, y) := ess∞({x}, {y}), x, y ∈ X. (1.3) Proposition 1.5 Let X be a doubling metric measure space satisfying a p-Poincaré inequality.

Then dq = dessfor every q∈ [p, ∞].

In the proof of Theorem1.2we will set Z:= (Z, d2). A posteriori, the chosen metric agrees

with the essential metric (1.3). However from this characterization it is unclear that the disc obtained in Theorem1.2is unique. This follows from the use of 2-modulus in the construction of the metric, and its quasi-invariance under quasisymmetric maps.

(4)

1.3 The Sobolev-to-Lipschitz property, thick geodecity, and volume rigidity

The Sobolev-to-Lipschitz property and thick geodecity are defined and equivalent in a much broader framework.

Definition 1.6 A metric measure space X is said to have the p-Sobolev-to-Lipschitz property, if every f ∈ N1,p(X) for which the minimal p-weak upper gradient gf satisfies gf ≤ 1 almost everywhere has a 1-Lipschitz representative.

To illustrate this equivalence, and contrast the difference of the Sobolev-to-Lipschitz property with merely being geodesic, recall that a proper metric space is geodesic if and only if, whenever a function f has (genuine) upper gradient 1, it is 1-Lipschitz. The p-Sobolev-to-Lipschitz property requires we can reach the same conclusion from the weaker assumption that f has p-weak upper gradient 1. The next result translates this condition to having an abundance of nearly geodesic curves in the space.

Theorem 1.7 Let X be an infinitesimally doubling metric measure space and p ≥ 1. Then

X has the p-Sobolev-to-Lipschitz property if and only if it is p-thick geodesic.

These properties are also related to volume rigidity, which asks whether any volume preserving 1-Lipschitz map of a given class is an isometry. Recall that a Borel map

f : (X, μ) → (Y , ν) between metric measure spaces is volume preserving if f∗μ = ν.

When X and Y are n-rectifiable, we equip them with the Hausdorff measureHn, unless otherwise explicitly stated. Volume preserving 1-Lipschitz maps between rectifiable spaces are essentially length preserving, see Proposition4.1, but need not be homeomorphisms. For volume preserving 1-Lipschitz homeomorphisms with quasiconformal inverse, the Sobolev-to-Lipschitz property guarantees isometry.

Proposition 1.8 Let f : X → Y be a volume preserving 1-Lipschitz homeomorphism

between n-rectifiable metric spaces. If f−1 is quasiconformal and Y has the n-Sobolev-to-Lipschitz property, then f is an isometry.

Note that the inverse of a quasiconformal map is not quasiconformal in general; cf. [44, Remark 4.2]. This is guaranteed under the assumption that the spaces are Ahlfors n-regular and X supports an n-Poincaré inequality. The assumption that Y has the Sobolev-to-Lipschitz property is essential for the validity of Proposition1.8; cf. Example4.4. Thus, Proposition 1.8connects volume rigidity, the Sobolev-to-Lipschitz property and thick geodecity under these fairly restrictive assumptions, and will be used in the proof of Theorem1.2.

1.4 Application: essential minimal surfaces and essential pull-back metrics

For the definitions and terminology in the forthcoming discussion, we refer the reader to Sect.2.

A smooth map u : M → X between Riemannian manifolds gives rise to a pull-back metric on M. In the case that M and X are merely metric spaces and u a continuous map, various nonequivalent definitions of pull-back metric are discussed in [37,38]. The metric

du(p, q) := infγ (u ◦ γ ), p, q ∈ M,

whereγ varies over all continuous curves in M joining p to q, was used in [33,38,42] to study solutions u: D → X of the Plateau problem in X, also referred to as minimal discs. If

(5)

X satisfies a quadratic isoperimetric inequality and u bounds a Jordan curve of finite length,

then dugives rise to a factorization

D X

Zu u

Pu ¯u

where:

(1) Zuis a geodesic disc which satisfies the same quadratic isoperimetric inequality as X . (2) Puis a minimal disc bounding∂ Zu and a uniform limit of homeomorphisms. (3) ¯u is 1-Lipschitz and (Pu◦ γ ) = (u ◦ γ ) for every curve γ in D.

The fact that Zu satisfies the same quadratic isoperimetric inequality as X reflects the fact that the intrinsic curvature of a classical minimal surface is bounded above by that of the ambient manifold; cf. [34].

Our definition of essential metrics suggests the following variant of pull-back metrics in the minimal disc setting. Namely for a family of curves in D let essu() be the essential infimum of(u ◦ γ ) with respect to 2-modulus; cf. Definition5.1. Then for p, q ∈ D we define the essential pull-back metric by



du(p, q) := lim

δ→0essu(Bδ(p), Bδ(q)). This construction gives rise to the following factorization result.

Theorem 1.9 Let X be a proper metric space satisfying a(C, l0)-quadratic isoperimetric

inequality. Let u: D → X be a minimal disc bounding a Jordan curve  which satisfies a chord-arc condition. Then there exists a geodesic disc Zuand a factorization

D X  Zu u  Pu u such that

(1) Zuhas the Sobolev-to-Lipschitz property and satisfies a(C, l0)-quadratic isoperimetric

inequality,

(2) Pua minimal disc bounding∂Zu and a uniform limit of homeomorphisms, and (3) u : Zu → X is 1-Lipschitz and for 2-almost every curve γ in D one has (Pu◦ γ ) =

(u ◦ γ ).

Furthermore the factorization is maximal in the following sense: If Z is a metric disc and u = ˜u ◦ P, where P : D → Z and ˜u : Z → X satisfy (2) and (3), then there exists a surjective 1-Lipschitz map f : Zu→ Z such that P= f ◦ Pu.

A Jordan curve is said to satisfy a chord-arc condition if it is bi-Lipschitz equivalent to

S1. The fact that the map Pu is a minimal disc implies additional regularity. Indeed, Pu is infinitesimally quasiconformal, Hölder continuous and its minimal weak upper gradient has global higher integrability; cf. [31–33,35].

Compared to the construction in [33], we trade off some regularity of u in exchange for the Sobolev-to-Lipschitz property on Zu. Indeed,u only preserves the length of almost every curve, while Zuis nicer in analytic and geometric terms. Note however that some geometric

(6)

properties of Zu, established for Zu in [33,42], remain open for Zu. In particular we do not know whetherH1(∂Z

u) < ∞.

Note that, in contrast to Theorem1.2, no characterization in terms of the Sobolev-to-Lipschitz property in Theorem1.9is possible. In fact it may happen that Zualready has the Sobolev-to-Lipschitz property but still Zuand Zuare not isometric, see Example6.4.

More generally, when M supports a Q-Poincaré inequality where Q≥ 1 satisfies (2.2), maps u∈ Nloc1,p(M; X), with a priori higher exponent p > Q, give rise to a p-essential pull-back metric on M and the resulting metric measure space has the Q-Sobolev-to-Lipschitz

property; cf. Definition1.6and Theorem5.16. Higher integrability of minimal weak upper gradients for quasiconformal maps is known to hold when M and X both have Q-bounded geometry, and also in the situation of [31, Theorem 1.4], but not in general; cf. [20,21, 26]. Together with the Poincaré inequality and Morrey’s embedding, higher integrability of minimal weak upper gradients implies finiteness of the essential pull-back metric. The assumption in Theorem1.9that is not only of finite length but satisfies a chord-arc condition is needed to guarantee this additional regularity.

1.5 Organization

The paper is organized as follows. After the preliminaries in Sect.2, we treat the equivalence of thick geodecity and the Sobolev-to-Lipschitz property in Sect.3, where we prove Theorem 1.7. Volume rigidity is discussed in Sect.4, which includes the proof of Proposition1.8.

The construction of essential metrics is presented in Sect.5. We develop some general tools in Sect.5.1, and prove Theorem1.4and Proposition1.5in Sect.5.2. In Sect.5.3we apply our constuction to Sobolev maps from a Poincaré space, and derive some basic properties of the resulting metric measure space; cf. Theorem5.15.

Finally, in Sect.6, we apply the results from Sects.5.2and5.3to obtain Theorems1.2 and1.9. First in Sect.6.1we prove Theorem1.2. Then in Sect.6.2we prove a weaker but more general result for discs, cf. Theorem6.3, and finally in Sect.6.3we discuss Theorem1.9.

2 Preliminaries

We refer to the monographs [2,22] for the material discussed below. Given a metric space

(X, d), balls in X are denoted B(x, r) and, if B = B(x, r) ⊂ X is an open ball and σ > 0, σ B denotes the ball with the same center as B and radius σr. The Hausdorff measure on X

is denotedHn or, if we want to stress the space or metric, byHnXorHnd. The normalizing constant is chosen so thatHnRn agrees with the Lebesgue measure.

By a measureμ on a metric space X we mean an outer measure which is Borel regular. A triple(X, d, μ) where (X, d) is a proper metric space and μ is a measure on X which is non-trivial on balls, is called a metric measure space. Note that all metric measure spaces in our convention are complete, separable and every closed bounded subset is compact. Furthermore all balls have positive finite measure and in particular the measures are Radon, cf. [22, Corollary 3.3.47]. We often abbreviate X = (X, d, μ) when the metric and measure are clear from the context.

Given a curveγ : [a, b] → X, we denote by (γ ) its length. When γ is absolutely continuous we writet| for the metric speed (at t ∈ [a, b]) of γ . Moreover, when γ is rectifiable (i.e.(γ ) < ∞) we denote by ¯γ : [a, b] → X the constant speed parametrization

(7)

ofγ . This is the (Lipschitz) curve satisfying

γ (a + (b − a)(γ |[a,t])/(γ )) = ¯γ(t), t ∈ [a, b].

The arc length parametrizationγsofγ is the reparametrization of ¯γ to the interval [0, (γ )]. With the exception of Proposition5.7, we assume curves are defined on the interval[0, 1].

Given a metric measure space X , a Banach space V , and p≥ 1, we denote by Lp(X; V ) and Llocp (X; V ) the a.e.-equivalence classes of p-integrable and locally p-integrable μ-measurable maps u: X → V . We also denote Lp(X) := Lp(X; R) and Llocp (X) :=

Llocp (X; R) and abuse notation by writing f ∈ Lp(X; V ) or f ∈ Llocp (X; V ) for maps f (instead of equivalence classes).

Properties of measures

A measureμ on X is called doubling if there exists C ≥ 1 such that

μ(B(x, 2r)) ≤ Cμ(B(x, r)) (2.1)

for every x ∈ X and 0 < r < diam X. The least such constant is denoted Cμand called the doubling constant ofμ. Doubling measures satisfy a relative volume lower bound

C  diam B diam B Qμ(B)

μ(B) for all balls B⊂ B⊂ X, (2.2)

for some constants C> 0 and Q ≤ log2Cμdepending only on Cμ. The opposite inequality with the same exponent Q need not hold. If there are constants C, Q > 0 such that

1

Cr

Q≤ μ(B(x, r)) ≤ CrQ, 0 < r < diam X,

we say thatμ is Ahlfors Q-regular. Ifμ is doubling and

Mrf(x) = sup

0<s<r− 

B(x,s)| f |dμ

denotes the (restricted) centered maximal function of f at x∈ X the sublinear operatorMr satisfies the usual boundedness estimates

Mrf Lp(X)≤ C f Lp(X)(p > 1), μ({Mrf > λ}) ≤ C f L

1(X)

λ , λ > 0, (2.3)

for some constant C depending only on Cμ and p. Consequently almost every point of a

μ-measurable set E ⊂ X is a Lebesgue density point.

If the measureμ satisfies the infinitesimal doubling condition lim sup

r→0

μ(B(x, 2r)) μ(B(x, r)) < ∞

forμ-almost every x ∈ X, the claim about density points still remains true.

Proposition 2.1 [22, Theorem 3.4.3] If(X, d, μ) is infinitesimally doubling metric measure

space and f ∈ L1 loc(X) then f(x) = lim r→0−  B(x,r) f dμ

(8)

forμ-almost every x ∈ X. In particular, μ-almost every point of a Borel set E ⊂ X is a Lebesgue density point.

Sobolev spaces

Let(X, d, μ) be a metric measure space, Y a metric space, and p ≥ 1. If u : X → Y is a map and g: X → [0, ∞] is Borel, we say that g is an upper gradient of u if

dY(u(γ (b), γ (a)) ≤ 

γ g (2.4)

for every curveγ : [a, b] → X. Recall that the line integral of g over γ is defined as 

γ g:=  (γ )

0

g(γs(t))dt

ifγ is rectifiable and ∞ otherwise. Suppose Y = V is a separable Banach space. A μ-measurable map u∈ Lp(X; V ) is called p-Newtonian if it has an upper gradient g ∈ Lp(X). The Newtonian seminorm is

u 1,p=  u p Lp(X;V )+ inf g g p Lp(X) 1/p , (2.5)

where the infimum is taken over all upper gradients g of u. The Newtonian space N1,p(X; V )

is the vector space of equivalence classes of p-Newtonian maps, where two maps u, v : X →

V are equivalent if u − v 1,p = 0. The quantity (2.5) defines a norm on N1,p(X; V ) and

(N1,p(X; V ), · 1,p) is a Banach space.

We say that aμ-measurable map u : X → V is locally p-Newtonian if every point x ∈ X has a neighbourhood U such that

u|U ∈ N1,p(U; V ),

and denote the vector space of locally p-Newtonian maps by Nloc1,p(X; V ). Minimal upper gradients

Let be a family of curves in X. We define the p-modulus of  as Modp() := inf  X ρpdμ : ρ : X → [0, ∞] Borel,  γρ ≥ 1 for every γ ∈   .

We say that a Borel function g: X → [0, ∞] is a p-weak upper gradient of u if there is a path family0with Modp(0) = 0 so that (2.4) holds for all curvesγ /∈ 0. The infimum in (2.5) need not be attained by upper gradients of u but there is a minimal p-weak upper gradient gu of u so that

gu Lp(X)= inf

g g Lp(X).

The minimal p-weak upper gradient is unique up to sets of measure zero.

If Y is a complete separable metric space then there is an isometric embeddingι : Y → V into a separable Banach space. We may define

(9)

We remark that maps u∈ Nloc1,p(X; Y ) have minimal p-weak upper gradients that are unique up to equality almost everywhere and do not depend on the embedding.

Poincare inequalities

A metric measure space X= (X, d, μ) supports a p-Poincaré inequality if there are constants

C, σ ≥ 1 so that −  B |u − uB|dμ ≤ C diam B  −  σ Bg pdμ 1/p (2.6) whenever u ∈ L1loc(X), g is an upper gradient of u, and B is a ball in X. Doubling metric measure spaces supporting a Poincaré inequality enjoy a rich theory.

Theorem 2.2 (Morrey embedding) Let(X, d, μ) be a complete doubling metric measure

space, where the measure satisfies (2.2) for Q ≥ 1, and suppose X supports a Q-Poincaré inequality. If p> Q, there is a constant C ≥ 1 depending only on the data of X so that for any u∈ Nloc1,p(X; Y ) and any ball B ⊂ X we have

dY(u(x), u(y)) ≤ C(diam B)Q/pd(x, y)1−Q/p  −  σ Bg p udμ 1/p , x, y ∈ B.

Here the data of X refers to p, the doubling constant of the measure and the constants in the Poincaré inequality.

Rectifiability and metric differentials

A separable metric space X is said to be n-rectifiable if there exist Borel sets Ei ⊂ Rnand Lipschitz maps gi : Ei→ X for each i ∈ N, satisfying

Hn X\ i∈N gi(Ei) = 0.

Unless otherwise stated we always equip n-rectifiable spaces with the Hausdorff n-measure

Hn.

Metric differentials play an important role in the study of rectifiable metric spaces. Let

E ⊂ Rn be measurable and f : E → X be a Lipschitz map into a metric space X. By a fundamental result of Kirchheim [25], f admits a metric differential mdx f at almost every point x∈ E. The metric differential is a seminorm on Rnand such that

| mdx f(y − z) − d( f (y), f (z))| = o(d(x, y) + d(x, z)) whenever y, z ∈ E. If ⊂ Rn is a domain and p≥ 1, a Sobolev map u ∈ Nloc1,p( ; X) admits an approximate

metric differential apmdxu for almost every x∈ ; cf. [24] and [31]. If ⊂ M is a domain in a Riemannian n-manifold, the approximate metric differential of u∈ N1,p( ; X) may be

defined, for almost every x∈ , as a seminorm on TxM by

apmdxu:= apmdψ(x)(u ◦ ψ−1) ◦ dψ(x) (2.7) for any chartψ around x; see the discussion after [14, Definition 2.1] and the references therein for the details.

(10)

We formulate the next results for maps defined on domains of manifolds; that is, we assume that ⊂ M for a Riemannian n-manifold M. Note that, although in the references the statements are proved in the Euclidean setting, the definition (2.7) easily implies the corresponding statements for general Riemannian manifolds. For the following area formula for Sobolev maps with Lusin’s property (N ), recall that a Borel map f : (X, μ) → (Y , ν) is said to have Lusin’s property (N ) ifν( f (E)) = 0 whenever μ(E) = 0.

Theorem 2.3 [24, Theorems 2.4 and 3.2] Let u∈ Nloc1,p( ; X) for some p > n. Then u has

Lusin’s property (N ), and

 ϕ(x)J(apmdxu)dx =  X ⎛ ⎝  x∈u−1(y) ϕ(x) ⎞ ⎠ dHn(y) (2.8)

for any Borel functionϕ : → [0, ∞].

Here the Jacobian of a seminorm s onRn is defined by

J(s) = α(n)

 Sn−1s(v)

−ndHn−1(v)−1 (2.9)

Following [31] we also define the maximal stretch of a seminorm s by

I+(s) = max{s(v) : |v| = 1}.

Note that the inequality J(s) ≤ I+n(s) always holds, and I+(apmdxu) = gu almost every-where for u ∈ Nloc1,n( ; X), see [31, Section 4]. We say that s is K -quasiconformal if

I+n(s) ≤ K J(s). A Sobolev map u ∈ Nloc1,n( ; X) is called infinitesimally K -quasiconformal,

if apmdxu is K -quasiconformal for almost every x∈ .

The (Reshetnyak) energy and (parametrized Hausdorff) area of u ∈ Nloc1,n( ; X) are defined as En+(u) :=  I n +(apmdxu)dx Ar ea(u) :=  J(apmdxu)dx. Quasiconformality

We refer to [17–20] for various definitions of quasiconformality and their relationship in metric spaces, and only mention what is sometimes known as geometric quasiconformality. A homeomorphism u : (X, μ) → (Y , ν) between metric measure spaces is said to be

K -quasiconformal with “index” Q≥ 1, if

ModQ ≤ K ModQu() (2.10)

for any curve family in X. Without further assumptions on the geometry of the spaces, the modulus condition (2.10) is fundamentally one-sided; cf. the discussion in the introduction. We refer to [44] for an equivalent characterization in terms of analytic quasiconformality and note that there is a corresponding result for monotone maps. Recall that a map is monotone if the preimage of every point is a connected set.

(11)

Proposition 2.4 [35, Proposition 3.4] Let M beD or S2. If u∈ N1,2(M; X) is a continuous, surjective, monotone and infinitesimally K -quasiconformal map into a complete metric space X , then

Mod2 ≤ K Mod2u() for any curve family in D.

We remark that in [35] the result is proven for M= D. The same proof however carries over to S2with minor modifications.

Quadratic isoperimetric inequality and minimal discs

Given a proper metric space X and a closed Jordan curve in X, we denote

(, X) := {u ∈ N1,2(D; X) : u spans }.

Here, a function u ∈ N1,2(D; X) is said to span  if u|

S1 agrees a.e. with a monotone

parametrization of. We call u ∈ (, X) a minimal disc spanning  if u has least area among all maps in(, X) and is furthermore of minimal Reshetnyak energy E+2 among all area minimizers in(, X). In [31] it has been shown that, as soon as(, X) = ∅, there exists a minimal disc u spanning and every such u is infinitesimally√2-quasiconformal. Definition 2.5 A metric space X is said to satisfy a(C, l0)-quadratic isoperimetric inequality if, for any Lipschitz curveγ : S1 → X with (γ ) < l0, there exists u ∈ N1,2(D; X) with u|S1 = γ and

Ar ea(u) ≤ C · (γ )2.

If X satisfies a quadratic isoperimetric inequality then minimal discs u : D → X lie in

Nloc1,p(D, X) for some p > 2 and, if furthermore  satisfies a chord-arc condition, then u∈ N1,p(D, X).

If a proper geodesic metric space Z is homeomorphic to a 2-manifold, then it satisfies a

(C, l0)-quadratic isoperimetric inequality if and only if every Jordan curve  in Z such that () < l0bounds a Jordan domain U ⊂ Z which satisfies

H2(U) ≤ C · ()2. (2.11)

This follows from the proof of [35, Theorem 1.4] together with the observation leading to [6, Corollary 1.5]. We will use this fact in Sect.6to establish the quadratic isoperimetric inequality for Zuin Theorem1.9.

3 A geometric characterization of the Sobolev-to-Lipschitz property

In this section we prove Theorem1.7. We assume throughout this section that X is a metric measure space and p ≥ 1. We will need the following consequence of the Sobolev-to-Lipschitz property for measurable maps with 1 as p-weak upper gradients.

Lemma 3.1 Suppose X has the p-Sobolev-to-Lipschitz property, and let V be a separable

Banach space. Then any measurable function f : X → V with 1 as p-weak upper gradient has a 1-Lipschitz representative.

(12)

Proof We prove the claim first when V = R. Fix x0∈ X and consider the function wk:= (k − dist(·, B(x0, k))+

for each k∈ N. Note that wkis 1-Lipschitz andwk∈ N1,p(X). The functions

fk:= min{wk, f }

have 1 as p-weak upper gradient (see [22, Proposition 7.1.8]) and fk ∈ N1,p(X). By the Sobolev-to-Lipschitz property there is a set N⊂ X with μ(N) = 0 and 1-Lipschitz functions ¯fksuch that fk(x) = ¯fk(x) if x /∈ N, for each k ∈ N. Since fk → f pointwise everywhere we have, for x, y /∈ N, that

| f (x) − f (y)| = lim

k→∞| ¯fk(x) − ¯fk(y)| ≤ d(x, y).

Thus f has a 1-Lipschitz representative.

Next, let V be a separable Banach space and let{xn} ⊂ V be a countable dense set. Given

f as in the claim, the functions fn := dist(xn, f ) have 1 as p-weak upper gradient, and thus there is a null set E ⊂ X and 1-Lipschitz functions ¯fn so that fn(x) = ¯fn(x) whenever

n∈ N and x /∈ E. For x, y /∈ N, we have

f (x) − f (y) V = sup

n | fn(x) − fn(y)| = supn | ¯fn(x) − ¯fn(y)| ≤ d(x, y).

We again conclude that f has a 1-Lipschitz representative.  Proposition 3.2 Let X be infinitesimally doubling and p-thick quasiconvex with constant C.

Then every u∈ N1,p(X) satisfying gu ≤ 1, has a C-Lipschitz representative.

Proof Let 0a path family of zero p-modulus such that

|u(γ (1)) − u(γ (0))| ≤ 

γ1= (γ )

wheneverγ /∈ 0. By Lusin’s theorem there is a decreasing sequence of open sets Um⊂ X such that, for each m∈ N, μ(Um) < 2−m and u|X\Um is continuous. Fix m ∈ N, and let x and y be distinct density points of X\ Um. Then for everyδ > 0 we have that

Modp((B(x, δ) \ Um, B(y, δ) \ Um; C)) > 0. Forγδ∈ (B(x, δ) \ Um, B(y, δ) \ Um; C) \ 0one has

|u(γδ(1)) − u(γδ(0))| ≤ (γ ) ≤ C · d(γδ(0), γδ(1)) ≤ Cd(x, y) + 2Cδ. Lettingδ → 0 and using the continuity of u|X\Um we obtain

|u(x) − u(y)| ≤ Cd(x, y). (3.1)

Letting m → ∞ one sees that (3.1) holds forμ-almost every x, y ∈ X and hence u has a

C-Lipschitz representative. 

Proof of Theorem1.7 One implication is directly implied by the preceeding proposition. Indeed, assume X is p-thick geodesic and let f ∈ N1,p(X) have 1 as a p-weak upper

gradient. By Proposition3.2, f has a C-Lipschitz representative for every C > 1. So the continuous representative of f is 1-Lipschitz.

Now assume X has the p-Sobolev-to-Lipschitz property but is not p-thick geodesic. Let

(13)

By looking at density points of E and F respectively we may assume without loss of generality that

0< dist(E, F) =: D and diam(E), diam(F) ≤(C − 1)D

4C .

There exists a non-negative Borel function g ∈ Lp(X) for which  γg = ∞ for every γ ∈ (E, F; C). Denote 0:=  γ :  γg= ∞  ,

whence(E, F; C) ⊂ 0, Modp(0) = 0 and for γ1, γ2 /∈ 0one hasγ1γ2 /∈ 0whenever the concatenationγ1γ2is defined. Define the functionv : X → [0, ∞] by

v(x) = lim n→∞inf  γ(1 + g/n) : γ ∈ (E, x)  .

The functionv is measurable by [22, Theorem 9.3.1] and satisfiesv|E≡ 0. We also have |v(y) − v(x)| ≥ C · D ≥ C ·  d(x, y) −(C − 1)D 2C  ≥C+ 1 2 d(x, y) (3.2) whenever y∈ F and x ∈ E.

For anyγ /∈ 0by closedness under composition one hasv(γ (0)) = ∞ iff v(γ (1)) = ∞. Furthermore ifv(γ (0)) = ∞, then

|v(γ (1)) − v(γ (0))| ≤ (γ ) = 

γ1 (3.3)

So g≡ 1 is a p-weak upper gradient of v.

If we had thatv ∈ N1,p(X), the Sobolev-to-Lipschitz property and (3.2) would lead to a

contradiction. A simple cut off argument remains to complete the proof. Definew : X → [0, ∞] by

w(x) := ((C + 2)D − d(E, x))+.

Thenw has 1 as a upper gradient, w ∈ N1,p(X) is compactly supported and for x ∈ E,

y∈ F one has

w(y) = (C + 2)D − d(E, y) ≥ C D ≥C+ 1

2 d(x, y). (3.4)

Set u:= min{v, w}. Then u ∈ N1,p(X) and gu ≤ 1. By the Sobolev-to-Lipschitz property

u has a 1-Lipschitz representative. But for every x∈ E, y ∈ F

|u(x) − u(y)| = u(y) ≥ C+ 1 2 d(x, y).

Asμ(E)μ(F) > 0 these two observations lead to a contradiction.  Example 3.3 For p ∈ [1, ∞) let Zp := {(x, y) ∈ R2 : |y| ≤ |x|p} be endowed with the intrinsic length metric and Lebesgue measure. Then Z is q-thick geodesic if and only if q > p + 1. This follows by Example 1 in [12]. Since Zp is doubling, it also has the

(14)

4 Volume rigidity

We prove Proposition1.8using the fact that volume preserving 1-Lipschitz maps between rectifiable spaces are essentially length preserving. We say that a map f : X → Y between

n-rectifiable spaces is volume preserving, if fHnX=HnY.

Proposition 4.1 Let f : X → Y volume preserving 1-Lipschitz map between n-rectifiable

metric spaces. There exists a Borel set N ⊂ X withHn(N) = 0 so that for every absolutely continuous curveγ in X with |γ−1N| = 0 we have

(γ ) = ( f ◦ γ ).

Proof Let Ei ⊂ Rn and gi : Ei → X be Borel sets and Lipschitz maps, respectively, satisfyingHn(E) = 0, where

E:= X \

i

gi(Ei).

We may assume that the sets gi(Ei) are pairwise disjoint. For each i, fix Lipschitz extensions ¯gi : Rn → l∞and ¯fi : Rn → lof gi and f ◦ gi, respectively (here we embed X and Y isometrically into l∞).

Lemma 4.2 Let i∈ N. For almost every x ∈ Ei, we have

mdx ¯gi= mdx ¯fi. (4.1)

Proof of Lemma4.2 Since f is 1-Lipschitz we have, at almost every point x ∈ Ei where both mdx ¯gi and mdx ¯fi exist, the inequality mdx ¯fi ≤ mdx ¯gi, which in particular implies

J(mdx ¯fi) ≤ J(mdx ¯gi) forLn-almost every x ∈ Ei. By the area formula [25, Corollary 8] and the fact that f is volume preserving, it follows that

 Ei J(mdx ¯gi)dLn(x) =Hn(gi(Ei)) =Hn( f (gi(Ei))) =  Ei J(mdx ¯fi)dLn(x);

cf. Theorem2.3. Consequently J(mdx ¯gi) = J(mdx ¯fi) forLn-almost every x∈ Ei. Thus

mdx ¯gi= mdx ¯fiforLn-almost every x ∈ Ei. 

For each i∈ N, let Ui denote the set of those density points x of Eifor which mdx ¯gi= mdx ¯fi. Lemma4.2implies thatLn(Ei\ Ui) = 0. Set

N := E ∪

i

gi(Ei\ Ui),

whenceHn(N) = 0. Let γ : [0, 1] → X be a Lipschitz path. For each i ∈ N, denote

Ki = γ−1(gi(Ui)) ⊂ [0, 1], and suppose γi : [0, 1] → Rn is a Lipschitz extension of the “curve fragment” gi−1◦ γ : Ki → Rn.

Let t∈ Kibe a density point of Ki, for which| ˙γt|, |( f ◦γ )t| and γi(t) exist, and γ (t) /∈ N. Thenγi(t) ∈ Uiand we have

mdγi(t)¯gi(γi(t)) = lim h→0

¯gi(γi(t) + hγi(t)) − ¯gi(γi(t)) l

(15)

= lim h→0 ¯gi(γi(t + h)) − ¯gi(γi(t)) l+ o(h) h = lim h→0 t+h∈Ki ¯gi(γi(t + h) − ¯gi(γi(t)) lh = lim h→0 t+h∈Ki d(γ (t + h), γ (t)) h = | ˙γt|.

The same argument with ¯fi in place of¯giyields

mdγi(t) ¯fi(γi(t)) = |( f ◦ γ )t|. Sinceγi(t) ∈ Ui, we have

| ˙γt| = mdγi(t)¯gi(γi(t)) = mdγi(t) ¯fi(γi(t)) = |( f ◦ γ )t|.

Suppose now thatγ : [0, 1] → X is a absolutely continuous path with |γ−1(N)| = 0. We may assume thatγ is constant speed parametrized. Since |γ−1(N)| = 0 it follows that the union of the sets Ki over i ∈ N has full measure in [0, 1]. Since, for each i ∈ N, a.e. Ki satisfies the conditions listed above, we may compute

(γ ) =  1 0 | ˙γt|dt =  i  Ki | ˙γt|dt =  i  Ki |( f ◦ γ )t|dt = ( f ◦ γ ).  The proof yields the following corollary.

Corollary 4.3 Let f : X → Y be a surjective 1-Lipschitz map between rectifiable spaces with

Hn(X) =Hn(Y ) < ∞. Then f is volume preserving and, in particular, ( f ◦ γ ) = (γ )

for∞-almost every γ in X.

We close this section with the proof of Proposition1.8, which connects volume rigidity and the Sobolev-to-Lipschitz property.

Proof of Proposition1.8 Since f−1 is quasiconformal, f−1◦ γ is rectifiable for n-almost every curveγ in Y . Proposition4.1implies that

gf−1 ≤ 1.

Since Y has the n-Sobolev-to-Lipschitz property, it follows that f−1has a 1-Lipschitz rep-resentative, cf. Lemma3.1. By continuity of f−1it coincides with this representative.  The Sobolev-to-Lipschitz property is crucial for the conclusion of Proposition1.8, as the next example shows.

Example 4.4 Let Y = (D, dw), where

w(x) = 1 − (1 − |x|)χ[0,1]×{0}, x ∈ D,

and the metric dwis given by,

dw(x, y) = inf

γ ∈(x,y) 

γw, x, y ∈ D.

The map id: D → Y is a volume preserving 1-Lipschitz homeomorphism and the inverse is quasiconformal, but not Lipschitz continuous.

(16)

Indeed, since[0, 1] × {0} has zero measure, we have J(apmd id) = 1 almost everywhere, whence the area formula implies thatH2(D) =H2

dw(D).

5 Construction of essential pull-back metrics

Let X be a metric measure space and let(X) denote the set of Lipschitz curves [0, 1] → X. Recall that p-almost every curve in X admits a Lipschitz reparametrization. In this section we construct essential pull-back distances by Sobolev maps. The key notion here is the essential infimum of a functional over a path family.

Definition 5.1 Let p ≥ 1, F : (X) → [−∞, ∞] be a function and  ⊂ (X) a path family. Define the p-essential infimum of F over by

essinfp  F= essinfp γ ∈ F(γ ) := sup Modp(0)=0 inf{F(γ ) : γ ∈  \ 0}

with the usual convention inf∅ = ∞.

It is clear that the supremum may be taken over curve familiesρfor non-negative Borel functionsρ ∈ Lp(X), where :=  γ :  γ ρ = ∞  .

Remark 5.2 We have the following alternative expression for essinfp: essinfp  F= max{λ > 0 : Modp( ∩ F(λ)) = 0} = inf{λ > 0 : Modp( ∩ F(λ)) > 0}. Here F(λ) := {γ ∈ (X) : F(γ ) < λ}.

Let Y be a complete metric space, p≥ 1, and u ∈ Nloc1,p(X; Y ). Given a path family  in

X we set

essu,p() := essinfp

γ ∈ (u ◦ γ ).

Definition 5.3 Let u∈ Nloc1,p(X; Y ). Define the essential pull-back distance

du,p : X × X → [0, ∞] by du,p(x, y) := lim

δ→0essu,p(( ¯B(x, δ), ¯B(y, δ)), x, y ∈ X.

In general, the essential pull-back distance may assume both values 0 and∞ for distinct points, and it need not satisfy the triangle inequality. We denote by du,pthe maximal pseu-dometric not greater than du,p , given by

du,p(x, y) = inf  n  i=1 du,p(xi, xi−1) : x0, . . . , xn ∈ X, x0 = x, xn = y  , x, y ∈ X.

(17)

The maximal pseudometric below du,p may also fail to be a finite valued. We give two situations that guarantee finiteness and nice properties of the arising metric space. Firstly, in Sect.5.2we consider the simple case u = id : X → X and use it to prove Theorem 1.4. Secondly, in Proposition5.12we prove that, when X supports a Poincaré inequality and u ∈ Nloc1,p(X; Y ) for large enough p, the distance in Definition 5.3is a finite valued pseudometric.

For both we need the notion of regular curves, whose properties we study next. 5.1 Regular curves

Throughout this subsection(X, d, μ) is a metric measure space. We define a metric on (X) by setting

d∞(α, β) := sup

0≤t≤1d(α(t), β(t))

for any two Lipschitz curvesα, β. By a simple application of the Arzela-Ascoli theorem it follows that((X), d) is separable.

Definition 5.4 A curveγ ∈ (X) is called (u, p)-regular if u ◦ γ is absolutely continuous and

essu,pB(γ, δ) ≤ (u ◦ γ ) for everyδ > 0.

Proposition 5.5 p-almost every Lipschitz curve is(u, p)-regular. For the proof we will denoteu,p(λ) = {γ : (u ◦ γ ) < λ}.

Proof of Proposition5.5 Denote by0the set of curves in(X) which are not (u, p)-regular. We may write0= 1∪ 2, where

1= {γ : u ◦ γ not absolutely continuous}, 2= 0\ 1.

By the fact that u ∈ Nloc1,p(X; Y ) we have Modp(1) = 0. It remains to show that Modp(2) = 0. For any δ > 0 and γ ∈ (X), set

ε(δ, γ ) := essu,pB(γ, δ) − (u ◦ γ )

δ(γ ) := sup{δ > 0 : ε(δ, γ ) > 0}.

Note that δ(γ ) > 0 if and only if γ ∈ 2. Using Remark5.2we make the following observation which holds forδ, ε > 0:

Modp(B(γ, δ) ∩ u,p((u ◦ γ ) + ε)) = 0 implies δ ≤ δ(γ ) and ε ≤ ε(δ, γ ). (5.1) Moreover, for anyγ ∈ 2andδ > 0, we have

Modp(B(γ, δ) ∩ u,p((u ◦ γ ) + ε(δ, γ ))) = 0. (5.2) Let{γi}i∈N ⊂ 2 be a countable dense set. For each i, k ∈ N and rational r > 0 let γi,k,r∈ B(γi, r) ∩ 2satisfy

(18)

By (5.2) it suffices to prove that

2

i,k∈N r,δ∈Q+

B(γi,k,r, δ) ∩ u,p((u ◦ γi,k,r) + ε(δ, γi,k,r)).

For anyγ ∈ 2 let i ∈ N be such that d∞(γ, γi) < δ(γ )/8. Choose rational numbers

r, δ ∈ Q+such that d∞(γi, γ ) < r < δ(γ )/8 and δ(γ )/4 < δ < δ(γ )/2, and a natural number k∈ N so that 1/k < ε(2δ, γ ).

We will show that

γ ∈ B(γi,k,r, δ) ∩ u,p((u ◦ γi,k,r) + ε(δ, γi,k,r)). (5.3) Indeed, sinceγi,k,r ∈ B(γi, r) ∩ 2, the triangle inequality yields

d∞(γ, γi,k,r) ≤ d∞(γ, γi) + d∞(γi, γi,k,r) < 2r < δ(γ )/4 < δ. In particular

γ ∈ B(γi,k,r, δ) (5.4)

and also

B(γi,k,r, δ) ⊂ B(γ, 2δ) (5.5) To bound the length of u◦ γ , observe that

(u ◦ γi,k,r) < inf{(u ◦ β) : β ∈ B(γi, r) ∩ 2} + 1/k < (u ◦ γ ) + 1/k

< (u ◦ γ ) + ε(2δ, γ )

Settingε := (u ◦ γ ) + 1/k − (u ◦ γi,k,r) > 0 it follows that

u,p((u ◦ γi,k,r) + ε) ⊂ u,p((u ◦ γ ) + ε(2δ, γ )) (5.6) Combining (5.5) and (5.6) with (5.2) we obtain

Modp(B(γi,k,r, δ) ∩ u,p((u ◦ γi,k,r) + ε)) = 0, which, by (5.1) yieldsε ≤ ε(δ, γi,k,r). Thus

(u ◦ γ ) < (u ◦ γ ) + 1/k = (u ◦ γi,k,r) + ε ≤ (u ◦ γi,k,r) + ε(δ, γi,k,r) which, together with (5.4), implies (5.3). This completes the proof.  For the next two results we assume that u ∈ Nloc1,p(X; Y ) is continuous. We record the following straightforward consequence of the definition of du,p and the continuity of u as a lemma.

Lemma 5.6 For each x, y ∈ X we have

d(u(x), u(y)) ≤ d

u,p(x, y)

For the next proposition, we denote byu,p(γ ) the length of a curve γ with respect to the pseudometric du,p.

Proposition 5.7 Ifγ : [a, b] → X is (u, p)-regular and a ≤ t ≤ s ≤ b then γ |[t,s]is (u, p)-regular and

(19)

Proof Denote η := γ |[s,t]. Let n :=  c|[s,t] : c ∈ B  γ,1 n  ∩ u,p  (u ◦ γ ) +1 n  .

Sinceγ is (u, p)-regular we have Modp(n) ≥ Modp  B  γ,1 n  ∩ u,p  (u ◦ γ ) +1 n  > 0.

We claim that for everyε > 0 there exists n0∈ N so that

n⊂ B  η,1 n  ∩ u,p((u ◦ η) + ε) (5.7)

for all n≥ n0. Indeed, otherwise there existsε0> 0 and a sequence γnk ∈ B  γ, 1 nk  ∩ u,p  (u ◦ γ ) + 1 nk  so that (u ◦ γnk|[s,t]) ≥ (u ◦ η) + ε0. Thus (u ◦ γ ) + 1 nk ≥ (u ◦ γnk) =   u◦ γnk|[s,t]  + u◦ γnk|[s,t]c  ≥  (u ◦ η) + ε0+   u◦ γnk|[s,t]c  yielding u◦ γ|[s,t]c+ 1 nk ≥   u◦ γnk|[s,t]c  + ε0. (5.8)

By taking lim infk→∞in (5.8) we obtain

u◦ γ|[s,t]c≥ u◦ γ|[s,t]c+ ε0, which is a contradiction.

Thus (5.7) holds true. Ifε, δ > 0, let n ∈ N be such that (5.7) holds andδ > 1/n. We have

0< Modpn ≤ Modp(B(η, δ) ∩ u,p((u ◦ η) + ε)),

implying essu,pB(η, δ) ≤ (u ◦ η) + ε. Since ε > 0 is arbitrary, η is (u, p)-regular. To prove the equality in the claim note that, sinceγ is (u, p)-regular we have

du,p(γ (t), γ (s)) ≤ du,p (γ (t), γ (s)) ≤ lim

δ→0essu,pB(γ |[s,t], δ) ≤ (u ◦ γ |[s,t]) for any s≤ t. It follows that

u,p(γ ) ≤ (u ◦ γ ). On the other hand Lemma5.6implies that

d(u(γ (t)), u(γ (s))) ≤ du,p(γ (t), γ (s))

(20)

5.2 The Sobolev-to-Lipschitz property in thick quasiconvex spaces

In this subsection let p∈ [1, ∞] and X = (X, d, μ) be p-thick quasiconvex with constant

C ≥ 1. Consider the map u = id ∈ Nloc1,p(X; X). We denote by dp the pseudometric du,p associated to u.

Lemma 5.8 dpis a metric on X , and satisfies d≤ dp ≤ Cd.

Proof Let x, y ∈ X be distinct and δ > 0. Since

Modp(B(x, δ), B(y, δ); C) > 0 it follows that

essid,p(B(x, δ), B(y, δ)) ≤ C(d(x, y) + 2δ),

implying did,p(x, y) ≤ Cd(x, y). Thus dp≤ Cd and in particular dpis finite-valued. Lemma 5.6implies that d(x, y) ≤ dp(x, y) for every x, y ∈ X. These estimates together prove the

claim. 

We are now ready to prove Theorem1.4. For the proof, we denote by Xp the space

(X, dp, μ) and by Bp(x, r) balls in X with respect to the metric dp;p and ModXp,prefer to the length of curves and p-modulus taken with respect to Xp.

Proof of Theorem1.4 By Lemma5.8it follows that Xpis an infinitesimally doubling metric measure space and ModXp,p = 0 if and only if Modp = 0. In particular if ∗denotes the set of curvesγ in X such that (γ ) = p(γ ) then by Propositions5.5and5.7

ModXp,p∗= Modp= 0. (5.9)

From (5.9) and the definition of modulus it follows that ModXp,p = Modp for every family of curves in X.

By (5.9) and Proposition 3.2every f ∈ N1,p(Xp) with gf ≤ 1 has a C-Lipschitz representative. If x, y ∈ X are distinct and δ, ε > 0 note that the curve family

1= {γ ∈ (B(x, δ), B(y, δ)) : (γ ) ≤ essid,p(B(x, δ), B(y, δ)) + ε}

satisfies

Modp1> 0 by Remark5.2and the fact thatε > 0. Note also that

(γ ) ≤ essid,p(B(x, δ), B(y, δ)) + ε ≤ did,p(γ (1), γ (0)) + ε

forγ ∈ 1. Let f ∈ N1,p(X

p) satisfy gf ≤ 1 almost everywhere. Let ¯f be the Lipschitz representative of f , and0a curve family with Modp0= 0 and

| f (γ (1) − f (γ (0))| ≤ p(γ ) = (γ ) wheneverγ /∈ 0. We have

Modp(1\ 0) > 0,

so that there existsγδ∈ 1\ 0. We obtain

(21)

Lettingδ → 0 yields | ¯f(x) − ¯f(y)| ≤ did,p(x, y) + ε. Since x, y ∈ X and ε > 0 are arbitrary it follows that ¯f is 1-Lipschitz with respect to dp. In particular Xp has the p-Sobolev-to-Lipschitz property and hence Theorem1.7implies it is p-thick geodesic. We

prove the minimality in Proposition5.9. 

The metric dpis the minimal metric above d which has the p-Sobolev-to-Lipschitz property. Proposition5.9provides a more general statement, from which the minimality discussed in the introduction immediately follows.

Proposition 5.9 Let Y be a p-thick geodesic metric measure space and f : Y → X be a

volume preserving 1-Lipschitz map. Then the map fp:= f : Y → (X, dp)

is volume preserving and 1-Lipschitz. Proof It suffices to show that

did,p( f (x), f (y)) ≤ d(x, y), x, y ∈ Y .

For any A> 1, the curve family

 := {γ ∈ (B(x, δ), B(y, δ)) : (γ ) ≤ Ad(γ (1), γ (0))}

has positive p-modulus in Y . Since f is volume preserving and 1-Lipschitz we have 0< ModY,p ≤ ModX,p f.

Ifγ ∈  then f ◦ γ ∈ (B( f (x), δ), B( f (y), δ)) and, moreover

( f ◦ γ ) ≤ (γ ) ≤ Ad(γ (1), γ (0)) ≤ Ad(x, y) + 2Aδ.

It follows that essid,p(B( f (x), δ), B( f (y), δ)) ≤ Ad(x, y) + 2Aδ, and thus

did,p( f (x), f (y)) ≤ Ad(x, y) + 2Aδ.

Since A> 1 and δ > 0 are arbitrary, the claim follows.  We have the following immediate corollary.

Corollary 5.10 Assume X is infinitesimally doubling and p-thick geodesic. Then d= dp. Before considering essential pull-back metrics by non-trivial maps, we prove Proposition1.5. The proof is based on the fact that spaces with Poincaré inequality are thick quasiconvex, and on the independence of the minimal weak upper gradient on the exponent.

Proposition 5.11 ([12]) Let p∈ [1, ∞] and X be a doubling metric measure space satisfying

a p-Poincaré inequality. There is a constant C≥ 1 depending only on the data of X so that X is p-thick quasiconvex with constant C.

The inverse implication in Proposition5.11only holds if p= ∞, see [11,12].

Proof of Proposition1.5 Assume X is a doubling metric measure space supporting a p-Poincaré inequality, and q ≥ p ≥ 1. Since Modq = 0 implies Modp = 0, see [2, Proposition 2.45], we have

(22)

for some constant C depending only on p and the data of X . Fix x0 ∈ X and consider the function

f : X → R, x → dp(x0, x).

Then f is Lipschitz and, by Propositions5.5and5.7, it has 1 as a p-weak upper gradient. Since X is doubling and supports a p-Poincaré inequality, [2, Corollary A.8] implies that the minimal q-weak and p-weak upper gradients of f agree almost everywhere, and thus 1 is a

q-weak upper gradient of f , i.e.

| f (γ (1)) − f (γ (0))| ≤ (γ ) ≤ q(γ )

for q-almost every curve. The space Xqhas the q-Sobolev-to-Lipschitz property, see Theorem 1.7, and the 1-Lipschitz representative of f agrees with f everywhere, since f is continuous. By this and the definition of dq we obtain

dp(x0, x) ≤ dq(x0, x), x ∈ X. Since x0∈ X is arbitrary the equality dp= dq follows.

For the remaining equality, note that d ≤ d ≤ dess ≤ Cd. Indeed, X supports an ∞-Poincaré inequality, whence [11, Theorem 3.1] implies the rightmost estimate with a constant

C depending only on the data of the∞-Poincaré inequality. As above, the function g: X → R, x → dess(x0, x)

satisfies

|g(γ (1)) − g(γ (0))| ≤ (γ ) ≤ ∞(γ )

for∞-almost every γ , from which the inequality dess ≤ d∞follows.  Note that in the proof of Proposition5.11we use that in p-Poincaré spaces the q-weak upper gradient does not depend on q≥ p. In general such an equality is not true, see [9], and we do not know whether we can weaken the assumptions in Proposition1.5from p-Poincaré inequality to p-thick quasiconvexity.

5.3 Essential pull-back metrics by Sobolev maps

Throughout this subsection(X, d, μ) will be a doubling metric measure space satisfying (2.2) with Q ≥ 1 and supporting a weak (1, Q)-Poincaré inequality, and Y = (Y , d) a proper metric space.

We will use the following observation without further mention. If p> Q, and u : X → Y has a Q-weak upper gradient in Llocp (X), then u has a representative ¯u ∈ Nloc1,p(X; Y ) and the minimal p-weak upper gradient of¯u coincides with the minimal Q-weak upper gradient of

u almost everywhere. See [2, Chapter 2.9 and Appendix A] and [22, Chapter 13.5] for more details.

The next proposition states that higher regularity of a map is enough to guarantee that the essential pull-back distance in Definition5.3is a finite-valued pseudometric.

Proposition 5.12 Let p> Q, and suppose u ∈ Nloc1,p(X; Y ). The pull-back distance du :=

du,Qin Definition5.3is a pseudometric satisfying du(x, y) ≤ C diam BQ/pd(x, y)1−Q/p  −  σ Bg p udμ 1/p (5.10)

(23)

whenever B⊂ X is a ball and x, y ∈ B, where the constant C depends only on p and the data of X .

For the proof of Proposition5.12we define the following auxiliary functions. Let x ∈ X,

δ > 0 and 0a curve family with ModQ0= 0. Set f := fx,δ,0 : X → R by

f(y) = inf{(u ◦ γ ) : γ ∈ ( ¯B(x, δ), y) \ 0}, y ∈ X.

Whenρ ∈ LQ(X) is a non-negative Borel function, whence ModQ(ρ) = 0, we denote

fx,δ,ρ := fx,δ,ρ.

Lemma 5.13 Let x, δ and ρ be as above. The function f = fx,δ,ρis finiteμ-almost every-where and has a representative in Nloc1,p(X) with p-weak upper gradient gu. The continuous

representative ¯f of f satisfies

| ¯f(y) − ¯f(z)| ≤ C(diam B)Q/pd(y, z)1−Q/p

 −  σ Bg p udμ 1/p . (5.11) Proof Let g ∈ Lp

loc(X) be a genuine upper gradient of u and let ε > 0 be arbitrary. We fix

a large ball B ⊂ X containing ¯B(x, δ) and note that there exists x0 ∈ ¯B(x, δ) for which

MB(g + ρ)Q(x0) < ∞, since (g + ρ)|B∈ LQ(B); cf. (2.3). Arguing as in [41, Lemma 4.6] we have that f(y) ≤ inf  γ(g + ρ) : γ ∈ x0y\ ρ  < ∞

for almost every y∈ B.

Letγ /∈ ρ be a curve such that f(γ (1)), f (γ (0)) < ∞ andγg < ∞. We may

assume that | f (γ (1)) − f (γ (0))| = f (γ (1)) − f (γ (0)) ≥ 0. If β is an element of ¯(B(x, δ), γ (0)) \ ρ such that(u ◦ β) < f (γ (0)) + ε then the concatenation γβ sat-isfiesγβ ∈ ( ¯B(x, δ), γ (1)) \ ρ. Thus

| f (γ (1)) − f (γ (0))| ≤ (u ◦ γβ) − (u ◦ β) + ε = (u ◦ γ ) + ε ≤ 

γ g+ ε. It follows that g∈ Llocp (X) is a Q-weak upper gradient for f . By [2, Corollary 1.70] we have that f(y) < ∞ for Q-quasievery y ∈ B, and f ∈ Nloc1,Q(X); cf. [22, Theorem 9.3.4].

Moreover, since g ∈ Llocp (X), f has a continuous representative ¯f ∈ Nloc1,p(X) which

satisfies (5.11), cf. [22, Theorem 9.2.14]. 

Proof of Proposition5.12 To prove the triangle inequality, let x, y, z ∈ X be distinct. Take

δ > 0 small, ρ ∈ LQ(X) non-negative, and let E ⊂ X be a set of Q-capacity zero such that

fx,δ,ρ and fy,δ,ρ agree with their continuous representatives outside E. Remember that

αβ /∈ ρwheneverα, β /∈ ρ. We have that

du,Q(x, z) + du,Q(z, y) ≥ inf

w∈ ¯B(z,δ)fx,δ,ρ∪E(w) + infv∈ ¯B(z,δ) fy,δ,ρ∪E(v)

≥ inf

w,v∈ ¯B(z,δ)\E[ fx,δ,ρ(w) + fy,δ,ρ(v)]. Together with the estimate (5.11) this yields

du,Q (x, z) + du,Q (z, y) ≥ inf

w∈ ¯B(z,δ)\E[ fx,δ,ρ(w) + fy,δ,ρ(w)] − Cδ

(24)

≥ inf{(u ◦ γ ) : γ ∈ ( ¯B(x, δ), ¯B(y, δ)) \ ρ} − Cδ1−Q/p, where C depends on u, x and y as well as the data. Since δ > 0 and ρ are arbitrary it follows that

du,Q(x, z) + du,Q(z, y) ≥ du,Q(x, y).

Moreover, for any y∈ ¯B(y, δ) \ E,

| fx,δ,ρ(y)| ≤ C diam Bp/Qd(x, y)1−Q/p  −  σ Bg p udμ 1/p ≤ C diam Bp/Q(d(x, y) + δ)1−Q/p  −  σ Bg p udμ 1/p

Taking supremum overρ, and letting δ tend to zero, we obtain (5.10).  Remark 5.14 A slight variation of the proof of Proposition5.12shows that also in the setting of Theorem1.4the essential pull-back distance defines a metric. In particular it satisfies the triangle inequality and one does not have to pass to the maximal semimetric below. In this case in the argument instead of the Morrey embedding one applies Proposition3.2. Fix a continuous map u∈ Nloc1,p(X; Y ), where p > Q, and denote by Yuthe set of equivalence classes[x] of points x ∈ X, where x and y are set to be equivalent if du,Q (x, y) = 0. The pseudometric du,Q defines a metric du on Yu by

du([x], [y]) := du,Q(x, y), x, y ∈ X. The natural projection map



Pu : X → Yu, x → [x] is continuous by (5.10). The map u factors as u= u◦ Pu, where

u: Yu → Y , [x] → u(x) is well-defined and 1-Lipschitz; cf. Lemma5.6.

Lemma 5.15 Under the given assumptions we obtain the following properties. (1) Pu ∈ Nloc1,p(X; Yu) and gPu = gu μ-almost everywhere.

(2) If u is proper then Yuis proper and the projection Pu: X → Yuis proper and monotone. Recall here that a map is called proper if the preimage of every compact set is compact or equivalently if the preimage of every singleton is compact.

Proof (1) Suppose γ is a (u, Q)-regular curve such that gu is an upper gradient of u along

γ . Then by Proposition5.7

du(Pu◦ γ (1), Pu◦ γ (0)) ≤ (u ◦ γ ) ≤ 

γgu.

By Proposition5.5this implies that Pu ∈ Nloc1,Q(X; Yu) and gPu ≤ gu. The opposite inequality follows becauseu is 1-Lipschitz.

(25)

(2) The factorization implies that K ⊂ Pu(u−1(u(K ))), for K ⊂ Yu and hence Yu is proper. Similarly Pu−1(K ) ⊂ u−1(u(K )) and hence Puis proper. The proof of monotonicity is an adaptation of the proof of [33, Lemma 6.3].

Assume Pu−1(y) is not connected for some y ∈ Yu. Then there are compact sets K1, K2for which dist(K1, K2) > 0 and Pu−1(y) = K1∪K2. Let S denote the closed and non-empty set of

points X whose distance to K1and K2agree. Let a be the minimum of x→ distu(y, Pu(x)) on S. Note here that Pu(S) is closed as Yu and Pu are proper and hence the infimum is attained and positive. Let ki ∈ Ki for i = 1, 2. Since du,Q(k1, k2) = 0, for every ε > 0 and small enoughδ > 0, Proposition5.5implies the existence of a(u, Q)-regular curve

γ ∈ (B(k1, δ), B(k2, δ)) with (u ◦ γ ) < ε. If δ is chosen small enough the curve must

intersect S at some point s:= γ (t). Since γ is (u, Q)-regular, γ |[0,t]is(u, Q)-regular and it follows that

a≤ du,Q(s, k1) ≤ du,Q(s, γ (0)) + Cδ1−Qp ≤ (u ◦ γ ) + Cδ1−Qp ≤ ε + Cδ1−Qp; cf. Proposition5.7 and (5.10). Choosing ε > 0 and δ > 0 small enough this yields a contradiction. Thus Pu−1(y) is connected for every y ∈ Yu.  Assume additionally that Y is endowed with a measureν such that Y = (Y , d, ν) is a metric measure space. We equip(Yu, du) with the measure νu := uν, which is characterized by the property

νu(E) = 

Y

#(u−1(y) ∩ E)dν(y), E ⊂ YuBorel; (5.12) cf. [13, Theorem 2.10.10]. Note that in generalνu is not aσ -finite measure. For the next theorem, we say that a Borel map u : X → Y has Jacobian Ju, if there exists a Borel function J u: X → [0, ∞] for which

 E J udμ =  Y #(u−1(y) ∩ E)dν(y) (5.13)

holds for every Borel set E ⊂ X. The Jacobian Ju, if it exists, is unique up to sets of

μ-measure zero.

Theorem 5.16 Assume u is proper, nonconstant and has a locally integrable Jacobian J u

which satisfies

guQ ≤ K Ju (5.14)

μ-almost everywhere for some K ≥ 1. Then the following properties hold.

(1) Yuis a metric measure space and has the Q-Sobolev-to-Lipschitz property.

(2) J u is the locally integrable Jacobian of Puand # Pu−1(y) = 1 for νu-almost every y∈ Yu.

Proof Note that   Pu−1E J udμ =  Y #(u−1(y) ∩ Pu−1E)dν(y) ≥  Y

#(u−1(y) ∩ E)dν(y) = νu(E) for every Borel set E ⊂ Yu. Thusνu is a locally finite measure, and the estimate above implies that Pu has a locally integrable Jacobian J Pu ≤ Ju. For any Borel set E ⊂ X we have

(26)

 E

J Pudμ = 

Yu

#(Pu−1(y) ∩ E)dνu(y) =  Y ⎛ ⎝  y∈u−1(z) #(Pu−1(y) ∩ E)⎠ dν(z) =  Y #(u−1(z) ∩ E)dν(z) =  E J udμ,

which yields J Pu = Ju almost everywhere. Since Pu is monotone and satisfies (5.13) we have that # Pu−1(y) = 1 for νu-almost every y∈ Yu.

By Lemma5.15to see that Yuis a metric measure space it remains to show that balls in

Yuhave positive measure. The idea of of proof is borrowed from [33, Lemma 6.11]. Assume

B= B(z, r) is a ball in Yusuch thatνu(B) = 0. Then we would have J Pu= Ju= 0 almost everywhere on U:= Pu−1(B). By (5.14) and Lemma5.15gP

u = gu = 0 almost everywhere on U . Thus Puis locally constant on U . But U is connected and hence Puis constant.

To see that U is connected let x ∈ U satisfy Pu(x) = z. Then, for y ∈ U and δ > 0 arbitrary, there is a(u, Q)-regular curve γ ∈ (B(x, δ), B(y, δ)) such that (u ◦ γ ) < r. Forδ small enough by Proposition5.7the image ofγ is contained in U. As this holds for all small enoughδ, the points x and y must lie in the same connected component of the open set U . Since y was arbitrary, U must be connected.

We have deduced that B = Pu(U) consists only of the single point z. So either Yu is disconnected or consists of a single point. The former is impossible because X is connected and the later by the assumption that u is nonconstant.

Before showing the Sobolev-to-Lipschitz property we first claim

ModQ ≤ K ModQPu() (5.15)

for every curve family in X. Indeed, suppose ρ ∈ LQ(Yu) is admissible for Pu(), and setρ1 = ρ ◦ Pu. Let0a curve family with ModQ0 = 0 for which γ is (Pu, Q)- and

(u, Q)-regular, and guis an upper gradient of Puand u alongγ , whenever γ /∈ 0. For any

γ ∈  \ 0we have 1≤   Pu◦γ ρ =  1 0 ρ1(γ (t))|(Pu◦ γ )t|dt ≤  1 0 ρ1(γ (t))gu(γ (t))|γt|dt, i.e.ρ1guis admissible for \ 0. We obtain

ModQ = ModQ( \ 0) ≤  X ρQ 1 g Q udμ ≤ K  X ρQ◦ P uJ udμ = K  Yu ρQdν u.

The last equality follows, since Lemma5.15implies Pu∗(Ju μ) = νu. Taking infimum over admissibleρ yields (5.15).

We prove that Yu has the Q-Sobolev-to-Lipschitz property. Suppose f ∈ N1,Q(Yu) sat-isfies gf ≤ 1. There is a curve family 0in Yu with ModQ(0) = 0 such that

| f (γ (1)) − f (γ (0))| ≤ u(γ ) wheneverγ /∈ 0. By (5.15) we have

Références

Documents relatifs

La fonction m

La fonction m

Les deux courbes indiquent les limites basses et hautes de l'évolution du poids d'un enfant : sa courbe de poids doit a priori se situer entre ces deux courbesh.

Les deux courbes indiquent les limites basses et hautes de l'évolution du poids d'un enfant : sa courbe de poids doit a priori se situer entre ces deux courbesh.

Exprimer, en fonction de x , le volume que peuvent occuper les bonbons dans la boîte.. Dans la pratique, x est compris entre 0,5

1 Pour chacune des questions suivantes, trois réponses sont proposées, une seule est exacte.. Pour chaque question, entoure la

1 Pour chacune des questions suivantes, trois réponses sont proposées, une seule est exacte.. Pour chaque question, entoure la

Quelles sont les valeurs minimales et maximales d'abscisses et d'ordonnées à entrer dans la fenêtre d'affichage de la